Next Article in Journal
Novel Approach for Integrated Environmental Management Systems
Previous Article in Journal
Cellulose Carriers from Spent Coffee Grounds for Lipase Immobilization and Evaluation of Biocatalyst Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Studies on the Valorization of Aluminum Production Residues into Bituminous Materials at Different Scales: A Review

Department of Applied Sciences, University of Quebec at Chicoutimi (UQAC), Saguenay, QC G7H 2B1, Canada
*
Authors to whom correspondence should be addressed.
Sustainability 2025, 17(21), 9634; https://doi.org/10.3390/su17219634
Submission received: 28 August 2025 / Revised: 1 October 2025 / Accepted: 14 October 2025 / Published: 29 October 2025
(This article belongs to the Section Waste and Recycling)

Abstract

To conserve natural resources and reduce waste generation, the effective valorization of industrial waste and byproducts in engineering applications is becoming increasingly important. Among these materials, aluminum production residues (APRs) offer a promising and sustainable solution for road pavement applications. Unlike previous reviews, this paper uniquely examines recent research on the use of various APRs in bituminous materials across multiple scales, with particular attention to their roles as additives and fillers. The APRs examined included red mud (RM), aluminum dross (AD), and spent pot lining (SPL) residues, as well as secondary aluminum waste (SAW). These materials have been employed as additives in asphalt binders (microscale), as fillers in asphalt mastics (mesoscale), and as additives or fillers in asphalt mixtures (macroscale). Overall, this review indicates that adopting appropriate treatment approaches for APRs as asphalt modifiers can enhance their dispersion, thermal stability, rheological behavior, and leaching performance. In particular, the use of RM has been shown to improve thermal stability, tensile strength, intermediate-temperature cracking resistance, and rutting resistance, largely due to the increased stiffness it imparts to asphalt mastic and mixture phases. However, there is no clear consensus among researchers regarding other properties, as performance outcomes depend strongly on multiple factors, particularly the physicochemical characteristics of the RM, filler–binder ratios, testing methods, and reference filler types. Other APRs—such as AD, SPL, and SAW—have also shown beneficial effects on the performance of asphalt mixtures. There is still limited research on the influence of APRs physicochemical variability on asphalt–filler interactions and the performance of bituminous materials. For the safe and large-scale adoption of APRs, it is essential to establish standardized characterization procedures, testing methods, and application guidelines while considering diverse climatic conditions. Comprehensive assessments of cost and environmental impacts should also be incorporated to support informed decision-making by engineers and industrial stakeholders.

1. Introduction

Today, rapid urbanization and industrialization generate enormous amounts of waste, leading to significant disposal challenges and harming the environment. Concurrently, the construction industry—particularly road pavement construction—consumes vast quantities of virgin raw materials, contributing to resource depletion and high energy consumption [1]. In response to these challenges, the valorization of waste materials in road pavements has emerged as a highly promising sustainable approach [2]. This strategy not only addresses the issue of waste disposal but also offers an opportunity to conserve natural resources, reduce energy consumption, and potentially lower construction costs. To this end, researchers have incorporated different waste materials—such as reclaimed asphalt pavement (RAP) [3], recycled concrete aggregate (RCA) [4], and crumb rubber (CR) [5]—as replacements for natural aggregates, as well as waste fillers—such as fly ash (FA) [6], ground granulated blast-furnace slag (GGBFS) [7], rice husk ash [8], brick dust (BD) [9], and borogypsum [10].
Driven by global industrial development and urbanization, increased demand for aluminum has led to a rise in the generation of byproducts [11]. Aluminum production residue (APR)—which includes byproducts from bauxite mining and refining operations as well as waste from post-consumer aluminum—has gained significant attention among researchers for valorization in construction materials, particularly road pavements, due to its high volume and potential usage as a sustainable resource [12].
Due to the growing interest in the use of waste materials in asphalt pavements, this research presents a comprehensive review of the incorporation of APRs into bituminous mixtures at different scales, including their application as asphalt binder modifiers, filler replacements, and mixture additives. The review aims to enhance understanding of their interactions with asphalt, highlight standard test methods for characterizing bituminous materials containing APRs, and assess the feasibility of using these residues in large-scale applications as a practical and sustainable approach. To this end, relevant literature was collected from three major bibliographic databases—Web of Science, Scopus, and Google Scholar—through systematic searches based on predefined keywords, as well as inclusion and exclusion criteria. The extracted data covered the type and dosage of APRs, various scales of bituminous materials, test methods, performance-related parameters, and key findings. Both qualitative and quantitative syntheses, with consideration of potential moderators, were employed to analyze the extracted information. This review begins by listing and comparing the physicochemical and mineralogical properties of commonly used APRs in bituminous materials. At the asphalt binder level (microscale), the role of APRs as additives is evaluated based on dispersion stability, rheological behavior, thermal properties, and chemical/nanoscale interactions. In the mastic phase (mesoscale), physicochemical interactions between APRs as fillers are investigated, along with their rheological and thermal effects. At larger scales (macroscale), the influence of APRs—either as additives or partial filler replacements—is assessed in terms of their effects on the volumetric properties, fatigue performance, low-temperature cracking resistance, permanent deformation, and moisture susceptibility of asphalt mixtures. Considering the potential toxicity of APRs, the environmental implications of incorporating these materials into asphalt mixtures are also critically examined.
The different types of APRs studied in this review can be classified as follows.

1.1. Red Mud

Red mud (RM), also known as bauxite residue, is the main waste material generated during aluminum production. It is produced during the alumina (Al2O3) refining process, which involves the reaction of bauxite with sodium hydroxide (NaOH) under high temperature and pressure [13]. The type of RM produced depends on the aluminum production method used, including the Bayer process (Figure 1a), the sintering process (Figure 1b), or a combination of both (Bayer-sintering) [14,15]. The Bayer process accounts for over 95% of global alumina production from bauxite, while the remainder comes from other methods [16,17]. For every ton of alumina produced, roughly 2–3 tons of bauxite are consumed, leading to the generation of approximately 0.4–2.0 tons of RM [18]. Approximately 150 million tons of RM are generated globally each year [19,20]. This high rate of generation has led to both environmental and economic challenges. Specifically, the high alkalinity and presence of heavy metals and trace amounts of radioactive elements in RM make its disposal a significant environmental challenge. These factors can lead to serious issues, such as groundwater contamination, often triggered by heavy rainfall or failures in containment structures [21,22]. Additionally, RM’s fine-grained particles are easily dispersed by wind, contributing to air pollution. [23]. As a sustainable solution, significant efforts have been made to identify suitable approaches for the safe and effective utilization of RM, which has been widely employed in the production of various construction materials, including lightweight aggregates [24], bricks [25], low-density foamed products [26], and traditional ceramics [27]. It has also been used as a partial replacement for cement in mortar and concrete formulations [28,29,30]. Moreover, due to its high alkalinity, RM has been valorized in the development of geopolymer composites [31,32,33,34]. In recent years, its utilization has expanded to the pavement sector, where it has been incorporated into road subgrade, base, and surface layer materials [35,36,37,38,39].

1.2. Aluminum Dross

Aluminum dross (AD) (Figure 1c), a byproduct of the aluminum melting process, occurs mainly in two forms, distinguished by their metal and salt content and by their origin in either primary or secondary aluminum production. White dross, which contains little to no salt, is produced during primary aluminum smelting, while black dross, which is rich in salts, is generated during secondary aluminum production [40,41,42]. Compared to black dross, the separation of white dross requires more energy and water and generates a greater volume of waste [43]. Notably, black dross typically contains less metallic aluminum than white dross [44]. Globally, it is estimated that up to 4 million tons of white dross and over 1 million tons of black dross are generated annually, with approximately 95% of this waste disposed of in landfills [45].
AD contains a certain volume fraction of toxic materials, and landfilling these hazardous substances is environmentally unsustainable. If not properly treated, toxic elements can leach into the surroundings and pose significant environmental risks [46]. One practical approach to mitigating the disposal issue is to utilize AD in building applications. This strategy not only reduces the environmental hazards associated with AD but also helps address natural resource scarcity. AD is primarily used in the production of ceramics [47,48] and bricks [49,50], as a partial replacement for cement in cementitious composites [51,52,53,54], and in soil stabilization [55].

1.3. Spent Pot Lining Residues

Spent pot lining (SPL) is hazardous waste generated from the internal lining of aluminum electrolytic pots. It comprises carbon-based materials and refractory bricks and is typically divided into two primary components: (a) spent cathode carbon block (SCCB) residues (Figure 1d) [56], which constitute the carbon-rich lower section of the lining, and (b) spent refractory materials (SRM) (Figure 1e), which include oxide-based bricks and insulation from the side and upper layers [57,58]. SPL is classified as a dangerous waste, primarily due to its high fluoride content of up to 150 kg per ton [59]. If not properly treated, SPL can lead to fluoride accumulation in the environment, causing lesions in plants and animals and posing serious health risks to humans [60,61].
Given the vast scale of the aluminum industry, the annual generation of SPL is substantial, with approximately 25 kg produced per tonne of primary aluminum, amounting to an estimated 1.6 million tonnes annually [62]. In recent years, researchers have increasingly focused on the safe and sustainable valorization of these residues in construction materials [63].

1.4. Secondary Aluminum Waste

Secondary aluminum waste (SAW), which refers to postconsumer aluminum materials that are recycled and reused, reportedly accounts for approximately one-third of total aluminum consumption [64,65]. Valorizing SAW yields monetary and economic benefits, reduces energy consumption, and helps increase landfill capacity. Aluminum shavings (Figure 1f) or chips, as part of SAW, are small, thin fragments of aluminum generated as byproducts during machining processes—such as cutting, drilling, or milling—in various industries. Their physicochemical characteristics—including size, density, shape, and chemical composition—vary depending on the specific machining process used. Due to the technical and environmental benefits offered by aluminum shavings, they have been widely employed in construction materials, primarily as additive fibers in concrete [66,67], cement and geopolymer mortars [68,69,70], and hot mix asphalt (HMA) [71].
This review not only examines various types of APRs but also emphasizes their incorporation into bituminous materials across different scales, distinguishing it from other studies. Different scales of bituminous materials are shown in Figure 2 [72]. The smallest scale is the asphalt binder phase, whose constituents are classified based on their polarity into four fractions—saturates, aromatics, resins, and asphaltenes—commonly referred to as SARA fractions [73]. Resins, saturates, and aromatics—collectively known as maltenes—represent the soluble molecular components and continuous phase of the asphalt binder [74]. The mastic phase incorporates fillers smaller than a specific size threshold (63 μm as per European standards [75], or 75 μm in North America [76]) and contains no air voids. It is regarded as an intermediate phase with relatively consistent properties and is commonly used to assess performance on larger scales. The fine aggregate matrix (FAM) phase bridges the gap between mastic and full asphalt mix; it consists of asphalt binder, fine aggregates typically smaller than 2.00–4.75 mm, filler, and air voids [77]. At the full scale—known as the asphalt mixture phase, which constitutes the road structure—coarse aggregates are incorporated, with the nominal maximum aggregate size defining the characteristic scale.

2. Physicochemical and Mineralogical Properties of Various Aluminum Production Residues

To ensure the effective valorization of APRs, it is essential to understand their physicochemical properties and mineralogical composition, as these characteristics significantly influence the performance of bituminous materials at various phases.

2.1. Chemical Composition of APRs

Figure 3a shows the chemical composition of RM from different processing methods—Bayer, Sintering, and a combination—as determined by X-ray fluorescence (XRF) analysis. As shown, RM is mainly composed of Al2O3, Fe2O3, and SiO2, as well as smaller amounts of CaO and Na2O. Among these, Al2O3 is the most abundant component in RM derived from the Bayer process but is less prevalent in the Sintering method. This variance depends on the bauxite source and extraction efficiency. The Fe2O3 content, which is generally higher in Bayer RM, can enhance the electrical and mechanical properties of asphalt mixtures [78,79]. Fe2O3 can also improve bonding with asphalt by providing active oxide surfaces that promote physical adsorption, hydrogen bonding, and interactions with polar groups in the asphalt binder [80]. Additionally, SiO2 is more abundant in RM from the Sintering and combination methods. The presence of SiO2 has been reported to improve interfacial bonding between the asphalt binder and aggregates, thereby increasing the mixture’s stiffness and mechanical strength [81]. A high content of CaO—an alkaline compound found in aggregates and fillers—enhances their bonding with the weakly acidic asphalt binder. This improvement is primarily attributed to the electrostatic potential generated at the interface between asphalt molecules and materials rich in CaO [82,83]. Compared to RM derived from Sintering or combination methods, Bayer RM generally contains lower levels of CaO. Studies have reported that blending Bayer RM with CaO-rich fillers, such as limestone, can improve the bonding performance of asphalt mixtures [84]. Furthermore, the Na2O content in some Bayer process residues can reach up to 12%, imparting high pH and alkalinity to RM and raising environmental concerns, particularly regarding soil and water contamination [85].
As shown in Figure 3b, AD contains exceptionally high levels of Al2O3 (60–80%), formed through the oxidation of molten aluminum. The remaining oxides—such as SiO2, Fe2O3, CaO, and MgO—originate from impurities, alloying agents, or furnace materials [86]. Additionally, the presence of SO2 suggests possible interactions with sulfur-bearing compounds during processing or storage [87]. Regarding SPL residue materials (Table 1), the main component in SRM is CaO, followed by SiO2 and Al2O3 [88]. In the case of SCCB, the predominant component is carbon (C)—determined by the loss on ignition method—followed by NaF and Fe2O3 [89].
Figure 3. (a) XRF analysis results of RM (the ternary phase diagram of the Fe2O3–SiO2–Al2O3 system was obtained from normalized values) [30,84,90,91,92,93,94,95,96,97,98,99,100] and (b) chemical composition of AD [101,102,103] (Flash arrows are positioned to facilitate the identification and interpretation of values).
Figure 3. (a) XRF analysis results of RM (the ternary phase diagram of the Fe2O3–SiO2–Al2O3 system was obtained from normalized values) [30,84,90,91,92,93,94,95,96,97,98,99,100] and (b) chemical composition of AD [101,102,103] (Flash arrows are positioned to facilitate the identification and interpretation of values).
Sustainability 17 09634 g003

2.2. Mineralogical Properties

The X-ray diffraction (XRD) patterns of each residue are presented in Figure 4. As can be seen, Sintering RM contains higher amounts of calcite (CaCO3), whereas Bayer RM is richer in goethite (FeO(OH)) and hematite (Fe2O3). These findings are consistent with XRF analysis results [104,105]. It has been reported that AD mainly contains the crystalline compounds Al2O3, aluminum nitride (AlN), magnesium aluminate spinel (MgAl2O4), sodium hexafluoroaluminate (Na3AlF6), and hydrated aluminum oxide (Al2O3H2O), as shown in Figure 4c, which are the key mineral phases present after the extraction of metallic aluminum [106]. Figure 4d shows that CaCO3 is the most prominent phase, as indicated by its highest diffraction peak, followed by calcium fluoride (CaF2) and SiO2. [88]. The XRD pattern of SCCB (Figure 4e) shows that the dominant crystalline phases present are carbon (C) and sodium fluoride (NaF) [89].

2.3. Microstructural Assessment

Scanning electron microscope (SEM) images of various residues are presented in Figure 5. RM exhibits small, rounded, agglomerated particles, mostly less than 10 μm in size, with an irregular granulometric distribution and rough texture. These characteristics may lead to a higher SSA and, consequently, increased asphalt binder absorption [107]. Similar observations have been reported in several other studies [108,109]. As illustrated in Figure 5b, the SEM image of AD reveals smooth, angular particles, along with plate-like fragments and whisker-like surface features [110]. In Figure 5c, the SRM particles exhibit a distinct lamellar morphology, with numerous surface grooves and fine pores. This structural complexity contributes to a higher SSA, which can enhance the absorption of free asphalt during mixing. As a result, it facilitates the formation of an asphalt mastic with stronger cohesive forces, thereby improving high-temperature stability [88]. Figure 5d shows that the surface of SCCB is relatively smooth, with well-defined edges and an absence of visible pore structures. Nonetheless, surface impurities are present, which may be attributed to residual sodium fluoride in the SCCB matrix [89].
Figure 5. SEM images of (a) RM [107], (b) AD [110], (c) SRM [88] (reprinted with permission), and (d) SCCB [89] (reprinted with permission).
Figure 5. SEM images of (a) RM [107], (b) AD [110], (c) SRM [88] (reprinted with permission), and (d) SCCB [89] (reprinted with permission).
Sustainability 17 09634 g005
Table 2 presents a comparative summary of the key physicochemical properties of RM and AD, based on values reported in the literature. The specific gravity of RM varies considerably, ranging from 2.60 to 3.50 g/cm3, with this variability attributed to differences in chemical composition resulting from various processing methods and the type of bauxite used. The higher specific gravity of RM compared with AD is primarily due to its greater content of high-density oxides, such as Fe2O3 and TiO2.
The average particle size (d50) of RM is typically fine, ranging from 2.75 μm to 7.20 μm, whereas AD consists of coarser particles. Regarding specific surface area (SSA), RM exhibits a wide range, from 8.3 m2/g to 35.4 m2/g, attributed to variations in mineralogy, particle size, and porosity. In contrast, AD had a significantly lower SSA, indicating reduced surface interaction potential. The higher SSA of RM provides a larger contact area with the asphalt binder, which enhances adhesion and improves surface interaction with the binder [111].
The Rigden void (RV) values of both residues indicate relatively porous structures, which depend on particle shape, size, distribution, and surface texture. A higher RV value increases the volume of asphalt that can be retained within the voids, thereby contributing to the greater stiffness of the asphalt mastic phase.
In terms of hydrophilicity, both RM and AD have been reported to exhibit moderate to high hydrophilic behavior. However, it has been reported that a higher hydrophilicity coefficient contributes to improved adhesion between the mineral filler and asphalt [112]. Furthermore, RM is characterized by strong alkalinity, with pH values ranging from 9.0 to 12.5, primarily due to the incorporation of alkaline reagents during the alumina extraction process [113]. Since the asphalt binder is slightly acidic, it tends to form stronger bonds with alkaline materials, offering enhanced resistance to stripping. In other words, a pH above 7 imparts a basic character to the residue surface, rendering it electrically negative, which can positively influence asphalt–filler adhesion.
Table 2. Physical and chemical properties of RM and AD.
Table 2. Physical and chemical properties of RM and AD.
APR TypeSpecific
Gravity (g/cm3)
Average Particle
Size, d50 (μm)
Specific Surface
Area, SSA (m2/g)
Rigden Void, RV (vol%)Hydrophilic
Coefficient, η
pH Range
RM3.10 [108,114], 3.50 [115], 2.60 [94], 2.97 [90],
2.80 [116], 2.55, 2.85 [91]
4.50 [108], 3.33 [114], 2.75 [115], 6.50 [94], 7.20 [116]12.10 [117], 8.30 [116], 30.99, 35.41 [91]44.05 [114], 47.90 [94], 44.80 [116]0.85 [108], 0.47 [94], 0.70 [96],
0.62 [97]
9.98 [108], 9.18 [114], 10.25 [109], 11.00 [115], 12.40 [116], 9.90, 10.00 [91]
AD2.90 [110], 2.79 [118]24.96 [110], 15.62 [119]1.03 [110], 1.26 [119]45.70 [110]0.67 [110]-

3. Valorization of APRs in Asphalt Binders (Microscale): Additive Applications

The use of waste additives in asphalt binders offers a sustainable approach to enhancing pavement performance while minimizing environmental impacts. Incorporating these materials—such as waste plastics, CR, and bio-based residues—has been shown to improve key binder properties, including viscosity, fatigue resistance, resistance to permanent deformation, and low-temperature cracking, ultimately contributing to longer-lasting pavements [120,121,122,123]. Among the various waste-derived additives, APRs have emerged as a promising option for asphalt modification due to their distinct physicochemical characteristics.

3.1. Use of RM as an Additive

Tao et al. (2017) [124] investigated the application of RM as an additive in asphalt binders through experimental studies and molecular simulations. They showed that RM reduced the penetration index and ductility while increasing the softening point, temperature sensitivity, and thermal stability of asphalt binders. From a chemical and microstructural perspective, RM’s large SSA enables it to adhere more uniformly to the asphalt binder surface. Furthermore, RM absorbs the oily components of asphalt, creating stronger molecular interactions that improve bonding. Consequently, the asphalt structure becomes thicker, adhesion between asphalt and mineral components is strengthened, and the modified asphalt binder demonstrates better stability and overall technical performance. As illustrated in Figure 6a, the RM-modified asphalt binder exhibits the formation of lumps and micro-blocks, which is attributed to π–π interactions and the accumulation of RM within the asphaltene- and oil-containing components.
In another study, Fu et al. (2020) [125] investigated the microstructural, chemical, and thermal characteristics of RM-modified asphalt under freeze–thaw (F–T) cycles. Their findings revealed that the addition of RM further decreased the heat absorption of the asphalt binder and that F–T cycles significantly reduced the heat absorption capacity of the modified asphalt binder. Microstructural analysis using atomic force microscopy showed that the number of bee structures observed noticeably decreased with the incorporation of RM. However, the remaining bee structures exhibited increased height and length. Furthermore, the F–T cycles had a significant impact on the morphology and uniformity of these structures. Molecular dynamics simulations demonstrated that asphaltenes adsorb onto RM surfaces, with stronger interactions observed with Al2O3 than with Fe2O3 across various temperatures. Therefore, increasing the Al2O3 content or reducing the Fe2O3 content in RM can enhance asphaltene adsorption. More specifically, Al2O3 surfaces exhibit strong adsorption energies with polar molecules in asphalt, promoting stable interfacial bonding and enhancing binder durability. The polar oxide surface facilitates hydrogen bonding and Lewis acid–base interactions with heteroatoms such as nitrogen and oxygen present in asphalt molecules, thereby improving adhesion and cohesion at the molecular level.
Figure 6b shows an SEM image of aluminum tailing slurry (ATS), categorized as RM, whose coarse surface texture and fine particle size facilitate the formation of a mechanical interlocking structure with the asphalt matrix, thereby enhancing interfacial bonding [126]. It has been shown that using ATS as an asphalt binder additive can reduce the quantity of bee structures and surface roughness, which may contribute to a decrease in the self-healing capacity of asphalt binders. Moreover, increasing the dosage of ATS in asphalt binders improves thermal stability, complex modulus, and rutting resistance. However, when the dosage exceeds 9%, the storage stability of the asphalt binder decreases significantly [127].
Figure 6. SEM image of (a) RM–modified asphalt binder [124] and (b) ATS [127].
Figure 6. SEM image of (a) RM–modified asphalt binder [124] and (b) ATS [127].
Sustainability 17 09634 g006
Compatibility between the additives and the asphalt binder is a critical factor that should be considered at the outset. For example, incorporating hydrophilic powders—such as RM—into a hydrophobic matrix, such as asphalt binder, can lead to interfacial defects and poor compatibility. Surface treatment of RM can enhance its efficiency as an asphalt modifier [128]. To improve interfacial interaction, Xiao et al. (2023) [129] investigated the use of organically modified RM as an additive in asphalt binders. RM was first treated with oxalic acid and a coupling agent and then analyzed using various microstructural techniques. Their results showed that surface modification successfully transformed the RM from hydrophilic to hydrophobic, yielding a more uniform and smaller particle size distribution. This modification improved both dispersion and interaction within the asphalt binder matrix. The enhanced performance can be attributed to the long organic chains of the modified RM, which intertwine with asphalt molecules, while physical adsorption strengthens the intermolecular forces between the RM particles and the asphalt binder. As a result, the modified asphalt binder exhibited increased resistance to high-temperature deformation and low-temperature cracking. Furthermore, the fine particle size of the modified RM contributed to the improved thermal stability and aging resistance of the asphalt binder. Han et al. (2025) [130] proposed an innovative approach to surface treatment of RM to improve the performance characteristics of RM-modified asphalt binders. They utilized dopamine, hypothesizing that its dual-functional groups could enhance compatibility between asphalt and RM. The catechol group forms stable bonds with metal oxides (e.g., Fe2O3, Al2O3) on the RM surface, improving dispersion in the asphalt matrix, while the aromatic ring of dopamine interacts with polycyclic aromatic hydrocarbons in asphalt binder via π–π stacking interactions [131,132]. Under mild alkaline conditions, dopamine undergoes self-polymerization into polydopamine, forming a cross-linked network that limits RM aggregation and strengthens cohesion in the asphalt matrix through physical anchoring. Therefore, they used solid waste RM as the base material, which was modified through dopamine self-polymerization and esterification reactions to synthesize a biomimetic composite with a high surface area inspired by the Siberian cocklebur (GA-CDA-RM). The preparation steps and mechanisms are illustrated in Figure 7a. Fluorescence microscopy images showed a more uniform dispersion of GA-CDA-RM compared to untreated RM in the asphalt binder. This treatment also enhanced the binder’s high-temperature rheological properties, rutting resistance, and low-temperature crack resistance while significantly reducing the leaching of heavy metal ions from the RM.
In another study [133], co-hydrothermal carbonization of RM with rice straw was used to produce modified forms of RM (c-HTC). Notably, H-HTC was developed through phosphoric acid–assisted hydrothermal carbonization (Figure 7b). The modified RMs exhibited significant detoxification and enhanced physicochemical properties, such as abundant surface functional groups, a clustered structure with tighter particle bonding (for c-HTC), and a unique flower-cluster structure (for H-HTC), as shown in the SEM images and Fourier transform infrared spectroscopy (FTIR) analysis (Figure 8a–c and Figure 9). The incorporation of this modified RM into asphalt binders substantially improved their high-temperature performance by increasing rutting resistance and enhancing thermal stability, as validated through molecular simulations. In another relevant study, Li et al. (2025) [134] developed a promising method using metallothioneins from plant oil residues to treat RM (Figure 7c). This treatment not only reduces harmful components in RM but also enhances the complex modulus, rutting resistance, and antiaging properties of asphalt binders. From a microstructural perspective (Figure 8d–g), RM modified with canola, soybean, and peanut oil residues forms small spherical structures, likely representing heavy metal ion–protein complexes. This confirms effective heavy metal immobilization and reduced environmental hazards. Additionally, the presence of these organic–inorganic microspheres improves the compatibility between RM and asphalt, enhancing the material’s elasticity and resistance to deformation.

3.2. Utilization of Other APRs as Modifiers

In addition to RM, SPL and SCCB residues have also been used to modify asphalt binders. Fan et al. (2024) [135] developed a novel thermoplastic polyester elastomer–spent refractory composite (TPEE–SRM) as an asphalt binder modifier, synthesized through an emulsion evaporation process, as illustrated in Figure 10a. This composite features a dual-layer structure consisting of a protective layer and a functional layer. Epoxy resins (EPs) incorporated into the protective layer formed a dense film that effectively sequesters fluoride. The organic modification of SRM altered its surface properties, enabling a better and more uniform dispersion of TPEE–SRM within the oily components of asphalt. The TPEE–SRM-modified asphalt binder exhibited enhanced complex modulus and rutting resistance compared to conventional styrene–butadiene–styrene (SBS)-modified asphalt. Furthermore, the TPEE–SRM/SBS composite-modified asphalt demonstrated superior fatigue resistance. In a study with a similar objective, researchers synthesized a bayberry-like hexamethyldisilazane and tannic acid–modified SRM (HMDS–TA–SRM) via a synchronous sedimentation–encapsulation reaction (Figure 10b). This modification enhanced the hydrophobicity of SRM and improved mechanical interlocking with the asphalt binder due to its distinctive nano-rough surface and mesoporous structure. The study demonstrated that asphalt binders modified with HMDS–TA–SRM exhibit excellent long-term performance and significantly reduce fluoride leaching to environmentally safe levels, with a maximum leaching rate of just 0.486% after 1 year. Additionally, the surface-encapsulated SRM showed outstanding fatigue resistance, attributed to the good organic compatibility of tannic acid and the oil-adsorptive surface created through further oil modification. It has also been reported that HMDS–TA–SRM SBS-modified asphalt exhibits significantly improved resistance to permanent deformation and enhanced long-term stability compared to pristine SBS-modified asphalt [136].
In another study, the authors addressed the hazardous nature of SCCB residues—known to contain toxic cyanides and fluorides—by developing a gel-crosslinked composite material, PSB, for asphalt modification. The treatment process involved using hydrogen peroxide to eliminate cyanides and calcium chloride to precipitate fluorides, followed by encapsulation with a polyvinyl alcohol–polyvinyl pyrrolidone (PVA–PVP) gel (Figure 10c). The resulting PSB effectively reduced cyanide concentrations to below environmental thresholds and enabled a slower, controlled release of fluorides, ensuring environmental safety. When incorporated into asphalt, particularly at a 1% PSB/SBS dosage, the modifier enhanced high-temperature rutting resistance, thermal stability, and elastic recovery when compared to conventional SBS-modified asphalt [89].

4. Studies on Asphalt Mastic (Mesoscale) and Binder–Filler Interactions

Several studies have explored the valorization of APRs as full or partial replacements for conventional fillers in asphalt mastics. To evaluate the effectiveness of these residues, a wide range of experimental tests have been conducted in accordance with standards, with the most used listed in Table 3.

4.1. Conventional Physical Properties

Conventional physical properties of APR asphalt mastics—including penetration, softening point, and ductility—have been evaluated in various studies. Yao et al. [84] replaced limestone (LS) with RM at 0%, 25%, 50%, 75%, and 100% levels, considering different filler-to-binder (F–B) ratios. They found that the penetration of an asphalt mastic containing RM was lower than that of an LS asphalt mastic. Of the five mixtures, the mastic with 100% RM had the most significant effect on reducing binder penetration. Additionally, the incorporation of RM increased the softening point of the asphalt binder. Zhang et al. (2018) [91] demonstrated that Sintering RM had a significant effect on the softening point, while Zhang et al. (2020) [105] reported that incorporating white mud (WM) into an RM asphalt mastic reduced the softening point and enhanced both penetration and ductility.

4.2. Rotational Viscosity

Many researchers have reported that incorporating RM as a filler in asphalt mastics increases rotational viscosity [84,90,92] and that adding hydrated lime (HL) further enhances this effect [97]. In contrast, introducing WM as a modifier into RM asphalt mastics slightly decreases viscosity [105].

4.3. Linear Viscoelastic Properties

Using temperature and frequency sweep tests, several studies have reported an increase in the complex modulus (G*) with the incorporation of RM into asphalt mastics [84,91,97,105]. This trend is consistent with the observed increase in the dynamic modulus of asphalt mixtures containing RM [149]. Similarly, an increase in G* was reported when conventional filler was replaced with SRM in an asphalt mastic [88].

4.4. Permanent Deformation (Rutting) Resistance

To evaluate the permanent deformation resistance of APR asphalt mastics, two key parameters—rutting index (G*/sin δ) and nonrecoverable creep compliance (Jnr)—were obtained through temperature sweep and multiple stress creep recovery (MSCR) tests, respectively. Studies have reported that incorporating RM as a filler leads to an increase in G*/sin δ [84,91,97] and a reduction in Jnr [92], indicating improved resistance to permanent deformation. Similar improvements have also been observed in asphalt mastics containing SRM [88].

4.5. Fatigue Resistance

Researchers have used G*sin(δ)—obtained from frequency and temperature sweep tests—as an index to evaluate the fatigue resistance of asphalt binders at intermediate temperatures. Lower values of G*sin(δ) typically indicate greater flexibility and improved fatigue resistance. Studies have reported that RM samples demonstrated either a higher [97] or a lower [105] G*sin(δ). Based on parameters obtained from linear amplitude sweep (LAS) tests, RM asphalt mastics exhibited the highest [95], lowest [150], or no clear trend [151] in terms of fatigue resistance performance compared to other samples.

4.6. Resistance to Low-Temperature Cracking

According to the bending beam rheometer (BBR) test, an asphalt mastic with a lower creep stiffness modulus (S) and a higher creep slope (m) indicates better resistance to cracking. Yao et al. (2020) [84] reported that incorporating RM increases the S value and decreases the m value, making mastics more susceptible to low-temperature cracking. However, Zhang et al. (2018) [90] did not reach a definitive conclusion on this effect. Furthermore, Zhang et al. (2020) [105], using a direct tension test, found that RM had a detrimental impact on low-temperature cracking resistance, but this effect could be improved by adding HL and WM.

4.7. Aging Resistance

Based on the carbonyl index obtained from FTIR analysis, the incorporation of diatomite (DT) and RM improved the aging resistance of an asphalt mastic by approximately 37.7% and 44.7%, respectively, compared to an LS asphalt mastic. In contrast, the use of FA resulted in reduced aging resistance. These findings have been supported by aging indices derived from both low- and high-temperature rheological properties [152].

4.8. Bonding, Wettability, and Moisture Susceptibility

Lima et al. (2020) [116] conducted a comparative study on the adhesive and wetting properties of asphalt mastics containing conventional fillers (LS and dolomite (DL)) and alternatives (RM and FA). They found that mastics with 40% RM achieved strong adhesion by passing the boiling water stripping test without chemical additives. Wettability tests showed that DL increased hydrophilicity (lower contact angles) while RM increased hydrophobicity (higher contact angles) (Figure 11a). However, this finding contradicts the observation reported in Section 2, indicating that RM exhibits moderate to high hydrophilic behavior.
In another study, Ou et al. (2024) [95] investigated the wettability and adhesive properties of asphalt mastics containing LS, RM, steel slag, and GGBFS. The contact angle of the mastic surface was measured using various probe liquids. The results showed that the RM asphalt mastic had the lowest surface free energy (SFE) and contact angle (Figure 11b) among the mixes. This may be due to the high SiO2 content, which affects the electrochemical reactions in the asphalt mastic, resulting in a low SFE. Additionally, the presence of Na2O can increase the risk of water damage at the bonding interface, as the chemical bonds formed between Na2O and asphalt are easily dissolved upon exposure to water [153]. This aligns with the findings of Zhang et al. [149], who measured the surface energy parameters of various fillers using an inverse gas chromatography test. They reported that LS exhibited the highest surface energy values while RM exhibited the lowest.
Many researchers have evaluated the moisture susceptibility of asphalt mastics containing RM; however, their findings remain inconclusive. Zhang et al. (2018) [90] reported that asphalt mastics incorporating RM filler exhibited better resistance to moisture-induced damage than LS fillers at the same F–B ratio, based on the binder bond strength (BBS) test. In contrast, Wang et al. (2021) [152] observed opposite results. A separate study [97] investigated the effect of Sintering RM on the moisture sensitivity of aggregate–mastic bonding using an automatic adhesion tester. Specimens were conditioned in a water bath for 1 and 7 days to simulate moisture damage. Bond strength tests showed that substituting LS with RM significantly decreased moisture resistance; however, this reduction was alleviated by incorporating an appropriate amount of HL. Furthermore, Zhang et al. (2020) [105] found that using WM as a modifier could counteract the negative effects of RM on bonding strength at the aggregate–mastic interface after moisture conditioning. In addition, Zhang et al. (2019) [92] demonstrated that the pull-off tensile strength measured with the BBS test—along with the self-healing capacity of asphalt mastics—depends strongly on the type of RM, the F–B ratio, and the testing conditions (dry or wet).
Table 4 summarizes studies on the asphalt mastic phase using different types of fillers and F–B ratios, along with key physicochemical, microstructural, and rheological findings.

5. Valorization of APRs in Asphalt Mixtures (Macroscale)

As asphalt mixture is the final product used in road surfacing, its mechanical performance and durability are key criteria for evaluating the feasibility of using APRs. Numerous studies have investigated the effects of these residues on the performance characteristics of asphalt mixtures through various experimental tests, with the most commonly used summarized in Table 5.

5.1. Studies on RM Hot Mix Asphalt Samples

5.1.1. Volumetric Properties

Choudhary et al. (2019) [114] reported that the RM mix exhibited a lower optimum binder content (OBC) than the SD mix. Moreover, unlike the SD mix, the OBC values for the RM mix did not vary significantly with an increase in filler content. This behavior was attributed to the finer particle size of RM compared to SD, which enabled it to act as a binder extender and enhanced the workability of the mixture. However, other studies have reported contrasting results, suggesting that the porous structure and high SSA of the RM lead to greater asphalt binder absorption [107,166].

5.1.2. Strength and Flow Measurement

Marshall stability reflects the resistance of asphalt mixes to deformation under load and is influenced by the viscosity of the asphalt mastic, which in turn depends on the interaction between asphalt and filler. Therefore, the type and amount of filler significantly affect the stability. Researchers have reported RM mixes exhibiting higher Marshall stability than OPC [108] and SD [107] mixes, which may be attributed to the higher bearing capacity of RM.
The flow value indicates the plastic behavior of asphalt mixtures and is defined as the total strain or deformation occurring in a specimen between the application of zero and the maximum load during the Marshall test. A higher flow value reflects greater plasticity, while a lower value suggests brittleness [167]. An RM asphalt mix has been shown to exhibit a higher flow value than an SD mix, indicating increased plasticity [107].

5.1.3. Permanent Deformation Resistance

The Marshall quotient (MQ), calculated as the ratio of Marshall stability (kN) to flow (mm), indicates the resistance of an asphalt mix to permanent deformation and shear stress [168]. A higher MQ signifies greater stiffness and enhanced load distribution capacity. Choudhary et al. (2022) [107] found that RM mixes exhibited higher MQ values, likely due to the lower asphalt film thickness of RM and its more integrated structure compared to SD mixes [169]. Lima et al. (2017) [170] conducted a permanent deformation test following French standard NF P 98-253-1. They found that replacing stone powder with RM as a filler improved the deformation resistance of asphalt mixtures, supporting the findings of other research [109]. Due to its lower SSA compared to stone powder, RM can reduce the asphalt binder content required, thereby limiting the thermal susceptibility of the mixture. The superior resistance to permanent deformation of RM mixes compared to conventional mixes has been demonstrated using the Hamburg wheel tracking (HWT) test [90,149]. Furthermore, Tian et al. (2021) [94] reported that porous asphalt mixtures incorporating RM and DT exhibited enhanced rutting resistance—based on dynamic stability and creep slope measurements—compared to those containing LS and FA (Figure 12). This improved performance was attributed to the strong adsorption capacity and hardening effect of these porous fillers on the asphalt binder.

5.1.4. Moisture Susceptibility

Choudhary et al. (2018) [108] revealed that fillers rich in insoluble calcium-based compounds provide superior resistance to moisture damage. Mixes containing carbide lime (CL), ordinary Portland cement (OPC), and limestone slurry dust exhibited higher retained Marshall stability (RMS) values compared to an RM mix, with the CL mix showing the highest RMS value. This was attributed to the inclusion of portlandite and calcite in the CL, which impart antistripping properties. However, Choudhary et al. (2022) [107] reported that the highly alkaline nature of RM, along with its content of adhesion-promoting minerals—such as calcite—contributed to its satisfactory moisture resistance compared to SD mixes [171]. These findings are consistent with those reported by Lima et al. (2020) [109].
Enhanced moisture resistance was also reported by Zhang et al. (2018) [90] and Shanchun et al. [172] (2022) when RM was used either as a replacement for LS filler or as an asphalt binder modifier in the mixture [96]. In contrast, Tian et al. (2021) [94] found that RM porous asphalt mixtures exhibited a lower tensile strength ratio (TSR) than LS mixtures. This reduction was attributed to the presence of Na2O in the RM, which tends to swell upon water exposure and increase moisture sensitivity—a finding consistent with mastic phase evaluations by Ou et al. (2024) [95]. Zhang et al. [149] also demonstrated that replacing LS filler with RM improved rutting resistance under dry conditions but negatively impacted its performance under moisture conditioning. To mitigate this, the addition of 10% HL or OPC significantly enhanced the moisture resistance of RM mixtures. This improvement was attributed to the presence of Ca(OH)2 in HL, which formed stable bonds with carboxylic acids and sulfoxides in the binder. Similarly, cement compounds such as 3CaO·SiO2, 2CaO·SiO2, and 3CaO·Al2O3 contributed to stronger adhesion with the asphalt binder, thereby improving resistance to moisture-induced damage.

5.1.5. Tensile Strength and Cracking Resistance at Intermediate Temperatures

Recent studies have reported that RM mixes exhibited superior indirect tensile strength compared to reference samples [107,108,114]. This improvement can be attributed to the fine particle size of RM, which promotes uniform distribution within a mixture and contributes to the formation of an integrated structure in the bituminous matrix, thereby enhancing stiffness [169]. Tian et al. (2021) [94] performed the semicircular bend (SCB) test and calculated quantitative indicators, including peak load, from load–displacement curves; fracture energy; slope at inflection point; and flexibility index (Figure 13). They concluded that asphalt mixtures containing DT and RM exhibited notable cracking resistance due to the anchoring and hardening effects provided by the fillers. However, RM—in comparison to DT—not only significantly hardened the asphalt binder but also demonstrated a certain toughening effect. The anchoring and toughening effects result from the interaction between RM—primarily composed of interparticle pores—and the crosslinked network structure of the incorporated SBS modifier.

5.1.6. Adhesion and Bonding

Passive adhesion is defined as the ability of bitumen to adhere to an aggregate surface under external influences, such as water and traffic [173]. Mixes with superior passive adhesion generally exhibit higher asphalt coverage in boiling water tests. Researchers have reported that RM exhibits good passive adhesion compared to most other waste fillers, including BD, CL, copper tailings, limestone slurry dust, rice straw ash, and glass powder [108]. This performance was attributed to the stiffening effect of RM’s fine particles on the asphalt binder. However, RM showed lower passive adhesion than mixes containing OPC [108], SD [107,166], and Kota dust (KS) [166]. The superior adhesion observed in the KS and SD mixes compared to RM was attributed to the higher content of calcium-based dolomite and calcite, respectively, in their composition [166].

5.1.7. Raveling Resistance

Based on the Cantabro abrasion test, researchers reported that RM mixes outperformed LS [90,94], DT, and FA [94] but underperformed compared to SD mixes [107].

5.1.8. Aging Resistance

Choudhary et al. [107] assessed the aging resistance of RM asphalt mixes using the mean Marshall stability ratio, defined as the ratio of the average Marshall stability after aging (oven-conditioned at 85 °C for 5 days) to that before aging. Their results indicated that RM mixes exhibited lower aging resistance than conventional mixes, which may be attributed to the lower asphalt film thickness in RM mixes [174]. In contrast, Zhang et al. (2018) [90] reported that porous asphalt mixtures containing RM with an F–B ratio of 0.9 demonstrated better aging resistance than those with LS. These findings align with those of Tian et al. (2021) [94], who observed that the inclusion of DT and RM improved the raveling resistance of aged porous asphalt mixtures by 39.6% and 43.2%, respectively, compared to an LS mix. In summary, Table 6 lists important studies on the valorization of RM in HMA.

5.1.9. Valorization of RM in Alternative Asphalt Mixture Types

Few studies have examined the use of RM in other types of asphalt mixtures—such as cold-mix asphalt (CMA) or warm-mix asphalt (WMA)—or its practical application in asphalt plants. Phan et al. (2024) [115] evaluated the performance of a CMA mixture containing RAP, RM as a filler, and recycled waste glass as a fine aggregate. They conducted a comprehensive series of tests, including workability and compactability, IDEAL-CT (Figure 14a), HWT, and dynamic modulus tests. In addition, a 400 m pedestrian road trial section was constructed to assess CMA performance under real outdoor conditions (Figure 14b). However, the study primarily focused on the effects of recycled glass, with the role of RM as a filler not specifically highlighted.

5.1.10. Environmental Aspects of RM Valorization

Due to the presence of certain heavy metals in its composition [175,176] and residual sodium hydroxide, especially from the Bayer process, RM exhibits strong alkaline properties and poses potential environmental concerns. Consequently, many researchers classify RM as hazardous waste; however, some studies have categorized it as noninert waste rather than strictly hazardous [177,178]. Lima et al. (2018) [178] analyzed water from moisture damage tests on RM asphalt mixes using a photocolorimeter to detect potential hazardous leaching. Although the results indicated only minimal leaching of RM, the study recommended more precise and in-depth evaluations using specialized equipment to better assess the risk of water contamination. In another study, Lima et al. (2021) [179] evaluated the environmental potential of asphalt mixtures using alternative fillers (RM and FA) alongside reference fillers (LS and DL) through a cradle-to-gate life cycle assessment. The analysis covered stages from raw material acquisition to asphalt plant production. Three scenarios were modeled: (1) residues with treatment, (2) residues without treatment, and (3) landfilling. The treatment process involved washing, drying, and crushing the RM, while FA underwent Sintering. Environmental indicators included global warming potential, acidification, and cumulative energy demand (CED). The results indicated that residue treatment significantly influenced environmental impacts and played a key role in assessing the feasibility of using RM and FA in asphalt pavements.
Li et al. (2021) [180] investigated the runoff purification performance of porous asphalt with various fillers—including LS, coal fly ash, DT, and RM—using a self-developed rainfall simulation device (Figure 15) introduced by Liu et al. (2021) [181]. The authors reported that the purification mechanism of the asphalt mix involved both physical adsorption and chemical degradation, including chemical precipitation and ion exchange. Alkaline precipitation from fillers and aggregates in porous asphalt leads to the formation of heavy metal precipitates. In terms of ion exchange, a strong interaction occurs between Na+ and K+ leached from red mud and NH4+ in runoff, enhancing the removal of total nitrogen. Furthermore, RM contains abundant porous structures and a large SSA, significantly improving the micropore characteristics of porous asphalt. Its rougher surface and larger surface area enhance interactions with pollutants, strengthen physical adsorption, and ultimately improve the overall purification performance of asphalt mixes.
In another study, Rashidian et al. (2025) [117] conducted a toxicity characteristic leaching procedure test to assess the level of heavy metal pollution leached from RM microsurfacing asphalt mixtures. The concentrations of heavy metals—As, Pb, Cd, Cr, Ni, Cu, Hg, and Zn—in the RM mixes were well below the allowable limits. This could be attributed to the alkaline nature of RM, which enhances its bonding with the asphalt binder and reduces leaching. Additionally, the stabilizing properties of the asphalt binder helped mitigate the release of harmful metals—a finding also supported by previous research [169,182].

5.2. Utilization of Aluminum Dross in HMA

Some researchers have evaluated the feasibility of incorporating AD as a filler in asphalt mixtures [101,119,183]. Ulga et al. (2024) [184] investigated the moisture susceptibility of HMA samples in which LS filler was partially replaced with AD at replacement levels of 10%, 20%, and 30%. The optimum replacement level was identified as 20%, at which point two key moisture resistance indicators—TSR and the index of retained strength—were found to be 13.42% and 8.73% higher, respectively, than those of the reference mixes. The same results were obtained when incorporating different dosages of RCA and AD in HMA [102].
Ergezer et al. (2025) [118] examined the impact of microwave heating on self-healing and ice melting in asphalt mixtures containing steel fiber (SF) and AD, based on the hypothesis that AD, like other iron-containing materials, absorbs microwaves and is thermally conductive [185]. AD powders were used as a substitute for LS filler at different content levels, and SF was added in different proportions by the weight of the mix. The results showed that increasing the AD and SF content significantly improved the average surface temperature and ice-melting speed by 112.86% and 73.6%, respectively, compared to the reference mix. Indirect tensile tests (IDTs) conducted before and after damage indicated that higher AD and SF contents enhanced healing after 60 s of microwave heating. However, extending the heating to 120 s with a high SF content reduced the healing effectiveness. Soos et al. (2017) [110] performed a dynamic modulus test to evaluate the linear viscoelastic behavior of asphalt mixtures containing LS filler, with AD used as a 50% and 100% replacement for LS. As shown in Figure 16, analysis of the Cole–Cole diagram and the master curve—developed using the sigmoidal model—confirmed the stiffening effect of AD. The results indicate that the increase in stiffness was more pronounced at high temperatures than at low temperatures.

5.3. Incorporation of Aluminum Shavings in HMA

Researchers have incorporated aluminum shavings into HMA due to their favorable reinforcing mechanisms [71]. Serin et al. (2021) [186] evaluated the cracking resistance of mixes containing aluminum chips derived from shavings and iron powders at proportions of 0.5%, 1.0%, and 1.5% by weight of the total mixture using the SCB test (Figure 17a). They reported that incorporating 0.5% aluminum chips improved the fracture energy, peak load, cracking resistance index, and flexibility index compared to the reference mixture. This was supported by SEM images (Figure 17b), which showed no evidence of segregation or loss of adhesion between asphalt binder and aluminum chips; instead, clear wetting and strong bonding were observed. Similarly, Al-Obaidi and Abed (2023) [187] found that adding waste aluminum scrap powder enhanced the cracking resistance of asphalt mixtures. In another study, Atakan et al. (2021) [188] investigated the microwave healing potential of porous asphalt mixes containing waste aluminum and steel shavings, produced with different aggregate types (basalt and limestone), using an IDT at low temperatures. The results showed that aluminum shavings exhibited better healing performance than steel shavings in both aggregate-type groups, consistent with the findings of Pamulapati et al. (2017) [189]. The authors attributed this to aluminum’s lower density, which promotes more uniform dispersion in the mixture and increases microwave absorption efficiency compared to steel [188]. Martinez et al. (2025) [190] showed that incorporating 0.1% aluminum shavings into HMA significantly improved tensile strength, fatigue life, and moisture damage resistance (indicated by higher TSR), reduced mass loss due to abrasion, and maintained a resilient modulus and stiffness comparable to the reference mix, despite a slight decrease in resistance to permanent deformation. However, higher aluminum shaving contents (≥0.5%) resulted in performance deterioration.

6. Challenges and Recommendations for Future Work

Although APRs show promising potential in bituminous materials, several challenges must be addressed before their large-scale adoption in pavement engineering. The physicochemical and mineralogical characteristics of APRs—particularly RM—vary significantly depending on their source, processing method, and treatment history. This variability affects their compatibility with asphalt binders and aggregates, as well as their mechanical and environmental performance. Consequently, inconsistent performance can lead to misinterpretation and confusion among engineers, making it difficult to standardize mixed designs and establish effective quality-control measures.
The hydrophilic nature and strong tendency for particle agglomeration in most APRs—especially RM, which is often reported as a self-agglomerating material—can result in poor dispersion within hydrophobic asphalt matrices, thereby reducing adhesion and workability. Such agglomeration also poses practical challenges during the construction stage, including storage, sieving, and mixing operations. Surface modification techniques—such as organic treatments and coupling agents—show promise for improving compatibility but require further optimization for large-scale field applications.
While researchers generally agree that APRs can enhance certain mixture properties, such as stiffness, stability, and rutting resistance, their effects on other critical indicators—including fatigue performance, low-temperature cracking resistance, and moisture susceptibility—remain inconsistent across studies. Moreover, research on their applicability in cold regions is lacking; in particular, low-temperature cracking behavior, freeze–thaw cycle durability, and fatigue performance at the asphalt mixture scale have not been sufficiently explored.
From an environmental perspective, although several studies have examined the leaching behavior and environmental impact of RM in asphalt mixtures—with promising results—little to no research has addressed these aspects for other APR types. Comprehensive evaluations of leaching potential, long-term environmental stability, and ecotoxicity are necessary to ensure safe and sustainable implementation.
Based on the challenges discussed above, the following recommendations for future research are proposed:
  • Investigate the linkage between the physicochemical properties of APRs and the rheological and mechanical performance of bituminous materials using advanced analytical tools, such as machine learning. This will help in developing robust characterization protocols and classification systems for APRs and standardizing their application in mix design, construction, and quality assessment.
  • Explore advanced surface modification techniques to enhance APR–asphalt binder compatibility, focusing on scalable, eco-friendly approaches that are feasible for large-scale applications.
  • Study the incorporation of APRs at the FAM scale as a more precise approach compared to the mastic phase to simulate the performance of large-scale asphalt mixtures.
  • Assess the feasibility of incorporating APRs into various asphalt mixture types—such as CMA and WMA—to achieve a more sustainable approach.
  • Focus on the use of APRs in road pavements for cold regions by designing comprehensive experimental testing protocols tailored to these environmental conditions.
  • Conduct life cycle assessments and life cycle cost analyses to evaluate the environmental and economic viability of APRs in large-scale industrial applications.

7. Conclusions

The valorization of APRs—including RM, AD, and SPL residues and SAW in bituminous materials—offers a promising pathway to more sustainable pavement engineering. This review presents a comprehensive investigation of various APRs from different stages of aluminum production in bituminous materials, aiming to clarify their reinforcing mechanisms and assess their effects on the performance of asphalt mixtures for large-scale applications. Treating and using different APRs as asphalt modifiers has shown potential; however, scaling up these applications for field use may present economic challenges. Across meso- and macroscale studies, RM has demonstrated the ability to improve thermal stability, tensile strength, cracking resistance at intermediate temperatures, and rutting resistance. Nonetheless, inconsistent results have been reported for other properties, such as fatigue behavior, low-temperature cracking, aging resistance, adhesion, and moisture susceptibility. These variations stem from differences in the physicochemical properties of RM as well as in F–B ratios, reference filler types, mix designs, and other parameters. Despite promising environmental assessment results for RM in asphalt mixtures, mainly due to its stabilizing effect on asphalt binders, which helps mitigate the release of harmful metals, environmental evaluations of other APRs at the macro scale remain largely unexplored. To ensure the safe and widespread adoption of APRs, it is essential to develop standardized characterization methods and application protocols, alongside a deeper understanding of the fundamental interactions between APRs and asphalt binders across various scales.

Funding

This research was funded by Rio Tinto and Mitacs under grant number IT45927.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

The authors thank Rio Tinto and Mitacs for their support of this study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hasan, A.; Hasan, U.; Whyte, A.; Al Jassmi, H. Lifecycle analysis of recycled asphalt pavements: Case study scenario analyses of an urban highway section. CivilEng 2022, 3, 242–262. [Google Scholar] [CrossRef]
  2. Ahmed, W.; Lu, G.; Ng, S.T.; Liu, G. Innovative valorization of solid waste materials for production of sustainable low-carbon pavement: A systematic review and scientometric analysis. Case Stud. Constr. Mater. 2025, 22, e04541. [Google Scholar] [CrossRef]
  3. Tsakoumaki, M.; Plati, C. A critical overview of using reclaimed asphalt pavement (RAP) in road pavement construction. Infrastructures 2024, 9, 128. [Google Scholar] [CrossRef]
  4. Fanijo, E.O.; Kolawole, J.T.; Babafemi, A.J.; Liu, J. A comprehensive review on the use of recycled concrete aggregate for pavement construction: Properties, performance, and sustainability. Clean. Mater. 2023, 9, 100199. [Google Scholar] [CrossRef]
  5. Zeleke, Y.S.; Tsegaye, D.B. Investigating the Effect of Crumb Rubber Using as an Aggregate in the Hot Mix Asphalt Mixture Properties. In Sustainable Development Research in Green Infrastructure, Water Resources, Manufacturing, and Process Engineering: Advancements of Science and Technology; Springer Nature: Cham, Switzerland, 2025; pp. 429–450. [Google Scholar]
  6. Russo, F.; Veropalumbo, R.; Pontoni, L.; Oreto, C.; Biancardo, S.A.; Viscione, N.; Pirozzi, F.; Race, M. Sustainable asphalt mastics made up recycling waste as filler. J. Environ. Manag. 2022, 301, 113826. [Google Scholar] [CrossRef] [PubMed]
  7. Nassar, A.I.; Mohammed, M.K.; Thom, N.; Parry, T. Characterisation of high-performance cold bitumen emulsion mixtures for surface courses. Int. J. Pavement Eng. 2018, 19, 509–518. [Google Scholar] [CrossRef]
  8. Al-Hdabi, A. Laboratory investigation on the properties of asphalt concrete mixture with Rice Husk Ash as filler. Constr. Build. Mater. 2016, 126, 544–551. [Google Scholar] [CrossRef]
  9. Arabani, M.; Tahami, S.A.; Taghipoor, M. Laboratory investigation of hot mix asphalt containing waste materials. Road Mater. Pavement Des. 2017, 18, 713–729. [Google Scholar] [CrossRef]
  10. Kütük-Sert, T.; Kütük, S. Physical and Marshall properties of borogypsum used as filler aggregate in asphalt concrete. J. Mater. Civil. Eng. 2013, 25, 266–273. [Google Scholar] [CrossRef]
  11. Nie, Q.; Hu, W.; Ai, T.; Huang, B.; Shu, X.; He, Q. Strength properties of geopolymers derived from original and desulfurized red mud cured at ambient temperature. Constr. Build. Mater. 2016, 125, 905–911. [Google Scholar] [CrossRef]
  12. Kumar, B.R.; Ramakrishna, G. Performance evaluation of red mud as a construction material—A review. Mater. Today Proc. 2023, 84, 263–278. [Google Scholar]
  13. Brunori, C.; Cremisini, C.; Massanisso, P.; Pinto, V.; Torricelli, L. Reuse of a treated red mud bauxite waste: Studies on environmental compatibility. J. Hazard. Mater. 2005, 117, 55–63. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, J.; Yao, Z.; Wang, K.; Wang, F.; Jiang, H.; Liang, M.; Wei, J.; Airey, G. Sustainable utilization of bauxite residue (Red. Mud) as a road material in pavements: A critical review. Constr. Build. Mater. 2021, 270, 121419. [Google Scholar] [CrossRef]
  15. Harmaji, A.; Jafari, R.; Simard, G. Valorization of residue from aluminum industries: A review. Materials 2024, 17, 5152. [Google Scholar] [CrossRef]
  16. Reddy, P.S.; Reddy, N.G.; Serjun, V.Z.; Mohanty, B.; Das, S.K.; Reddy, K.R.; Rao, B.H. Properties and assessment of applications of red mud (bauxite residue): Current status and research needs. Waste Biomass Valorization 2021, 12, 1185–1217. [Google Scholar] [CrossRef]
  17. Sun, C.; Chen, J.; Tian, K.; Peng, D.; Liao, X.; Wu, X. Geochemical characteristics and toxic elements in alumina refining wastes and leachates from management facilities. Int. J. Environ. Res. Public Health 2019, 16, 1297. [Google Scholar] [CrossRef]
  18. Bánvölgyi, G.; Huan, T.M.; Director, F. De-watering, disposal and utilization of red mud: State of the art and emerging technologies. TRAVAUX 2010, 35, 431–443. [Google Scholar]
  19. Evans, K. The history, challenges, and new developments in the management and use of bauxite residue. J. Sustain. Metall. 2016, 2, 316–331. [Google Scholar] [CrossRef]
  20. Wang, L.; Hu, G.; Lyu, F.; Yue, T.; Tang, H.; Han, H.; Yang, Y.; Liu, R.; Sun, W. Application of red mud in wastewater treatment. Minerals 2019, 9, 281. [Google Scholar] [CrossRef]
  21. Hua, Y.; Heal, K.V.; Friesl-Hanl, W. The use of red mud as an immobiliser for metal/metalloid-contaminated soil: A review. J. Hazard. Mater. 2017, 325, 17–30. [Google Scholar] [CrossRef]
  22. Mukiza, E.; Zhang, L.; Liu, X.; Zhang, N. Utilization of red mud in road base and subgrade materials: A review. Resour. Conserv. Recycl. 2019, 141, 187–199. [Google Scholar] [CrossRef]
  23. Deelwal, K.; Dharavath, K.; Kulshreshtha, M. Evaluation of characteristic properties of red mud for possible use as a geotechnical material in civil construction. Int. J. Adv. Eng. Technol. 2014, 7, 1053. [Google Scholar]
  24. Molineux, C.J.; Newport, D.J.; Ayati, B.; Wang, C.; Connop, S.P.; Green, J.E. Bauxite residue (red mud) as a pulverised fuel ash substitute in the manufacture of lightweight aggregate. J. Clean. Prod. 2016, 112, 401–408. [Google Scholar] [CrossRef]
  25. Xu, Y.T.; Yang, B.; Liu, X.M.; Gao, S.; Li, D.S.; Mukiza, E.; Li, H.J. Investigation of the medium calcium based non-burnt brick made by red mud and fly ash: Durability and hydration characteristics. Int. J. Miner. Metall. Mater. 2019, 26, 983–991. [Google Scholar] [CrossRef]
  26. Liu, T.; Li, X.; Guan, L.; Liu, P.; Wu, T.; Li, Z.; Lu, A. Low-cost and environment-friendly ceramic foams made from lead–zinc mine tailings and red mud: Foaming mechanism, physical, mechanical and chemical properties. Ceram. Int. 2016, 42, 1733–1739. [Google Scholar] [CrossRef]
  27. Sglavo, V.M.; Maurina, S.; Conci, A.; Salviati, A.; Carturan, G.; Cocco, G. Bauxite ‘red mud’in the ceramic industry. Part. 2: Production of clay-based ceramics. J. Eur. Ceram. Soc. 2000, 20, 245–252. [Google Scholar] [CrossRef]
  28. Tang, W.C.; Wang, Z.; Liu, Y.; Cui, H.Z. Influence of red mud on fresh and hardened properties of self-compacting concrete. Constr. Build. Mater. 2018, 178, 288–300. [Google Scholar] [CrossRef]
  29. Ghalehnovi, M.; Roshan, N.; Hakak, E.; Shamsabadi, E.A.; De Brito, J. Effect of red mud (bauxite residue) as cement replacement on the properties of self-compacting concrete incorporating various fillers. J. Clean. Prod. 2019, 240, 118213. [Google Scholar] [CrossRef]
  30. Singh, V.; Bano, S.; Chauhan, V.B.; Pal, P.; Kumar, A.; Srivastava, J.B. Red mud as a sustainable road construction material: An. experimental investigation. Constr. Build. Mater. 2024, 411, 134549. [Google Scholar] [CrossRef]
  31. Mudgal, M.; Singh, A.; Chouhan, R.K.; Acharya, A.; Srivastava, A.K. Fly. ash red mud geopolymer with improved mechanical strength. Clean. Eng. Technol. 2021, 4, 100215. [Google Scholar] [CrossRef]
  32. Li, Y.; Min, X.; Ke, Y.; Liu, D.; Tang, C. Preparation of red mud-based geopolymer materials from MSWI fly ash and red mud by mechanical activation. Waste Manag. 2019, 83, 202–208. [Google Scholar] [CrossRef]
  33. Sun, Z.; Tang, Q.; Xakalashe, B.S.; Fan, X.; Gan, M.; Chen, X.; Ji, Z.; Huang, X.; Friedrich, B. Mechanical and environmental characteristics of red mud geopolymers. Constr. Build. Mater. 2022, 321, 125564. [Google Scholar] [CrossRef]
  34. Harmaji, A.; Jafari, R.; Simard, G. Durable bauxite-residue geopolymers: Enhancing scaling resistance with fly ash and waste glass powder. Constr. Build. Mater. 2025, 491, 142661. [Google Scholar] [CrossRef]
  35. Sahoo, S.; Mohanty, C. Construction of road sub-base by using industrial waste. Int. J. Eng. Res. Dev. 2016, 3, 20–26. [Google Scholar]
  36. Singh, K.; Pandey, R.K.; Mishra, C.S.; Rai, A.K.; Bind, Y.K. Analysis on utilization of cement kiln dust stabilized red mud for road construction. Int. J. Civil. Eng. Technol. 2014, 5, 56–61. [Google Scholar]
  37. Satayanarayana, P.V.V.; Naidu, G.; Adiseshu, S.; Rao, C. Characterization of lime stabilized red mud mix for feasibility in road construction. Int. J. Eng. Res. Dev. 2012, 3, 20–26. [Google Scholar]
  38. Liu, X.M.; Tang, B.W.; Yin, H.F.; Emile, M. Durability and environmental performance of Bayer. red mud-coal gangue-based road base material. Chin. J. Eng. 2018, 40, 438–445. [Google Scholar]
  39. Wong, T.L.X.; Hasan, M.R.M.; Peng, L.C. Recent development, utilization, treatment and performance of solid wastes additives in asphaltic concrete worldwide: A review. J. Traffic Transp. Eng. 2022, 9, 693–724. [Google Scholar]
  40. Wilson, S. Residues from Aluminium Dross Recycling in Cement; Citeseer: University Park, PA, USA, 2007. [Google Scholar]
  41. Hwang, J. Recovery of metals from aluminum dross and saltcake. J. Miner. Mater. Charact. Eng. 2006, 5, 47. [Google Scholar] [CrossRef]
  42. Muñoz Perez, S.P.; Garcia Chumacero, J.M.; Charca Mamani, S.; Villena Zapata, L.I. Influence of the secondary aluminum chip on the physical and mechanical properties of concrete. Innov. Infrastruct. Solut. 2023, 8, 45. [Google Scholar]
  43. Mahinroosta, M.; Allahverdi, A. Hazardous aluminum dross characterization and recycling strategies: A critical review. J. Environ. Manag. 2018, 223, 452–468. [Google Scholar] [CrossRef]
  44. Udvardi, B.; Géber, R.; Kocserha, I. Examination of the utilization of aluminum dross in road construction. In Proceedings of the IOP Conference Series: Materials Science and Engineering, Hiroshima, Japan, 26–28 March 2019; IOP Publishing: Bristol, UK. [Google Scholar]
  45. Busari, A.; Joseph, F.; Ajayi, S.; Alayande, T.; Nwachukwu, J.; Agbama, D. Index properties of aluminum dross modified pavement geo-material. J. Phys. Conf. Series. 2019, 1378, 022102. [Google Scholar] [CrossRef]
  46. Lou, B.; Shen, H.; Liu, B.; Liu, J.; Zhang, S. Recycling secondary aluminum dross to make building materials: A review. Constr. Build. Mater. 2023, 409, 133989. [Google Scholar] [CrossRef]
  47. Ewais, E.M.; Besisa, N.H. Tailoring of magnesium aluminum titanate based ceramics from aluminum dross. Mater. Des. 2018, 141, 110–119. [Google Scholar] [CrossRef]
  48. Liu, Y.; Yang, J.; Shen, H.; Zhang, J.; Li, W.; Zhang, X.; Liu, J.; Liu, B.; Zhang, S. Synthesis of porous glass ceramics with hierarchical and interconnected pores from secondary aluminum dross and waste glass. Ceram. Int. 2022, 48, 34364–34373. [Google Scholar] [CrossRef]
  49. Ni, H.; Wu, W.; Lv, S.; Wang, X.; Tang, W. Formulation of non-fired bricks made from secondary aluminum ash. Coatings 2021, 12, 2. [Google Scholar] [CrossRef]
  50. Zhang, Z.J.; Li, S.Z.; Wang, L.L.; Li, M.K.; Huang, K.P.; Wu, W.; Liu, J.; Yi, X.M. Preparation of Aluminum Dross Microporous Bricks and the Pore Formation Mechanisms. JOM 2023, 75, 4701–4713. [Google Scholar] [CrossRef]
  51. Parra-Molina, D.; Rojas-Manzano, M.A.; Gómez-Gómez, A.; Muñoz-Vélez, M.F.; Maury-Ramírez, A. Mechanical Performance of Mortars with Partial. Replacement of Cement by Aluminum Dross: Inactivation and Particle Size. Sustainability 2023, 15, 14224. [Google Scholar] [CrossRef]
  52. Elinwa, A.U.; Mbadike, E. The use of aluminum waste for concrete production. J. Asian Archit. Build. Eng. 2011, 10, 217–220. [Google Scholar] [CrossRef]
  53. Javali, S.; Chandrashekar, A.R.; Naganna, S.R.; Manu, D.S.; Hiremath, P.; Preethi, H.G.; Vinod Kumar, N. Eco-concrete for sustainability: Utilizing aluminium dross and iron slag as partial replacement materials. Clean Technol. Environ. Policy 2017, 19, 2291–2304. [Google Scholar] [CrossRef]
  54. Reddy, M.S.; Neeraja, D. Mechanical and durability aspects of concrete incorporating secondary aluminium slag. Resour.-Effic. Technol. 2016, 2, 225–232. [Google Scholar]
  55. Busari, A.A.; Akinwumi, I.I.; Awoyera, P.O.; Olofinnade, O.M.; Tenebe, T.I.; Nwanchukwu, J.C. Stabilization effect of aluminum dross on tropical lateritic soil. Int. J. Eng. Res. Afr. 2018, 39, 86–96. [Google Scholar] [CrossRef]
  56. Zhou, L.; Yao, Z.; Sun, K.; Tian, Z.; Li, J.; Zhong, Q. Methodological Review of Methods and Technology for Utilization of Spent Carbon Cathode in Aluminum Electrolysis. Energies 2024, 17, 4866. [Google Scholar] [CrossRef]
  57. Tu, H.; Li, X.; Muhammad, Y.; Jin, X.; Lin, H.; Pei, R.; Zhao, Z.; Li, J. Feasibility of reuse of modified electrolytic aluminium spent refractory material in asphalt and assessment of its environmental stability. J. Clean. Prod. 2023, 389, 136072. [Google Scholar] [CrossRef]
  58. Birry, L.; Poirier, S. The LCL&L process: A sustainable solution for the treatment and recycling of spent pot lining. Light Met. 2020, 2020, 1237–1242. [Google Scholar]
  59. Zhao, X.; Ma, L. Hazardous waste treatment for spent pot liner. In Proceedings of the IOP Conference Series: Earth and Environmental Science, Kolaka, Indonesia, 7–9 February 2018; IOP Publishing: Bristol, UK, 2018. [Google Scholar]
  60. Choubisa, S.L.; Choubisa, D. Status of industrial fluoride pollution and its diverse adverse health effects in man and domestic animals in India. Environ. Sci. Pollut. Res. 2016, 23, 7244–7254. [Google Scholar] [CrossRef]
  61. Lisbona, D.F.; Somerfield, C.; Steel, K.M. Leaching of spent pot-lining with aluminum anodizing wastewaters: Fluoride extraction and thermodynamic modeling of aqueous speciation. Ind. Eng. Chem. Res. 2012, 51, 8366–8377. [Google Scholar] [CrossRef]
  62. Nunez, P. Sustainable spent pot line management guidance. Light Met. 2020, 2020, 1225–1230. [Google Scholar]
  63. Camara, N.; Ouellet-Plamondon, C. Scenarios to Valorize Treated Spent Pot Lining in Cement and Concrete. In Proceedings of the Canadian Society of Civil Engineering Annual Conference, Niagara Falls, ON, Canada, 26–29 May 2021; Springer: Singapore, 2021. [Google Scholar]
  64. Wallace, G. Production of secondary aluminium. Fundam. Alum. Metallurgy 2011, 70–82. [Google Scholar]
  65. Aleksić, J.; Vargas, D.B. Aluminium Demand Will Rise 40% by 2030, Here’s How to Make It Sustainable. World Economic Forum. Available online: https://www.weforum.org/agenda/2023/11/aluminium-demand-how-to-make-it-sustainable/ (accessed on 28 November 2023).
  66. Elsayed, M.; Abd-Allah, S.R.; Said, M.; El-Azim, A.A. Structural performance of recycled coarse aggregate concrete beams containing waste glass powder and waste aluminum fibers. Case Stud. Constr. Mater. 2023, 18, e01751. [Google Scholar] [CrossRef]
  67. Ofuyatan, O.M.; Ivoke, A.A.; Olowofoyeku, A.M.; Adesina, A.; Oluwafemi, J. Effect of waste aluminium shavings on the bond characteristics of laterized concrete. Adv. Mater. Res. 2019, 8, 25–36. [Google Scholar]
  68. Yusuf, M.O. Performance of aluminium shaving waste and silica fume blended mortar. Mag. Civil Eng. 2023, 122, 100–112. [Google Scholar]
  69. Filazi, A.; Yilmazel, R.; Pul, M. The Effect of Material Ratios on Technical Properties in Fly Ash and Blast Furnace Slag Based Aluminum Chip Reinforced Geopolymer Mortars. Int. Conf. Pioneer Innov. Stud. 2023, 1, 344–350. [Google Scholar] [CrossRef]
  70. Gulmez, N. Roles of aluminium shavings and calcite on engineering properties of cement-based composites. J. Clean. Prod. 2020, 277, 124104. [Google Scholar] [CrossRef]
  71. Ahmed, L.S.A.; Mahmood, L.O.T. Influence of Mineral fibers Properties on the performance of Hot Mix Asphalt for the Surface Layer of Pavement. Department of Civil Engineering, Faculty of Engineering, Koya University, Erbil, Iraq. Int. J. Eng. Trends Technol. 2015, 26, 145–151. [Google Scholar] [CrossRef]
  72. Lackner, R.; Spiegl, M.; Blab, R.; Eberhardsteiner, J. Is low-temperature creep of asphalt mastic independent of filler shape and mineralogy?—Arguments from multiscale analysis. J. Mater. Civil. Eng. 2005, 17, 485–491. [Google Scholar] [CrossRef]
  73. Salehfard, R.; Behbahani, H.; Dalmazzo, D.; Santagata, E. Effect of colloidal instability on the rheological and fatigue properties of asphalt binders. Constr. Build. Mater. 2021, 281, 122563. [Google Scholar] [CrossRef]
  74. Salehfard, R.; Yeganeh, S.; Dalmazzo, D.; Underwood, B.S.; Santagata, E. Linking chemical structure to the linear and nonlinear properties of asphalt binders. Transp. Res. Rec. 2024, 2678, 1483–1495. [Google Scholar] [CrossRef]
  75. EN 12697–2 (2015); Bituminous Mixtures-Test Methods-Part 2: Determination of Particle Size Distribution. European Committee for Standardization: Brussels, Belgium, 2015.
  76. Wang, D.; Cannone Falchetto, A.; Zhang, F.; Riccardi, C.; Sun, Y. Fabrication and characterization of multiphase bituminous materials for cold region pavements. J. Infrastruct. Preserv. Resil. 2023, 4, 22. [Google Scholar] [CrossRef]
  77. Behbahani, H.; Salehfard, R. A review of studies on asphalt fine aggregate matrix. Arab. J. Sci. Eng. 2021, 46, 10289–10312. [Google Scholar] [CrossRef]
  78. Arabani, M.; Mirabdolazimi, S. Experimental investigation of the fatigue behaviour of asphalt concrete mixtures containing waste iron powder. Mater. Sci. Eng. A 2011, 528, 3866–3870. [Google Scholar] [CrossRef]
  79. Ullah, S.; Yang, C.; Cao, L.; Wang, P.; Chai, Q.; Li, Y.; Wang, L.; Dong, Z.; Lushinga, N.; Zhang, B. Material design and performance improvement of conductive asphalt concrete incorporating carbon fiber and iron tailings. Constr. Build. Mater. 2021, 303, 124446. [Google Scholar] [CrossRef]
  80. He, Y.; Zeng, Q.; Liu, Y.; Liu, P.; Zeng, Y.; Xu, Z.; Liu, Q. Evaluation of the Composite Mechanism of Nano-Fe2O3/Asphalt Based on Molecular Simulation and Experiments. Materials 2021, 14, 3425. [Google Scholar] [CrossRef] [PubMed]
  81. Sherre, T.K.; Liao, M.-C. Characteristics of recycled mineral fillers and their effects on the mechanical properties of hot-mix asphalt when used as limestone filler replacements. J. Mater. Civil Eng. 2022, 34, 04021395. [Google Scholar] [CrossRef]
  82. Zhu, S.; Kong, L.; Peng, Y.; Chen, Y.; Zhao, T.; Jian, O.; Zhao, P.; Sheng, X.; Li, Z. Mechanisms of interface electrostatic potential induced asphalt-aggregate adhesion. Constr. Build. Mater. 2024, 438, 137255. [Google Scholar] [CrossRef]
  83. Du, Z.; Zhu, X. Molecular dynamics simulation to investigate the adhesion and diffusion of asphalt binder on aggregate surfaces. Transp. Res. Rec. 2019, 2673, 500–512. [Google Scholar] [CrossRef]
  84. Yao, L.; Gao, W.; Ma, X.; Fu, H. Properties analysis of asphalt binders containing Bayer. red mud. Materials 2020, 13, 1122. [Google Scholar] [CrossRef]
  85. Padhan, A.; Paul, B. Unlocking the potential of red mud: Advanced strategies for economic optimization and sustainable recovery of critical minerals. J. Environ. Manag. 2025, 389, 126040. [Google Scholar] [CrossRef]
  86. Shen, H.; Liu, B.; Ekberg, C.; Zhang, S. Harmless disposal and resource utilization for secondary aluminum dross: A review. Sci. Total Environ. 2021, 760, 143968. [Google Scholar] [CrossRef]
  87. Mintsis, M.Y.; Galevsky, G.V.; Rudneva, V.V.; Galevsky, S.G. Formation and emission of sulfur dioxide in aluminum production. In Proceedings of the IOP Conference Series: Materials Science and Engineering, Bangkok, Thailand, 24–26 February 2018; IOP Publishing: Bristol, UK, 2018. [Google Scholar]
  88. Meng, Y.; Huang, J.; Lai, J.; Li, Y.; Pan, Z.; Lu, T. Feasibility and synergies of aluminum electrolysis spent refractory material and graphene oxide for rubber modified asphalt mixture. Constr. Build. Mater. 2025, 481, 141647. [Google Scholar] [CrossRef]
  89. Li, X.; Ren, D.; Fan, S.; Lin, Z.; Meng, F.; Wu, H.; Li, J. Synthesis of gel crosslinked composite material from spent cathode carbon blocks and its application in modified asphalt binders. J. Clean. Prod. 2024, 434, 140253. [Google Scholar] [CrossRef]
  90. Zhang, H.; Li, H.; Zhang, Y.; Wang, D.; Harvey, J.; Wang, H. Performance enhancement of porous asphalt pavement using red mud as alternative filler. Constr. Build. Mater. 2018, 160, 707–713. [Google Scholar] [CrossRef]
  91. Zhang, J.; Liu, S.; Yao, Z.; Wu, S.; Jiang, H.; Liang, M.; Qiao, Y. Environmental aspects and pavement properties of red mud waste as the replacement of mineral filler in asphalt mixture. Constr. Build. Mater. 2018, 180, 605–613. [Google Scholar] [CrossRef]
  92. Zhang, H.; Li, H.; Abdelhady, A.; Mo, D.; Yang, B.; Harvey, J. Optimum filler–bitumen ratio of asphalt mortar considering self-healing property. J. Mater. Civil Eng. 2019, 31, 04019166. [Google Scholar] [CrossRef]
  93. Lima, M.S.S.; Thives, L.P. Mechanical feasibility of using red mud as filler in asphalt mixtures to improve permanent deformation. Transportes 2020, 28, 1–13. [Google Scholar] [CrossRef]
  94. Tian, Y.; Sun, L.; Li, H.; Zhang, H.; Harvey, J.; Yang, B.; Zhu, Y.; Yu, B.; Fu, K. Laboratory investigation on effects of solid waste filler on mechanical properties of porous asphalt mixture. Constr. Build. Mater. 2021, 279, 122436. [Google Scholar] [CrossRef]
  95. Ou, L.; Zhu, H.; Chen, R.; Su, C.; Yang, X. Effect of industrial solid waste as fillers on the rheology and surface free energy of asphalt mastic. Materials 2024, 17, 1125. [Google Scholar] [CrossRef]
  96. Wu, D.; Yin, Y.; Fu, T.; Liu, F. Simulation study on basic road performance and modification mechanism of red mud modified asphalt mixture. Sci. Eng. Compos. Mater. 2024, 31, 20240003. [Google Scholar] [CrossRef]
  97. Zhang, J.; Sun, C.; Li, P.; Liang, M.; Jiang, H.; Yao, Z. Experimental study on rheological properties and moisture susceptibility of asphalt mastic containing red mud waste as a filler substitute. Constr. Build. Mater. 2019, 211, 159–166. [Google Scholar] [CrossRef]
  98. Liu, W.; Chen, X.; Li, W.; Yu, Y.; Yan, K. Environmental assessment, management and utilization of red mud in China. J. Clean. Prod. 2014, 84, 606–610. [Google Scholar] [CrossRef]
  99. Jiang, Y.; Liu, Y.; Lin, C. Chemistry and mineralogy of red mud and its utilization in Zhengzhou Alumina Refinery. Light Met. 2007, 10, 18–21. [Google Scholar]
  100. Nan, X.L.; Zhang, T.A.; Liu, Y.; Dou, Z.H.; Zhao, Q.Y. Main categories of red mud and its environmental impacts. Chin. J. Process Eng. 2009, 9, 459–464. [Google Scholar]
  101. Busari, A.A.; Adeboje, A.; Modupe, A.E.; Fadugba, G.O.; Loto, R.T.; Abatan, G.O. Assessment of the engineering properties of modified Asphalt using Aluminium dross as a filler. In Proceedings of the IOP Conference Series: Materials Science and Engineering, Suzhou, China, 17–19 March 2021; IOP Publishing: Bristol, UK, 2021. [Google Scholar]
  102. Ugla, S.K.; Ismael, M.Q. Evaluating the moisture susceptibility of asphalt mixtures containing RCA and modified by waste alumina. Civil. Eng. J. 2023, 9, 250–262. [Google Scholar]
  103. Ozerkan, N.; Maki, O.; Anayeh, M.; Tangen, S.; M Abdullah, A. The effect of aluminium dross on mechanical and corrosion properties of concrete. Int. J. Innov. Res. Sci. Eng. Technol. 2014, 3, 9912–9922. [Google Scholar]
  104. Kang, S.-P.; Kwon, S.-J. Effects of red mud and alkali-activated slag cement on efflorescence in cement mortar. Constr. Build. Mater. 2017, 133, 459–467. [Google Scholar] [CrossRef]
  105. Zhang, J.; Li, P.; Liang, M.; Jiang, H.; Yao, Z.; Zhang, X.; Yu, S. Utilization of red mud as an alternative mineral filler in asphalt mastics to replace natural limestone powder. Constr. Build. Mater. 2020, 237, 117821. [Google Scholar] [CrossRef]
  106. Hu, S.; Wang, D.; Hou, D.; Zhao, W.; Li, X.; Qu, T.; Zhu, Q. Research on the preparation parameters and basic properties of premelted calcium aluminate slag prepared from secondary aluminum dross. Materials 2021, 14, 5855. [Google Scholar] [CrossRef]
  107. Choudhary, J.; Kumar, B.; Gupta, A. Performance evaluation of bauxite residue modified asphalt concrete mixes. Eur. J. Environ. Civil. Eng. 2022, 26, 978–994. [Google Scholar] [CrossRef]
  108. Choudhary, J.; Kumar, B.; Gupta, A. Application of waste materials as fillers in bituminous mixes. Waste Manag. 2018, 78, 417–425. [Google Scholar] [CrossRef]
  109. Lima, M.S.; Thives, L.P. Evaluation of red mud as filler in Brazilian dense graded asphalt mixtures. Constr. Build. Mater. 2020, 260, 119894. [Google Scholar] [CrossRef]
  110. Soós, Z.; Géber, R.; Tóth, C.; Igazvölgyi, Z.; Udvardi, B. Utilization of aluminium dross as asphalt filler. Építőanyag J. Silic. Based Compos. Mater. 2017, 69, 89–93. [Google Scholar] [CrossRef]
  111. Shishehboran, M.; Ziari, H.; Korayem, A.H.; Hajiloo, M. Environmental and mechanical impacts of waste incinerated acidic sludge ash as filler in hot mix asphalt. Case Stud. Constr. Mater. 2021, 14, e00504. [Google Scholar] [CrossRef]
  112. Cui, Y.; Si, C.; Li, S.; Jia, Y.; Guo, B. Iron tailings as mineral fillers and their effect on the fatigue performance of asphalt mastic. Materials 2024, 17, 2927. [Google Scholar] [CrossRef] [PubMed]
  113. Kumar, A.; Kumar, S. Development of paving blocks from synergistic use of red mud and fly ash using geopolymerization. Constr. Build. Mater. 2013, 38, 865–871. [Google Scholar] [CrossRef]
  114. Choudhary, R.; Kumar, A.; Rahman, G. Rheological and mechanical properties of bauxite residue as hot mix asphalt filler. Int. J. Pavement Res. Technol. 2019, 12, 623–631. [Google Scholar] [CrossRef]
  115. Phan, T.M.; Kang, S.P.; Vo, H.V.; Park, D.W. Evaluation on performances of cold asphalt mixture containing recycled waste glass and red mud. Case Stud. Constr. Mater. 2024, 20, e03194. [Google Scholar] [CrossRef]
  116. Lima, M.S.S.; Thives, L.P.; Haritonovs, V.; Gschösser, F. The influence of alternative fillers on the adhesive properties of mastics fabricated with red mud. Materials 2020, 13, 484. [Google Scholar] [CrossRef]
  117. Rashidian, S.; Hosseini, S.A.; Korandeh, M.E. Investigation of red mud effects on microsurfacing asphalt performance: Quality enhancement and industrial waste reduction. Constr. Build. Mater. 2025, 483, 141798. [Google Scholar] [CrossRef]
  118. Ergezer, F.; Terzi, S. Investigation into the deicing efficiency and self-healing capability of asphalt mixtures with aluminum dross powder and steel fiber. Road Mater. Pavement Des. 2025, 26, 1–21. [Google Scholar] [CrossRef]
  119. Udvardi, B.; Géber, R.; Kocserha, I. Investigation of aluminum dross as a potential asphalt filler. Int. J. Eng. Manag. Sci. 2019, 4, 445–451. [Google Scholar] [CrossRef]
  120. Xu, F.; Zhao, Y.; Li, K. Using. waste plastics as asphalt modifier: A review. Materials 2021, 15, 110. [Google Scholar] [CrossRef]
  121. Ling, T.; Lu, Y.; Zhang, Z.; Li, C.; Oeser, M. Value-added application of waste rubber and waste plastic in asphalt binder as a multifunctional additive. Materials 2019, 12, 1280. [Google Scholar] [CrossRef]
  122. Elahi, Z.; Mohd Jakarni, F.; Muniandy, R.; Hassim, S.; Ab Razak, M.S.; Ansari, A.H.; Ben Zair, M.M. Waste cooking oil as a sustainable bio modifier for asphalt modification: A review. Sustainability 2021, 13, 11506. [Google Scholar] [CrossRef]
  123. Zhang, R.; You, Z.; Ji, J.; Shi, Q.; Suo, Z. A review of characteristics of bio-oils and their utilization as additives of asphalts. Molecules 2021, 26, 5049. [Google Scholar] [CrossRef] [PubMed]
  124. Tao, F.; Hui-ming, B. Molecular simulation study on modification mechanism of red mud modified asphalt. In Proceedings of the IOP Conference Series: Earth and Environmental Science, Pekanbaru, Indonesia, 25–26 July 2017; IOP Publishing: Bristol, UK, 2017. [Google Scholar]
  125. Fu, T.; Wei, J.; Bao, H.; Liang, J. Multiscale study on the modification mechanism of red mud modified asphalt. Adv. Mater. Sci. Eng. 2020, 2020, 2150215. [Google Scholar] [CrossRef]
  126. Zhou, X.; Wang, Z.; Wang, X.; Guo, H.; Ji, X.; Liu, J. Utilization of calcium carbide slag as alternative filler in asphalt mastic: Filler characteristics, rheological and adhesion properties. J. Clean. Prod. 2022, 380, 134980. [Google Scholar] [CrossRef]
  127. Qin, Y.; Xie, K.; Meng, Y.; Fu, T.; Fang, G.; Luo, X.; Wang, Q. Feasibility and environmental assessment of reusing aluminum tailing slurry in Asphalt. Constr. Build. Mater. 2024, 411, 134737. [Google Scholar] [CrossRef]
  128. Yue, H.; Zeng, Z.; Chen, M.; Xiao, J.; Bi, Y.; Zhang, H.; Ding, T.; Zhang, J. The Impacts of Red. Mud on Road Performance and Aging Resistance of Asphalt Mixture. J. Test. Eval. 2024, 52, 2451–2467. [Google Scholar] [CrossRef]
  129. Xiao, J.; Zhang, J.; Zhang, H.; Bi, Y.; Yue, H.; Xu, R. Preparation and characterization of organic red mud and its application in asphalt modification. Constr. Build. Mater. 2023, 367, 130269. [Google Scholar] [CrossRef]
  130. Han, F.; Li, C.; Luo, Y.; Lv, Y.; Tan, D.; Zhao, Z.; Li, J. Study on Properties of Siberian Cocklebur Bionic Harmless Red Mud Modified Asphalt. J. Appl. Polym. Sci. 2025, e57179. [Google Scholar] [CrossRef]
  131. Lee, H.; Dellatore, S.M.; Miller, W.M.; Messersmith, P.B. Mussel-inspired surface chemistry for multifunctional coatings. Science 2007, 318, 426–430. [Google Scholar] [CrossRef]
  132. Ren, D.; Jiang, H.; Fan, S.; Song, L.; Luo, Y.; Lai, F.; Pei, R.; Yang, Q.; Li, J. Polydopamine-modified black phosphorus/microcapsule composite material used to enhance the solar thermal conversion and self-healing performance of SBS modified asphalt. J. Appl. Polym. Sci. 2025, 142, e56316. [Google Scholar] [CrossRef]
  133. Li, C.; Guo, J.; Wu, Y.; Chen, J.; Bai, X.; Zou, H.; Li, J.; Li, J. Synthesis of flower cluster structure from red mud and rice straw through co-hydrothermal carbonization for modified asphalt applications. Constr. Build. Mater. 2025, 471, 140745. [Google Scholar] [CrossRef]
  134. Li, J.; Han, F.; Lv, Y.; Wei, X.; Deng, S.; Liu, Y. Research on the harmless treatment of RM with MTs in plant oil residue and its anti-aging in asphalt. Constr. Build. Mater. 2025, 483, 141797. [Google Scholar] [CrossRef]
  135. Fan, S.; Tu, H.; Muhammad, Y.; Ren, D.; Lai, F.; Yang, Q.; Zhao, Z.; Li, J. Resource utilization of emulsion evaporation modified aluminum industrial refractory waste in the preparation of composite modified asphalt. J. Appl. Polym. Sci. 2024, 141, e56092. [Google Scholar] [CrossRef]
  136. Li, J.; Tu, H.; Li, X.; Chen, Y.; Zhou, T.; Li, Z.; Zhao, Z.; Liu, Y. Resourceful modification mechanism of modified electrolytic aluminum spent refractory material and its long-term safety performance assessment. Constr. Build. Mater. 2024, 418, 135254. [Google Scholar] [CrossRef]
  137. ASTM D5/D5M–20; Standard Test Method for Penetration of Bituminous Materials. ASTM: West Conshohocken, PA, USA, 2020.
  138. ASTM D36/D36M-14; Standard Test Method for Softening Point of Bitumen (Ring-and-Ball Apparatus). ASTM: West Conshohocken, PA, USA, 2014.
  139. ASTM D113-17; Standard Test Method for Ductility of Asphalt Materials. ASTM: West Conshohocken, PA, USA, 2017.
  140. ASTM D4402/D4402M-15; Standard Test Method for Viscosity Determination of Asphalt at Elevated Temperatures Using a Rotational Viscometer. ASTM: West Conshohocken, PA, USA, 2015.
  141. ASTM D2872-22; Standard Test Method for Effect of Heat and Air on a Moving Film of Asphalt (Rolling Thin-Film Oven Test). ASTM: West Conshohocken, PA, USA, 2022.
  142. ASTM D6521-22; Standard Practice for Accelerated Aging of Asphalt Binder Using a Pressurized Aging Vessel (PAV). ASTM: West Conshohocken, PA, USA, 2022.
  143. AASHTO T 315-22; 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.
  144. AASHTO T350-19; Standard Method of Test for Multiple Stress Creep Recovery (MSCR) Test of Asphalt Binder Using a Dynamic Shear Rheometer. American Association of State Highway and Transportation Officials: Washington, DC, USA, 2019.
  145. AASHTO TP 101-14; Standard Method of Test for Estimating Damage Tolerance of Asphalt Binders Using the Linear Amplitude Sweep. American Association of State Highway and Transportation Officials: Washington, DC, USA, 2014.
  146. AASHTO T313-22; Standard Method of Test for Determining the Flexural Creep Stiffness of Asphalt Binder Using the Bending Beam Rheometer. American Association of State Highway and Transportation Officials: Washington, DC, USA, 2022.
  147. ASTM D6732-02; Standard Test Method for Determining the Failure Properties of Asphalt Binder in Direct Tension. ASTM: West Conshohocken, PA, USA, 2002.
  148. AASHTO TP-91; Standard Method of Test for Determining Asphalt Binder Bond Strength by Means of the Asphalt Bond Strength (ABS) Test. American Association of State Highway and Transportation Officials: Washington, DC, USA, 2011.
  149. Zhang, J.; Li, P.; Wang, K.; Ma, C.; Liang, M.; Jiang, H.; Yao, K.; Su, C.; Yao, Z. Adhesive behavior and pavement performance of asphalt mixtures incorporating red mud as a filler substitute. Constr. Build. Mater. 2021, 298, 123855. [Google Scholar] [CrossRef]
  150. Chaudhary, M.; Saboo, N.; Gupta, A. Introduction of a new parameter to quantify the fatigue damage in asphalt mastics and asphalt binder. Coatings 2021, 11, 828. [Google Scholar] [CrossRef]
  151. Chaudhary, M.; Saboo, N.; Gupta, A.; Steineder, M.; Hofko, B. Effect of analysis procedure and sample geometry on the fatigue life results of asphalt mastics from linear amplitude sweep test. Mech. Time-Depend. Mater. 2023, 27, 1097–1121. [Google Scholar] [CrossRef]
  152. Wang, H.; Li, H.; Zhang, H.; Zhang, X.; Guo, B.; Yu, B.; Tian, Y. Experimental study on the aging behavior of modified asphalt with different types of fine solid wastes under different aging conditions. Constr. Build. Mater. 2021, 291, 123308. [Google Scholar] [CrossRef]
  153. Xing, Z.; Hébert, R.; Beaucour, A.L.; Ledésert, B.; Noumowé, A. Influence of chemical and mineralogical composition of concrete aggregates on their behaviour at elevated temperature. Mater. Struct. 2014, 47, 1921–1940. [Google Scholar] [CrossRef]
  154. Chaudhary, M.; Saboo, N.; Gupta, A.; Hofko, B.; Steineder, M. Assessing the effect of fillers on LVE properties of asphalt mastics at intermediate temperatures. Mater. Struct. 2020, 53, 96. [Google Scholar] [CrossRef]
  155. Chaudhary, M.; Saboo, N.; Gupta, A.; Miljković, M. Contribution of mineral filler to the fatigue damage behaviour of bituminous mastic. Constr. Build. Mater. 2022, 334, 127120. [Google Scholar] [CrossRef]
  156. ASTM D6927-15; Standard Test Method for Marshall Stability and Flow of Asphalt Mixtures. ASTM: West Conshohocken, PA, USA, 2015.
  157. ASTM D6931-12; Standard Test Method for Indirect Tensile (IDT) Strength of Bituminous Mixtures. ASTM: West Conshohocken, PA, USA, 2012.
  158. AASHTO TP124-16; Standard Method of Test for Determining the Fracture Potential of Asphalt Mixtures Using Semicircular Bend Geometry (SCB) at Intermediate Temperature. American Association of State Highway and Transportation Officials: Washington, DC, USA, 2016.
  159. ASTM D8225-19; Standard Test Method for Determination of Cracking Tolerance Index of Asphalt Mixture Using the Indirect Tensile Cracking Test at Intermediate Temperature. ASTM: West Conshohocken, PA, USA, 2019.
  160. AASHTO T 324-11; Standard Method of Test for Hamburg Wheel-Track Testing of Compacted Hot Mix Asphalt (HMA). American Association of State Highway and Transportation Officials: Washington, DC, USA, 2011.
  161. AFNOR NF P 98-250-2; Essais Relatifs Aux Chaussées: Préparation des Mélanges Hydrocarbonés—Partie 2: Compactage des plaques. AFNOR Éditions: Paris, France, 1997.
  162. ASTM D1075-11; Standard Test Method for Effect of Water on Compressive Strength of Compacted Bituminous Mixtures. ASTM: West Conshohocken, PA, USA, 2011.
  163. AASHTO T 283-14; Standard Method of Test for Resistance of Compacted Asphalt Mixtures to Moisture-Induced Damage. American Association of State Highway and Transportation Officials: Washington, DC, USA, 2014.
  164. ASTM D3625-12; Standard Test Method for Effect of Water on Bituminous-Coated Aggregate Using Boiling Water. ASTM: West Conshohocken, PA, USA, 2012.
  165. AASHTO TP 108-14; Standard Method of Test for Determining the Abrasion Loss of Asphalt Mixture Specimens (Cantabro Test). American Association of State Highway and Transportation Officials: Washington, DC, USA, 2014.
  166. Choudhary, J.; Kumar, B.; Gupta, A. Effect of filler on the bitumen-aggregate adhesion in asphalt mix. Int. J. Pavement Eng. 2020, 21, 1482–1490. [Google Scholar] [CrossRef]
  167. Asphalt Institute. Mix Design Methods for Asphalt Concrete and Other Hot-Mix Types: Manual Series No. 2 (MS-2), 6th ed.; Asphalt Institute: Lexington, KY, USA, 1997. [Google Scholar]
  168. Zoorob, S.; Suparma, L. Laboratory design and investigation of the properties of continuously graded Asphaltic concrete containing recycled plastics aggregate replacement (Plastiphalt). Cem. Concr. Compos. 2000, 22, 233–242. [Google Scholar] [CrossRef]
  169. Modarres, A.; Bengar, P.A. Investigating the indirect tensile stiffness, toughness and fatigue life of hot mix asphalt containing copper slag powder. Int. J. Pavement Eng. 2019, 20, 977–985. [Google Scholar] [CrossRef]
  170. Lima, M.; Thives, L.P.; Haritonovs, V. Rutting performance of bituminous mixtures composed with red mud. In Bearing Capacity of Roads, Railways and Airfields; CRC Press: Boca Raton, FL, USA, 2017; pp. 349–355. [Google Scholar]
  171. Pasandín, A.; Pérez, I. The influence of the mineral filler on the adhesion between aggregates and bitumen. Int. J. Adhes. Adhes. 2015, 58, 53–58. [Google Scholar] [CrossRef]
  172. Shanchun, F.; Dongdong, W.; Hongyan, S. Study on Water Damage Resistance of Red Mud Mineral Powder Asphalt Mixture. Multipurp. Util. Miner. Resour. Kuangchan Zonghe Liyong 2022, 43, 25–30, 36. [Google Scholar]
  173. Tarrer, A. The Effect of the Physical and Chemical Characteristics of the Aggregate on Bonding; Strategic Highway Research Program (SHRP): Washington, DC, USA, 1991. [Google Scholar]
  174. Kandhal, P.S.; Chakraborty, S. Effect of asphalt film thickness on short-and long-term aging of asphalt paving mixtures. Transp. Res. Rec. 1996, 1535, 83–90. [Google Scholar] [CrossRef]
  175. Antunes, M.; Conceição, F.; Navarro, G. Caracterização da lama vermelha brasileira (resíduo do refino da bauxita) e avaliação de suas propriedades para futuras aplicações. Int. Workshop Adv. Clean. Prod. 2011, 3, 10. [Google Scholar]
  176. Collazo, A.; Fernández, D.; Izquierdo, M.; Nóvoa, X.R.; Pérez, C. Evaluation of red mud as surface treatment for carbon steel prior painting. Progress. Org. Coat. 2005, 52, 351–358. [Google Scholar] [CrossRef]
  177. Nunn, R.F. Advances in red mud dewatering and disposal technologies. Light Met. Warrendale 1998, 107–114. [Google Scholar]
  178. Lima, F. Utilização da Lama Vermelha e do Resíduo Caulinítico na Produção de Pigmento Pozolânico Para Argamassas e Concretos de Cimento Portland; UFPA: Belém, PA, Brazil, 2006. [Google Scholar]
  179. Lima, M.S.S.; Hajibabaei, M.; Thives, L.P.; Haritonovs, V.; Buttgereit, A.; Queiroz, C.; Gschösser, F. Environmental potentials of asphalt mixtures fabricated with red mud and fly ash. Road Mater. Pavement Des. 2021, 22 (Suppl. S1), S690–S701. [Google Scholar] [CrossRef]
  180. Li, H.; Liu, J.; Zhang, H.; Harvey, J. Investigation on the effect of fine solid wastes on the runoff purification performance of porous asphalt mixture. J. Environ. Manag. 2021, 300, 113612. [Google Scholar] [CrossRef] [PubMed]
  181. Liu, J.; Li, H.; Sun, L.; Zhang, H.; Xie, N.; Tian, Y.; Jia, M.; Fu, K.; Yu, B.; Zhu, Y. Leachate risks of fine solid wastes in porous asphalt pavement and runoff purification effects of diatomite filler. J. Clean. Prod. 2021, 297, 126623. [Google Scholar] [CrossRef]
  182. Modarres, A.; Ramyar, H.; Ayar, P. Effect of cement kiln dust on the low-temperature durability and fatigue life of hot mix asphalt. Cold Reg. Sci. Technol. 2015, 110, 59–66. [Google Scholar] [CrossRef]
  183. Busari, A.; Snyman, J.; Kupolati, W.; Ndambuki, J.; Sadiku, E.; Loto, R.; Sagay, A. Data on the engineering properties of aluminum dross as a filler in asphalt. Data Brief. 2020, 31, 105934. [Google Scholar] [CrossRef]
  184. Ugla, S.K.; Ismael, M.Q. Evaluating the moisture susceptibility of asphalt mixtures containing aluminum dross as a filler. J. Eng. 2024, 30, 190–206. [Google Scholar] [CrossRef]
  185. Yinfei, D.; Jiacheng, W.; Jiaqi, C. Cooling asphalt pavement by increasing thermal conductivity of steel fiber asphalt mixture. Sol. Energy 2021, 217, 308–316. [Google Scholar] [CrossRef]
  186. Serin, S.; Emiroğlu, M.; Gönül, V.E. Investigation of the fracture energy of hot mixtures asphalt incorporating metallic wastes via semi-circular bending test. Constr. Build. Mater. 2021, 300, 124006. [Google Scholar] [CrossRef]
  187. Al-Obaidi, M.K.; Abed, A.H. Appraisal of asphalt mixtures’ cracking resistance modified with waste aluminum scrap powder (wasp) at intermediate temperatures. In Proceedings of the E3S Web of Conferences, Semarang, Indonesia, 8–9 August 2023; EDP Sciences: London, UK, 2023. [Google Scholar]
  188. Atakan, M.; Yıldız, K. Self-healing potential of porous asphalt concrete containing different aggregates and metal wastes through microwave heating. Politek. Derg. 2022, 25, 623–631. [Google Scholar] [CrossRef]
  189. Pamulapati, Y.; Elseifi, M.A.; Cooper, S.B., III; Mohammad, L.N.; Elbagalati, O. Evaluation of self-healing of asphalt concrete through induction heating and metallic fibers. Constr. Build. Mater. 2017, 146, 66–75. [Google Scholar] [CrossRef]
  190. Bastidas-Martínez, J.G.; Rondón-Quintana, H.A.; Mojica-Herrera, L.F. Behavior of a hot-mix asphalt with aluminum shavings. Constr. Build. Mater. 2025, 474, 140905. [Google Scholar] [CrossRef]
Figure 1. Different types of aluminum production residues: (a) Bayer RM, (b) Sintering RM, (c) AD, (d) SCCB, (e) SRM, and (f) aluminum shavings.
Figure 1. Different types of aluminum production residues: (a) Bayer RM, (b) Sintering RM, (c) AD, (d) SCCB, (e) SRM, and (f) aluminum shavings.
Sustainability 17 09634 g001
Figure 2. Multiscale schematic of bituminous materials [72] (reprinted with permission).
Figure 2. Multiscale schematic of bituminous materials [72] (reprinted with permission).
Sustainability 17 09634 g002
Figure 4. Mineral compositions of (a) Bayer RM [104], (b) Sintering RM [105], (c) AD [106], (d) SRM [88], and (e) SCCB [89].
Figure 4. Mineral compositions of (a) Bayer RM [104], (b) Sintering RM [105], (c) AD [106], (d) SRM [88], and (e) SCCB [89].
Sustainability 17 09634 g004
Figure 7. (a) Preparation steps and mechanisms of modified RM [130] (reprinted with permission), (b) preparation steps of c-HTC and H-HTC [133], and (c) preparation process of modified RM (CZ, DD, and HS refer to canola oil residue, soybean oil residue, and peanut oil residue, respectively) and the interaction mechanism between metallothioneins and metal ions [134].
Figure 7. (a) Preparation steps and mechanisms of modified RM [130] (reprinted with permission), (b) preparation steps of c-HTC and H-HTC [133], and (c) preparation process of modified RM (CZ, DD, and HS refer to canola oil residue, soybean oil residue, and peanut oil residue, respectively) and the interaction mechanism between metallothioneins and metal ions [134].
Sustainability 17 09634 g007
Figure 8. SEM images of (a) unmodified RM; (b) c-HTC; and (c) H-HTC [133]; (d) unmodified RM and (e) RM modified with canola oil residue, (f) soybean oil residue, and (g) peanut oil residue [134] (reprinted with permission).
Figure 8. SEM images of (a) unmodified RM; (b) c-HTC; and (c) H-HTC [133]; (d) unmodified RM and (e) RM modified with canola oil residue, (f) soybean oil residue, and (g) peanut oil residue [134] (reprinted with permission).
Sustainability 17 09634 g008
Figure 9. FTIR spectra of RM, c-HTC, and H-HTC [133].
Figure 9. FTIR spectra of RM, c-HTC, and H-HTC [133].
Sustainability 17 09634 g009
Figure 10. (a) Preparation process of the thermoplastic polyester elastomer–spent refractory material (TPEE–SRM) composite using epoxy resins (EPs) and triethylenetetramine (TETA) as a curing agent [135] (reprinted with permission), (b) modification process of SRM and the performance mechanism of HMDS–TA–SRM [136], and (c) preparation process and mechanism of PSB [89].
Figure 10. (a) Preparation process of the thermoplastic polyester elastomer–spent refractory material (TPEE–SRM) composite using epoxy resins (EPs) and triethylenetetramine (TETA) as a curing agent [135] (reprinted with permission), (b) modification process of SRM and the performance mechanism of HMDS–TA–SRM [136], and (c) preparation process and mechanism of PSB [89].
Sustainability 17 09634 g010
Figure 11. (a) Contact angle of an asphalt mastic incorporating 40% RM [116], (b) contact angle of asphalt mastics containing different fillers [95].
Figure 11. (a) Contact angle of an asphalt mastic incorporating 40% RM [116], (b) contact angle of asphalt mastics containing different fillers [95].
Sustainability 17 09634 g011
Figure 12. (a) Dynamic stability and (b) creep slope results from the Hamburg wheel tracking test [94].
Figure 12. (a) Dynamic stability and (b) creep slope results from the Hamburg wheel tracking test [94].
Sustainability 17 09634 g012
Figure 13. Quantitative indices and their variations obtained from the SCB test: (a) peak load, (b) fracture energy, (c) slope at inflection point, and (d) flexibility index [94].
Figure 13. Quantitative indices and their variations obtained from the SCB test: (a) peak load, (b) fracture energy, (c) slope at inflection point, and (d) flexibility index [94].
Sustainability 17 09634 g013
Figure 14. (a) IDEAL-CT test setup and (b) testbed road after two years of service [115].
Figure 14. (a) IDEAL-CT test setup and (b) testbed road after two years of service [115].
Sustainability 17 09634 g014
Figure 15. Setup of a rainfall simulation device [180].
Figure 15. Setup of a rainfall simulation device [180].
Sustainability 17 09634 g015
Figure 16. (a) Cole–Cole diagram and (b) dynamic modulus master curve of HMA containing varying contents of aluminum dross [110].
Figure 16. (a) Cole–Cole diagram and (b) dynamic modulus master curve of HMA containing varying contents of aluminum dross [110].
Sustainability 17 09634 g016
Figure 17. (a) SCB setup (dimensions are in mm) and (b) SEM images of the reference sample (top) and aluminum chip–containing sample (bottom) [186].
Figure 17. (a) SCB setup (dimensions are in mm) and (b) SEM images of the reference sample (top) and aluminum chip–containing sample (bottom) [186].
Sustainability 17 09634 g017
Table 1. Chemical composition of SRM and SCCB.
Table 1. Chemical composition of SRM and SCCB.
SRMRef.
CaOSiO2MgOAl2O3Fe2O3MnO
54.30.820.620.520.160.15[88]
SCCB
CNaFFe2O3CaOSO3K2O[89]
79.69.233.002.031.641.12
Table 3. Series of tests and simulations for evaluating the performance of APR asphalt mastics.
Table 3. Series of tests and simulations for evaluating the performance of APR asphalt mastics.
PropertyTest and Simulation MethodsStandard Code
PenetrationNeedle penetrationASTM D5 [137]
Softening pointRing and ballASTM D36 [138]
DuctilityElongationASTM D113 [139]
ViscosityRotational viscosityASTM D4402 [140]
Short-term agingRolling thin-film oven testASTM D2872 [141]
Long-term agingPressure aging vesselASTM D6521 [142]
Linear viscoelastic behaviorFrequency and temperature sweep tests AASHTO T315 [143]
Permanent deformation resistance and elastic behaviorMultiple stress creep recovery testAASHTO T350 [144]
Fatigue resistance Linear amplitude sweep testAASHTO TP 101 [145]
Low temperature stiffnessBending beam rheometerAASHTO T313 [146]
Cracking resistance at low temperatureDirect tension testASTM D6732-02 [147]
Adhesion and bonding performanceBinder bond strengthAASHTO TP-91 [148]
Table 4. Physicochemical findings and rheological properties of asphalt mastics with different types of fillers.
Table 4. Physicochemical findings and rheological properties of asphalt mastics with different types of fillers.
Ref.Filler TypeF–B RatiosPhysicochemical and Microstructural AssessmentRheological Characterization
[114]RM and stone dust (SD)0.6, 0.9, 1.2, and 1.5 (w/w)✓ RM consisted of rounded, agglomerated particles with a rough texture, whereas SD contained flaky, irregularly shaped particles.
✓ RM was significantly finer than SD and exhibited higher specific gravity and Rigden voids.
✓ G* was comparable for both fillers, but the SD mastic was stiffer, at a F–B ratio of 1.5.
✓ SD mastics had a lower phase angle, indicating higher elasticity compared to RM mastics.
✓ SD mastics had lower Jnr values, at 3.2 kPa (Jnr 3.2), suggesting higher stiffness than RM mastics.
✓ At lower F–B ratios, SD mastics had a higher recovery. However, at an F–B ratio of 1.5, RM mastics exhibited greater recovery due to a faster increase in recovery with the F–B ratio.
[154]RM, marble dust (MD), limestone (LS), granite (GR), basalt (BA), and quartz (QZ)0.5, 1, and 1.5 (w/w)✓ RM exhibited the lowest fineness modulus, the highest Rigden value (RV), and the highest filler grain coefficient, indicating a well-graded and fine particle distribution.✓ The RM asphalt mastic exhibited the highest G*LVE of all the samples.
[151]RM and LS0.5, 1, and 1.5 (w/w)✓ Compared to LS, RM showed higher specific gravity and RV but a lower fineness modulus.✓ There was no consistent superiority between RM and LS fillers in terms of fatigue life, as performance varied depending on test conditions, analysis methods, and filler-to-binder ratios.
✓ The pseudo-strain energy method consistently produced the highest fatigue life estimates for both RM and LS asphalt mastics, followed by dissipated energy and R-based approaches.
✓ Hyperbolic geometry proved more effective in capturing fatigue damage in asphalt mastics for both types of fillers.
[150]RM and conventional filler10%, 20%, and 30% (v/v) ✓ Increased RM content resulted in greater fatigue damage.
✓ The newly proposed fatigue parameter (F) proved to be an effective indicator of fatigue damage in both asphalt mastics and binders.
[155]RM, LS, and MD10%, 20%, and 30% (v/v)✓ RM was much finer, with approximately 40% of RM particles smaller than 10 μm, compared to ~12% for the other fillers.✓ Mastics containing 30% RM filler were excessively stiff and required more time for homogenization, unlike the other mixes.
✓ RM showed the most significant sensitivity of the G* and Glover–Rowe parameters to changes in filler volume fraction.
✓ The fine, porous structure of RM and its strong binder interaction significantly influenced fatigue damage evolution, whereas MD and LS fillers had more moderate effects, with temperature being a more dominant factor in their performance.
✓ The strong contribution of RM was linked to its intensive interaction with asphalt binder, with a critical threshold observed at a 20% volume fraction, likely due to particle interlocking.
[116]RM, LS, dolomite, and fly ash (FA) 20% and 40% (v/v)✓ RM exhibited the highest surface area, RV, and pH among the fillers.✓ RM asphalt mastics demonstrated excellent adhesion to aggregates. Mastics containing 40% RM achieved the required minimum aggregate coverage (≥85%) in the boiling water stripping test without the need for any chemical adhesion improver, indicating robust adhesive properties.
✓ The wettability test results showed that mastics with RM had favorable contact angles, indicating good affinity between the mastics and water, which is important for durability in wet conditions.
[95]RM, LS, steel slag (SS), and ground granulated blast-furnace slag (GGBFS)1 (w/w)✓ RM showed a denser microstructure than the other fillers, with fine, agglomerated particles.✓ RM asphalt mastic exhibited the lowest G* and the highest phase angle.
✓ RM asphalt mastic showed better fatigue performance than LS asphalt mastic.
✓ Unlike SS, RM showed the lowest recovery and the highest Jnr, indicating the poorest resistance to rutting.
[152]RM, LS,
FA, and diatomite (DT)
33% (v/v)✓ The RV values of RM and DT were significantly higher than those of LS and FA, indicating a more porous structure.✓ RM asphalt mastic retained its low-temperature flexibility better than LS mastic after prolonged aging.
✓ The stress sensitivity of the RM asphalt mastic in the fatigue test was higher than that of LS or FA mastics, due to differences in the effect of physical hardening caused by the porous structure of RM.
✓ In unaged samples, RM asphalt mastic exhibited a higher percentage of recovery than LS mastic.
[84]RM and LS 0.3, 0.6, 0.9, 1.2, 1.5, 1.8,
and 2.1 (w/w)
✓ Incorporating and increasing the amount of RM led to a rise in rotational viscosity, G*, rutting index, and creep stiffness, all of which increased further with a higher F–B ratio.
✓ Increasing the substitution of LS with RM increased creep stiffness (S), while the rate of change in stiffness (m-value) decreased.
[90]LS and RM0.3, 0.6, and 0.9 (w/w) ✓ Based on the rutting index, the high-temperature performance of RM was superior to that of LM, especially when the F–B ratio was at least 0.6.
✓ Incorporating RM increased percentage recovery and decreased Jnr, indicating improved elasticity and resistance to permanent deformation.
✓ Increasing the F–B ratio reduced low-temperature cracking, and RM did not consistently outperform LS in low-temperature performance.
[91]Conventional filler, Sintering RM, and Bayer RM1 (w/w)✓ Sintering RM contained more aggregated particles than Bayer RM, despite their similar particle sizes. The pore volume of RM was more than five times greater than that of conventional mineral filler.✓ Sintering RM more effectively enhanced G* and rutting resistance. It also improved elastic recovery, unlike Bayer RM, which had a minimal effect.
[97]RM, LS, hydrated lime (HL), and FA1 (w/w)✓ RM had smaller, rougher, and more porous particles than LS, with fine, near-spherical particles prone to agglomeration, affecting asphalt binder absorption and asphalt mastic rheology.✓ The viscosity of RM asphalt mastic was over four times higher than that of LS mastic.
✓ The fatigue parameter (G*sin δ) of RM mastic showed the highest values, especially at low frequencies.
[92]RM, LS, FA, and DT0.11, 0.22, 0.33, 0.44, and 0.55 (v/v) ✓ Larger RV in fillers such as DT and RM increased asphalt absorption and particle friction, leading to higher viscosity and lower Jnr 3.2 than LS and FA asphalt mastics.
[105]RM and LS1 (w/w)✓ RM consisted of near-spherical, fine particles with coarse, edgeless surfaces, which tended to agglomerate into larger clusters.✓ RM improved the elastic recovery of asphalt mastic but negatively affected cracking and fatigue resistance, which could be mitigated by adding white mud.
✓ RM significantly increased the viscosity of mastic, and adding HL amplified this effect, while white mud slightly reduced it.
[88]SRM0%, 50%, and 100% (v/v)✓ Both SRM and graphene oxide (GO) featured wrinkled and grooved surfaces that effectively adsorbed free asphalt without undergoing any chemical reaction with it.✓ The synergistic effect of GO as an asphalt modifier and SRM as a filler improved asphalt mastic stiffness, reduced permanent deformation, and lowered stress sensitivity, mainly due to the CaO and Al2O3 content in SRM and the surface structure of GO.
Table 5. Common experimental tests used to evaluate asphalt mixtures containing APRs.
Table 5. Common experimental tests used to evaluate asphalt mixtures containing APRs.
PropertyTestStandard Code
Strength and flowMarshall stability and flow testASTM D6927 [156]
Tensile strengthIndirect tensile strength testASTM D6931-12 [157]
Cracking resistanceSemicircular bend testAASHTO TP124-16 [158]
Indirect tensile asphalt cracking testASTM D8225 [159]
Rutting resistanceHamburg wheel tracking testAASHTO T 324–11 [160]
French wheel tracker testNF P 98-250-2 [161]
Moisture resistanceMoisture susceptibility testASTM D1075-11 [162]
AASHTO T283 [163]
Adhesion and bondingBoiling water testASTM D3625-12 [164]
Raveling resistanceCantabro abrasion loss testAASHTO TP108-14 [165]
Table 6. Key studies on the valorization of RM in HMA.
Table 6. Key studies on the valorization of RM in HMA.
Reference FillerGradation TypeType of Asphalt BinderEffect of RMRef.
OBC% Stability and FlowRutting ResistanceTensile Strength and Cracking ResistanceMoisture SusceptibilityAdhesion Raveling ResistanceAging Resistance
OPCDense-graded60–70 penetration grade- [108]
SDDense-gradedVG-30 [114]
SDDense-gradedVG-30 [107]
SDDense-gradedVG-30 [166]
SDDense-gradedPG 64-22 [109]
LSDense-graded60–80 penetration grade [149]
LSOpen-gradedSBS asphalt (PG 76-22) [90]
LSOpen-gradedSBS asphalt (PG 82-28) [94]
Note: ↑ indicates an increase, ↓ indicates a decrease, and - indicates no change.
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

Salehfard, R.; Jafari, R. Studies on the Valorization of Aluminum Production Residues into Bituminous Materials at Different Scales: A Review. Sustainability 2025, 17, 9634. https://doi.org/10.3390/su17219634

AMA Style

Salehfard R, Jafari R. Studies on the Valorization of Aluminum Production Residues into Bituminous Materials at Different Scales: A Review. Sustainability. 2025; 17(21):9634. https://doi.org/10.3390/su17219634

Chicago/Turabian Style

Salehfard, Reza, and Reza Jafari. 2025. "Studies on the Valorization of Aluminum Production Residues into Bituminous Materials at Different Scales: A Review" Sustainability 17, no. 21: 9634. https://doi.org/10.3390/su17219634

APA Style

Salehfard, R., & Jafari, R. (2025). Studies on the Valorization of Aluminum Production Residues into Bituminous Materials at Different Scales: A Review. Sustainability, 17(21), 9634. https://doi.org/10.3390/su17219634

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