Next Article in Journal
Surface Enhancement of CoCrMo Bioimplant Alloy via Nanosecond and Femtosecond Laser Processing with Thermal Treatment
Previous Article in Journal
A Practical 1D Approach for Real-Time Prediction of Argon Flow and Pressure in Continuous Casting of Steel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Development Status of Production Purification and Casting and Rolling Technology of Electrical Aluminum Rod

School of Materials and Metallurgy, Guizhou University, Guiyang 550025, China
*
Author to whom correspondence should be addressed.
Metals 2025, 15(9), 981; https://doi.org/10.3390/met15090981
Submission received: 8 August 2025 / Revised: 28 August 2025 / Accepted: 30 August 2025 / Published: 1 September 2025

Abstract

As the demand for lightweight and high-performance conductive materials grows in power transmission systems, aluminum alloy rods have emerged as a cost-effective and scalable alternative to copper conductors. This review systematically examines the development status and technological progress in the purification and casting–rolling processes used in the production of Electrical Round Aluminum Rods (ERARs). It explores current challenges in improving electrical conductivity and mechanical strength while addressing issues such as hydrogen and oxide inclusion removal, grain refinement, and impurity segregation. Key purification techniques—including flux refining, gas treatment, filtration, and rotary injection—are compared in terms of performance, cost, and environmental impact. The paper also analyzes different casting–rolling methods, including continuous casting and rolling, twin-roll casting, and extrusion processes, with attention to process optimization and equipment design. Furthermore, emerging applications of artificial intelligence (AI) in predictive modeling, defect detection, and process parameter optimization are highlighted, offering a novel perspective on intelligent and sustainable ERAR production. This paper aims to provide insights for facilitating the industrial-scale production and performance enhancement of ERAR materials.

1. Introduction

Aluminum is one of the most abundant elements in Earth’s crust and the most widely used non-ferrous metal [1], owing to its combination of low density, low cost, high reflectivity, workability, relatively high alloying strength, and high electrical conductivity [2,3,4,5]. Its electrical conductivity is 64.94% IACS, while its thermal conductivity is approximately 237 W·m−1·K−1 at room temperature [2,3]. Although copper is traditionally regarded as the standard conductor, aluminum offers several key advantages. Per unit weight, aluminum provides nearly twice the conductivity of copper while being significantly more cost-effective [6,7]. Applications requiring high electrical conductivity account for approximately 20% of total aluminum usage now [8]. Moreover, only 0.01% of Earth’s crust is made up of copper, and its availability is vulnerable to geopolitical tensions [9].
With the expansion of industrial scale and the growth of long-distance transmission demand, reducing line losses and improving transmission stability and capacity have become the core goals of power system optimization. In this context, aluminum alloy wire, with its excellent electrical conductivity and cost advantages [10], has become the core material carrier of the strategy of “saving copper with aluminum”. ERAR is a kind of raw material used to make aluminum alloy cables. The preparation of round aluminum rods for electricians can roughly be summarized into two steps: purification process and casting and rolling process. The purification process improves conductivity by increasing the component ratio and purity of ERAR and optimizing the microcrystalline structure. The casting and rolling process, on the other hand, largely influences mechanical properties such as elongation and tensile strength, as well as service life [11,12,13]. This paper reviews the development achievements of these two processes, providing a reference basis for scientific research. Therefore, this review aims to provide a comprehensive overview of the current state of purification and cast-rolling technologies for ERAR, to identify key challenges in improving electrical and mechanical properties, and to explore how AI-driven methodologies can contribute to future process optimization.
The literature review was primarily conducted using the Web of Science (WoS) Core Collection and the China National Knowledge Infrastructure (CNKI) databases. The search covered publications from 2010 to 2024, using keywords such as “electrical round aluminum rod”, “aluminum alloy”, “electrical conductivity”, “continuous casting and rolling”, “material microstructure and properties”, and “process optimization”.

2. Aluminum Alloy Cable Market Size and Classification

Due to their favorable properties, including low density and high electrical conductivity, aluminum alloy conductors have seen rapid development, particularly in the production of electrical aluminum rods. With the expansion of the aluminum alloy cable market and the ongoing structural upgrading of the electrical aluminum rods industry, the range of products used in electrical applications has become increasingly diverse. Consequently, the establishment of detailed specifications and standardized constraints for electrical aluminum rods models used in electrical systems has become a pressing necessity.

2.1. Aluminum Alloy Cable Market Size

Compared with traditional pure copper and aluminum core cables, aluminum alloy cables show significant comprehensive performance advantages. Its density (2.7 g/cm3) is only 30% of that of copper. Based on maintaining a 61.8% IACS conductivity, the conductor cross-sectional area can be optimized to 1.6 times that of a copper core cable, and the weight per unit length can be reduced by 50% [14,15]. By alloying with elements such as Fe, Mg, and Si, aluminum alloy cables achieve a tensile strength of 113.8 MPa and an elongation of 30%, effectively overcoming the defects of the insufficient mechanical properties of pure aluminum materials (see Table 1). This improved corrosion resistance and creep resistance make them a compelling alternative in the cable industry.
According to market data, China’s total cable industry output reached RMB 228.4 billion in 2020, representing a 13.3% year-on-year increase, and accounting for 37.8% of power transmission sector’s total output. Among this, aluminum alloy cables contributed RMB 72.3 billion, representing 31.6% of the cable industry’s total output. Driven by the ongoing “aluminum replacing copper” strategy, the replacement rate of aluminum alloy cables has approached the international averages. Consequently, the Chinese market continues to expand (Figure 1), and the recent investments of the State Grid Corporation of China have been increasing continuously (Figure 2). The application share of aluminum alloy cables in China in new energy grid connection and ultra-high-voltage (UHV) power transmission exceeds 45% [16].

2.2. Classification and Performance Requirements of Round Aluminum Rods for Electricians

ERAR is extensively used in cables, wires, transformers, reactors, generators, and busbars in the aluminum electrolysis industry. With the rapid development of ultra-high-voltage (UHV) grids, long-distance transmission systems, new energy infrastructure, electric vehicles, and the communication industry, the demand for ERAR has grown significantly, presenting strong market potential. Figure 3 [18] show the average market price of a 9.5 mm round aluminum pole for electrical engineering. China’s power grid investment reached RMB 520 billion in 2023 and exceeded RMB 600 billion in 2024. It is projected to surpass RMB 650 billion in 2025, indicating a consistent year-on-year growth trend in grid investment. It can be seen through the SMM website; we can clearly see the price of 9.5 mm electrician round aluminum rods peaked in 2022 at 20,860 RMB/t. In 2023, the average price declined to 19,588 RMB/t. It is projected to rise again to 20,278 RMB/t in 2024. Overall, the price has remained around 20,000 RMB/t over the past three years [18].
High-strength and high-conductivity ERAR are essential to ensure the safety, reliability, and energy efficiency of downstream applications. However, owing to the impurities and rolling process conditions, it remains challenging to simultaneously achieve optimal electrical and mechanical properties. As one of the products of aluminum deep processing, ERAR offers low weight, corrosion resistance, and excellent electrical conductivity and thermal conductivity, making it widely applied in wires and cables, electronics, communications, and automobiles. With the development of new energy and aerospace, its potential applications are also expanding.
In order to meet the expanding market demand for aluminum alloy cables, the development of ERAR products has progressed rapidly, and many ERAR grades have been put into the market (see Table 2). Different countries employ distinct systems for standardizing aluminum conductor grades. However, the same type of aluminum alloy can be found with corresponding grades in the standards of various countries. As a developed country, the United States has a mature system for aluminum alloy grades, which adopts the naming convention of AA (representing the Aluminum Association) followed by four digits. It classifies aluminum alloys into eight series, and each series has a similar composition. The naming rule in Europe is to use “EN AW-” as a prefix, with the numerical part corresponding to the American grade (Table 3 compares the specific differences between European and American standards corresponding to the same brand ERAR). Notably, Europe places greater emphasis on processing states in its standards, directly incorporating them into grades, and imposes stricter limits on impurity elements while allowing slightly broader ranges for other components than the American system.

3. Development of the Purification Process

Purification is essential for producing high-purity ERAR, directly affecting downstream casting and rolling. Hydrogen and oxide inclusions, formed by the reaction between liquid aluminum and water vapor, significantly reduce ERAR’s mechanical properties. Therefore, purification—via physical and chemical methods—is required to eliminate hydrogen and other impurities from the melt (Typical impurities in aluminum conductors are shown on Table 4). Common techniques include fluxing, gas purification, filtration, and rotary injection, and are classified as adsorptive, non-adsorptive, or composite purification [20].

3.1. Fluxing Method

The fluxing method uses molten salts to remove impurities through chemical reactions and physical adsorption, improving melt cleanliness and fluidity. Fluxes are broadly categorized into refining agents for impurity removal and intermediate alloys for alloying and grain refinement.

3.1.1. Refining Agent

Fluxing is a key method for purifying aluminum melts, removing impurities through chemical reactions and physical adsorption between molten salts and contaminants. The impurity removal mechanism is illustrated in Figure 4 [23] and Figure 5 [24]. Early refining agents were mainly chloride based (e.g., KCl, NaCl, MgCl2), which reacted with Al2O3 to form low-melting-point compounds, aiding inclusion flotation.
Kulikov et al. [25] conducted a systematic comparative study of five widely used aluminum refining fluxes (SF1–SF5), employing XRD, DTA, and TGA analyses alongside direct chill casting experiments under near-industrial conditions. Their research examined the chemical composition of flux, moisture content, melting points, particle size distributions, and refining efficiency.
The experiments revealed that each flux contained crystalline hydrates, which decomposed at high temperatures to release water vapor. This vapor reacted with molten aluminum to produce hydrogen and oxide inclusions, further worsening melt contamination. Figure 6 and Figure 7 show that the refining agents were composed entirely of chlorides, highlighting the direct link between flux function and chemical makeup. Meanwhile, Figure 8 presents DTA–TGA thermal analysis curves that precisely identified the temperature ranges and weight losses associated with moisture release, confirming that water content is a critical negative factor in refining performance. Due to poor wettability with aluminum, these agents achieved only 30% impurity removal and were mainly used as surface cover agents to prevent oxidation.
Incorporating fluoride compounds into refining agents markedly improves impurity removal efficiency and overall refining performance. Adding fluorides such as Na3AlF6 and MgF2 to the KCl–NaCl system reduces interfacial tension and wetting angle, forms a protective surface layer, and increases impurity removal efficiency to 70–90% [24]. Figure 9 [23] illustrates the impact of various fluorides on interfacial tension with molten aluminum. Experimental evaluations identified KF combined with K3AlF6 or KAlF4 as the most effective formulations, as shown in Figure 10 [23], which depicts aluminum droplet coalescence in KAlF4 solutions. Shi et al. [26] reported that chloride–fluoride synergy enhances alumina dissolution and improves flux resistance to melt corrosion. This composite system effectively removes impurities and hydrogen from A356 alloy, promotes grain refinement, and improves mechanical properties. Consequently, refining agent design has evolved beyond traditional chloride systems toward more advanced, multifunctional formulations.
Composite refining agents have gradually replaced traditional chloride-based salts due to their superior performance. These agents integrate multiple functions through synergistic components: chlorides offer surface coverage, fluorides aid slag-crystal separation, carbonates release gas bubbles that promote flotation, and nitrates improve melt fluidity via exothermic reactions. For example, the Mg3N2 flux developed by Chen [27] refines and floats impurities by releasing nitrogen gas. GFLUX-J100 [28] reduces aluminum loss during refining, while AJ401K [29] lowers oxidation loss and enhances purification efficiency.
Further innovations include potassium cryolite-based fluxes, which showed excellent performance in aluminum alloy refining (Tkacheva et al. [30]), and a Na2SO4–NaCl system that effectively removes oxide inclusions (Widyantoro et al. [31]). Rare-earth elements (REEs) offer promising potential: they form stable compounds with hydrogen and oxygen; reduce surface tension; and enrich at grain boundaries, inhibiting defect formation [32]. For instance, La and Ce improve bonding strength and mechanical properties, while the LaF3-containing flux developed by Guan [21] enables efficient melt purification by releasing atomic-state RE species.
From a theoretical standpoint, Zhang et al. [33] optimized a NaF–Na2CO3–CaF2–Na3AlF6 composite using thermodynamic analysis, achieving low wetting angles with Al2O3 and excellent refining performance. Current research emphasizes optimizing flux composition ratios and developing multifunctional agents capable of degassing, impurity removal, and grain refinement.
In summary, as shown in Table 5, refining agents vary in cost, efficiency, and environmental impact. While chloride salts remain widely used due to their low cost and simplicity, they produce excessive smoke and slag, posing environmental and recovery challenges. By incorporating thermally stable additives like carbonates, sulfates, and nitrates, composite agents significantly reduce emissions and slag—by over 50% and 30%, respectively—while improving refining efficiency. The inclusion of rare-earth elements has become a leading research direction, though it also increases costs. Balancing performance and economic feasibility remains a key challenge. Nano-composite refining agents, which utilize nanoparticle-loaded salt carriers for deep impurity removal, show great promise; but high costs and nanoparticle instability currently limit their practical application.

3.1.2. Intermediate Alloy

Intermediate alloys play a vital role in aluminum melt purification by pre-combining active or easily oxidized elements with the aluminum matrix, thereby avoiding challenges associated with direct elemental addition. They help offset compositional losses from flux reactions and enhance product stability.
Moreover, these alloys serve as heterogeneous nucleation sites, promoting grain refinement and reducing melt viscosity, which in turn mitigates casting defects such as shrinkage porosity, segregation, and thermal cracking. Compared to direct refining agents, intermediate alloys offer superior compatibility and efficiency. Common industrial systems include Al–Ti–B, Al–Ti–C, and rare-earth-based alloys.
Titanium is widely used for grain refinement through the formation of TiAl3 intermetallic compounds. When the melt temperature exceeds 665 °C and titanium content is above 0.15%, TiAl3 acts as nucleation sites, significantly enhancing alloy solidification. Yang et al. [34] confirmed the segregation phenomenon of titanium during solidification, with higher concentrations within the boundary of grains than the matrix.
Adding boron improves both impurity control and nucleation. It forms stable high-melting borides (e.g., B4V3, MnB2, FeB2) with trace impurities. Chen et al. [35] reported that Al–B master alloys reduced grain size by over 40% in Al–Si systems, though their effectiveness in Al–Cu and Al–Mg alloys is limited [36]. Table 6 demonstrates the impact of boron on the electrical conductivity of aluminum alloys [37].
Rare-earth elements (REEs) provide new possibilities for intermediate alloy design. Elements such as La and Ce form stable compounds with oxygen and hydrogen, enabling deep melt purification. Their preferential segregation at grain boundaries—5 to 8 times higher than in the matrix—can inhibit harmful FeSi phases. Liu et al. [38] demonstrated that 0.20% scandium reduces grain size by over 50%, while Zeng et al. [39] showed that 0.15% lanthanum increases tensile strength by 18% and boosts conductivity by 5%. Table 7 visually showcases the impact of adding various rare-earth elements on the electrical conductivity and mechanical properties of aluminum alloys [40].
To meet the demands of advanced aluminum alloys, future development of master alloys will focus on optimizing compositions through REE additions, designing dispersion-strengthened systems, and improving processing methods. Although Al-Ti-B and Al-Ti-C alloys remain effective refiners, challenges such as impurity introduction, particle agglomeration, and high additive cost must be addressed through continued innovation.

3.2. Gas Purification Method

Gas purification technology removes hydrogen and non-metallic inclusions from molten aluminum by injecting fine gas bubbles into the melt. The purification mechanism involves the adsorption of hydrogen atoms and micron-sized inclusions onto the rising bubbles, which then carry these impurities to the melt surface for removal [41].
Depending on the type of gas used, the technique can be classified into inert gas refining (N2, Ar), reactive gas refining (Cl2, F2), and mixed-gas systems. Based on the injection method, gas purification is typically categorized into static and rotary injection techniques.
Inert gas refining typically employs nitrogen as the primary medium, with hydrogen-free bubbles introduced into the aluminum melt via a graphite nozzle. According to Sievert’s law, dissolved hydrogen atoms diffuse into the bubbles and rise through the melt, while oxidized inclusions adhere to the bubble surfaces and are subsequently removed.
Nitrogen is widely used in industrial applications due to its low cost and easy availability. However, at temperatures of 700–720 °C, it can react with aluminum to form aluminum nitride (AlN) inclusions, thereby compromising melt cleanliness. To address this issue, researchers have developed argon-based refining systems.
Experimental results [42] indicate that increasing the argon flow rate leads to smaller bubble sizes, higher hydrogen removal efficiency, and improved preservation of the alumina surface layer. Nevertheless, uneven bubble size distribution continues to limit purification efficiency, and issues such as compositional segregation remain significant.
While reactive gas refining improves hydrogen removal efficiency, it is highly corrosive, reducing equipment lifespan and posing occupational health risks. To balance purification performance with environmental and safety concerns, mixed-gas refining has been developed. Typical gas formulations include 15 vol.% Cl2 + 11 vol.% CO + 74 vol.% N2 and 2 vol.% CCl2F2 + 23 vol.% CO2 + 75 vol.% N2 [43].
Qin Yiming et al. [44] experimentally measured the inclusion content following nitrogen and mixed-gas refining and concluded that the mixed-gas approach achieved superior impurity removal. Despite being both efficient and environmentally friendly, mixed-gas refining suffers from inconsistent performance and fails to achieve optimal purification. To address these limitations, researchers have developed the rotary gas injection method.

3.3. Filtration Method

The filtration method achieves the separation of inclusions by passing molten aluminum through a porous medium, effectively removing secondary-phase particles in the size range of 1–30 μm. Common industrial filter materials include glass fiber filters, packed bed filters, and ceramic foam filters. The purification mechanism involves a synergistic combination of mechanical interception and surface adsorption processes [45].
Glass fiber filters employ a braided structure, offering advantages such as low cost, ease of operation, and moderate effectiveness in reducing oxidized inclusion content. However, they are primarily suitable for removing large-particle inclusions and require frequent replacement.
In comparison, packed bed filters consist of 2–8 mm Al2O3 spheres stacked to form a porous medium with enhanced adsorption capacity due to their increased specific surface area [45]. When integrated into in-line purification systems directly connected to melting furnaces, these filters can operate under argon protection to prevent secondary oxidation. Moreover, the filter beds are typically reusable [46].
Despite these advantages, bed filter systems still face limitations, including difficulty in temperature control, challenges in replacing the filter medium, and the formation of dead zones that lead to uneven filtration.
Ceramic foam filters utilize a three-dimensional mesh structure that enables finer purification of molten aluminum. Compared with the SNIF and MCF filtering methods, Wu [42] indicates that CFF can achieve a better impurity removal effect (see Figure 11). Research by Qiu et al. [21] indicates that, under identical functional pore sizes, filters with higher surface roughness exhibit greater filtration efficiency. This improvement is attributed to increased flow resistance caused by the rough interface, which enhances particle adsorption.
Comparative experiments conducted by Schoß et al. [48] demonstrated that ceramic foam filters made of Al2O3 and ZrO2 exhibit superior inclusion removal performance compared to those made of SiC. Although ceramic foam filters are widely applied in the production of standard aluminum alloy wires, secondary refining of ultra-fine wires still requires the use of deep-bed filters.
The filtration method requires regular replacement of filters, which incurs certain cost burdens. Moreover, it is difficult to remove tiny bubbles and impurities. Therefore, it is usually combined with other purification methods to improve overall performance. Its effect is relatively poor when used alone. Currently, the ERAR purification process is gradually evolving towards a mixed purification process that combines multiple purification methods. In the future, the filtration method will also find considerable application in ERAR production.

3.4. Rotary Injection Method

With the development of gas injection technology from the fixed tube type to rotary type, the rotary injection method breaks large bubbles into micro-bubbles with smaller diameter through shearing force generated by a high-speed rotating nozzle. With its large specific surface area, high interface energy, and low buoyancy characteristics, it significantly improves the bubbles’ stability, adsorption capacity, and melt residence time, thus enhancing the purification effect. In industrial applications, researchers have formed diversified and improved processes such as SILF, ALPUR, FILD, RDU, GBF, MINT, the Alcoa segregation technique, Heprojet, and LARSTM by combining with other aluminum alloy purification technologies. On the basis of inheriting the core principle of rotary injection, these methods achieve a breakthrough in performance through structural innovation and process combination, forming a unique technical system.
The SNIF method (Spinning Nozzle Inert Gas Flotation) is a common online processing technique for aluminum melts. It includes configurations such as single-chamber single-rotor, dual-chamber dual-rotor, and triple-chamber triple-rotor systems [49]. Figure 12 and Figure 13 [49] show the diagram of the dual-chamber dual-rotor SNIF device and a schematic of the rotating nozzle, which is the core component of the SNIF system. Its unique structure allows the rotating nozzle to maximize the dispersion of the purging medium through effective stirring. Its principle involves using graphite rotating nozzles to mix inert gas with trace chlorine gas and break it into microbubbles. With melt stirring to promote the uniform diffusion of bubbles, the degassing rate can reach 70% [49].
With the ALPUR method through the special nozzle structure to strengthen the agitation of the melt, so that the bubble and the metal liquid have full contact, degassing efficiency reached 60–75%. The FILD law integrates inert gas refining and flux filtration technology; adopting a two-chamber structure for gas refining and alumina ball filtration in turn, the processing speed can reach three times that of the ALPUR method, and the filter layer has the dual functions of mechanical interception and chemisorption.
As an online processing device developed by ALcoa, although the Alcoa segregation technique requires preheating of the filter bed and periodic replacement of alumina balls (see Figure 14 and Figure 15), it can reduce the concentration of Na to 0.00001% at a processing rate of 23 t/h through double filtration and the injection of refining gas. Table 8 summarizes the typical purity levels achieved during this process, along with the corresponding impurity contents present at each purity grade. RDU method uses a graphite rotor to produce spiral-rising microbubbles (diameter 38 mm), achieving a 95% inclusion removal rate and 0.07 mL/100 g gas impurity control level. The GBF method uses a 700 r/min high-speed nozzle to generate uniform microbubbles, and the degassing rate is 69–73%. Compared to SNIF, which focuses more on the uniformity of bubble distribution, GBF emphasizes the efficient generation of small bubbles. The Heprojet method combined with rotary injection and flux refining can achieve grain refinement while achieving a relatively high degassing rate, and has the advantages of less flux consumption and no pollution. The LARSTM rule, by preheating gas to maintain the melt temperature, in the process of flotation refining controls the hydrogen content in the range of 0.18–0.26 mL/100 g and reduces more than 50% slag content.
Each method continues to innovate in the aspects of bubble control, reaction interface optimization, and equipment structure design, forming a complete technical spectrum covering different purity needs, production conditions, and environmental requirements. Among them, the SNIF and RDU methods are outstanding in the control of ultra-low hydrogen content, FILD and ALPUR rules have advantages in the processing capacity per unit time, while the Heprojet and LARSTM methods focus on environmental protection and energy saving and melt insulation and other specific needs, reflecting the differentiated path of technological development. At present, the rotary jetting technology has been widely applied in the production of high-end aluminum alloys both at home and abroad. Compared with the traditional chlorinated salt flux refining method, the dehydrogenation rate of the former can only reach about 30%, while the rotary jetting technology can achieve over 70%, effectively meeting the purity requirements of high-end aluminum alloys and offering a more environmentally friendly production process. However, in the production process of ERAR, the rotary jetting method has not yet become the mainstream purification method. Nevertheless, in the high-end ERAR field in China, the combined refining process of the rotary jetting and flux methods has begun to be piloted in some areas. It is believed that in the future, the combined refining process of flux and jetting will be more widely used.

3.5. Other Auxiliary Methods

Vacuum degassing refining is based on Sievert’s law and promotes the escape of dissolved gases from aluminum melt by reducing system pressure. Combined with mechanical stirring, the degassing efficiency can be improved. Experimental results [50] indicate that when the rotational speed is maintained below 50 rpm, the adsorption between oxide inclusions and hydrogen bubbles is maximized. However, at speeds exceeding this threshold, increased melt turbulence leads to higher inclusion retention rates. The study by Zhou et al. [51] demonstrated that vacuum stirring exhibits superior purification performance compared to non-vacuum treatment, effectively reducing pore defects and enhancing matrix continuity. Figure 16 [51] illustrates the gas flow behavior during the vacuum refining process.
Ultrasonic refining employs high-frequency elastic waves to generate cavitation effects within the aluminum melt. The resulting cavitation bubbles absorb dissolved hydrogen and promote dendrite fragmentation, thereby enhancing grain refinement. A trace amount of titanium from the transducer dissolves into the melt, forming second-phase intermetallic compounds that further contribute to grain refinement [52]. Through parameter optimization, Zhou et al. [53] demonstrated that ultrasonic treatment at 2500 W and 20 kHz minimizes both hydrogen and oxide content in aluminum alloys. This technology is advancing toward intelligent monitoring, with real-time feedback improving processing accuracy. When optimized in terms of process parameters and technical integration, ultrasonic and vacuum refining can significantly enhance impurity removal efficiency. These techniques may be effectively combined with other purification processes, such as rotary injection and flux treatment, to achieve deep purification of aluminum melts.
Zone refining relies on the differential solubility of impurities in the solid and liquid phases, by which directional separation is achieved through localized melting and resolidification. High-frequency induction heating combined with a moving temperature field can maintain the intrinsic properties of the material, but it needs to be combined with inert gas protection to prevent oxidation. The numerical simulation of the Spim model [54] provides a theoretical basis for equipment improvement. Although effective impurity removal can be achieved under laboratory conditions, its industrial application is still limited by high energy consumption, intermittent operation, and challenges in controlling oxidation, and it is still in the exploratory stage.

4. Cast-Rolling Process

Following purification, the high-purity aluminum melt must still undergo casting and rolling to produce the final ERAR product. While purification is essential for quality control, the subsequent solidification and deformation stages are equally critical. With advances in manufacturing technology, previously distinct casting and rolling stages have increasingly merged into an integrated continuous casting and rolling process. This evolution not only enhances production efficiency but also improves the mechanical and electrical performance of ERAR products.
The aluminum alloy casting and rolling process is a continuous process in which aluminum melt is directly cast and rolled simultaneously. The core principle is that the molten aluminum liquid is rapidly solidified and rolled by a pair of high-speed rotating cooling rolls for aluminum alloy used as wire and cable. There are four general processing and casting methods: the continuous casting and rolling method, the two-roll casting and rolling method, the direct water-cooled semi-continuous casting and extrusion method, and the continuous casting and extrusion process. Each process has its own unique advantages (see Table 9), which will be introduced in turn.

4.1. Continuous Casting and Rolling Process

The continuous casting and rolling process, as a core technology for producing aluminum alloy wire, has become the global mainstream manufacturing method since the launch of the Properzi production line in Italy in 1948. In China, aluminum alloy wire rod is also mainly produced by this method. The typical process includes refining and purifying the electrolytic aluminum melt, continuous casting, and hot continuous rolling, resulting in is an aluminum wire with a diameter of 9.8 mm. Despite the huge scale of domestic application, there is still a gap between the equipment accuracy and process control and the international advanced level. At present, the main equipment types include five-wheel continuous casting machine and horizontal continuous casting machine. The five-wheel continuous casting machine is equipped with a diameter of 1600 mm crystallization wheel and tension wheel system to achieve wheel-belt casting [55]. The horizontal continuous casting machine integrates a tundish, crystallizer, intelligent control system, and other modules to ensure process stability.
At present, the most widely used casting and rolling process is the continuous casting and continuous rolling process. However, the mechanical properties of aluminum alloys are significantly affected by their final heat treatment conditions. Heat treatment is usually carried out after this process to obtain electrical aluminum alloy rods with stronger mechanical properties. Thus, understanding the basic mechanical properties of aluminum alloys under various heat treatment conditions has become a key area of current research [56]. Liu et al. [57] selected a common precipitation hardening alloy, 7075 aluminum alloy (the components are shown in Table 10), for heat treatment research. Analysis of the microstructures presented in Figure 17 and Figure 18 indicates that a higher heat treatment temperature facilitates grain growth, the dissolution of secondary phases, and a more advanced state of recrystallization. Reference [58] also analyzed and studied the heat treatment of 7075 aluminum alloy. It further concluded that for 7075 aluminum alloy plates, a solution treatment at 510 °C for 30 min followed by artificial aging at 120 °C for 24 h can achieve the best stamping effect.
The 6xxx series of aluminum alloys have excellent corrosion resistance and are widely used in many fields. However, compared with 7075 and 2219 aluminum alloys, 6061 aluminum alloy mainly contains Mg and Si(the components are shown in Table 11), while 7050 aluminum alloy mainly consists of Zn and Mg, and 2219 alloy is mainly Cu. Therefore, 6061 aluminum alloy mainly achieves strengthening by forming Mg2Si precipitates, which is different from the strengthening of 7075 aluminum alloy by the η phase (MgZn2) and the θ phase strengthening (Al2Cu) of 2219 aluminum alloy. Thus, there are significant differences in the heat treatment parameters between 6061 and 7075 or 2219 alloys. To study the specific effects of different heat treatment parameters on the properties of 6061 aluminum alloy, Jia et al. [57] designed an orthogonal experiment. The experimental results indicate that increasing the solution temperature within the range of 510 °C to 540 °C enhances the plasticity of the alloy. However, when the temperature reaches 570 °C, significant grain coarsening occurs and grain boundary brittleness increases, resulting in a notable reduction in plasticity. Furthermore, it was observed that prolonging the aging time leads to a decrease in elongation initially, followed by a subsequent increase. The mechanical properties of the aluminum alloy after heat treatment are shown in Table 12. Therefore, heat treatment can effectively improve the mechanical properties of the material.
All the above-mentioned studies have focused on discussing the impact of heat treatment on the strength of aluminum alloys. However, as ERAR materials, they not only require high strength but also have extremely high demands for high electrical conductivity. It is an established fact that the 1xxx, 5xxx, 6xxx, and 8xxx aluminum series have been extensively utilized for electrical applications. In recent decades, in the field of high voltage transmission lines, the 6201 Al alloy as a high-performance all-aluminum alloy conductor (AAAC) is even expected to replace the traditional Al conductor steel rein-forced (ACSR), which usually consists of commercially pure Al alloys, such as 1350 Al alloy, and an iron core. Therefore, Mao et al. [59] conducted experimental research on the 6201aluminum alloy within the 6xxx series of aluminum alloys. The research results show that the process route of under-aged (UA) + RS + re-aged at 160 °C (RA160) can achieve a conductivity of 50.6% IACS and a tensile strength of 363 MPa. The combination of peak aging (PA) + RS + RA160 can achieve an excellent combination of 51.7% IACS and 352 MPa (the tensile properties of the 6201Al alloy in each state are detailed in Table 13). This research has significantly broken through the trade-off relationship between strength and conductivity in traditional processes. Although both studies focused on the 6xxx series of aluminum alloys, Mao’s research was more in-depth compared to that of Jia et al. Mao was not confined to experiments on heat treatment only. By combining rotary swaging (RS) deformation with aging heat treatment, the comprehensive performance of 6201 aluminum alloy wire in terms of strength, electrical conductivity, and ductility was systematically optimized. This research also provides theoretical and technological basis for improving the mechanical and electrical properties of metallic materials through microstructure design in a coordinated manner.
In recent years, most research has focused on short-time heat treatment of aluminum alloys. This study [60] explored the effect of short-time T6 heat treatment (ST6) on the microstructure and mechanical properties of semi-solid formed graphene nanoplatelet (GNP) reinforced A356 aluminum alloy composites. The results showed that compared with conventional T6 treatment, ST6 treatment could more effectively refine the eutectic silicon phase, promote its spheroidization, and inhibit grain growth, while forming more intermetallic compounds. This treatment enabled the composites to achieve higher yield strength and tensile strength, which were 28% and 36% higher than those of T6 treatment, respectively, with slightly better elongation at break. Short-time heat treatment significantly saves energy and time while enhancing strength, making it suitable for industrial fields that require high-strength and lightweight materials. This research achievement provides ideas for the innovative development of new aluminum-based cable materials.
In general, this process has the advantages of moderate equipment investment, high production efficiency, and stable yield [61], but problems such as macro segregation and coarse grains still restrict the improvement of product performance. International studies have shown that increasing the casting speed to 1.2 m/min and increasing the cooling water flow rate by 30% can increase the material elongation by 15% [62]. The process monitoring system developed by Wang Yanan et al. [63] captures real-time parameters such as casting temperature and rolling pressure. With a defect prediction accuracy exceeding 85%, it provides robust data support for process optimization. Future technological development will focus on the development of intelligent control systems and the application of environmentally friendly lubricants, and deeper research on the mechanisms of multi-field coupling in casting and rolling to improve process stability.

4.2. Two-Roll Cast-Rolling Process

By pouring molten aluminum into counter-rotating rolls to achieve rapid solidification, the twin-roll casting and rolling process reduces the number of production steps from 25 to 11, significantly shortening the overall production cycle. This technology offers several advantages, including a highly flexible production line adaptable to multi-specification products, as well as reduced energy consumption and capital investment. The 6082 aluminum alloy rods produced via this method meet the mechanical property requirements specified in EN754-2 [64], and after-heat treatment, comply with EN755-2 [65] standards (The detailed data are presented in Table 14 and Table 15) [66]. Furthermore, compared with the traditional direct extrusion process (the detailed data are presented in Table 16), the Two-Roll Cast-Rolling (TRCR) process significantly enhances metal yield (the product yield can reach 95–97%) and reduces energy consumption (the electricity consumption for smelting has been reduced by approximately 89%). Despite these benefits, central segregation remains the primary technical bottleneck limiting broader adoption. As illustrated in Figure 19 [67], the mechanism involves the formation of columnar crystals under the rapid cooling action of the rolls. During this process, low-melting-point alloying elements tend to migrate toward the center due to dendritic growth, resulting in severe compositional segregation. To address this challenge, researchers both in China and abroad have conducted systematic investigations combining numerical simulations with experimental validation. As a result, the solidification model developed by Yu Wei’s team [67] can predict the critical thresholds for surface and linear segregation successfully.
In addition, the surface quality of aluminum materials produced by traditional twin-roll casting is affected by various process parameters. Cho [68] analyzes from a novel perspective that the spontaneously formed Al coating during the twin-roll casting process can effectively enhance interfacial heat transfer, thereby improving the surface quality of the finished aluminum products. As shown in the SEM image in Figure 20 and Figure 21 [68], the surface quality of the product is significantly improved when covered with an aluminum coating.
In terms of process improvement, domestic studies have confirmed that the application of ultrasonic and electromagnetic composite fields can effectively inhibit segregation [61]. Current technological development focuses on process parameter optimization and intelligent model development, combined with computer simulation to promote the process in the direction of green automation. The twin-roll casting and rolling process can significantly shorten the process cycle and effectively reduce the capital investment of the ERAR production line. Once the segregation prediction and prevention capabilities of the twin-roll casting and rolling process are further strengthened, this process also has the potential for widespread application.

4.3. Direct Water-Cooled Semi-Continuous Casting and Extrusion Process

The production process of direct water-cooled semi-continuous casting—the extrusion process is as follows: electrolytic aluminum liquid → semi-continuous casting → sawing → homogenization treatment → extrusion → φ9.5 mm aluminum alloy wire. This process seamlessly combines semi-continuous casting with hot extrusion and eliminates the intermediate cooling and reheating links in the traditional process, effectively shortening the process cycle [69].

4.3.1. Direct Water-Cooled Semi-Continuous Casting

In direct water-cooled semi-continuous casting (DC casting), the refined aluminum melt is poured into a crystallizer, where partial solidification occurs. The partially solidified ingot is then withdrawn from the crystallization zone by a pulling mechanism and further cooled using a water spray system, enabling continuous production. The equipment configuration is illustrated in Figure 22 [70]. When the cooling rate of the ingot reaches a balance with the traction speed, aluminum with uniform composition can be produced stably. Compared with the traditional casting process, the DC method has the advantages of high production efficiency and a high degree of automation. Engler et al. [71] systematically investigated the microstructural evolution of AA3105 aluminum alloy during direct chill (DC) casting and subsequent homogenization, using a combination of microstructural characterization techniques and numerical simulations. However, this process still suffers from significant drawbacks: large cross-section ingots tend to develop coarse columnar grains due to thermal gradients between the core and surface during solidification, leading to pronounced anisotropy in mechanical properties. To address this challenge, the research team successfully refined the grain structure by introducing physical field intervention methods such as ultrasonic field and electromagnetic stirring [72].
As the core equipment of DC casting, the design optimization of crystallizer plays a decisive role in product quality. At present, the domestic mold research and development is still mainly imitation, and there are technical bottlenecks in heat flow balance control and surface microstructure design. With the advancement of the Internet of Things and big data technologies, intelligent casting systems are now capable of real-time acquisition of process parameters via multisource sensors. By integrating machine learning algorithms, these systems can perform defect prediction, issue early warnings, and autonomously optimize process parameters. The particle method 3D model based on boundary conditions developed in the literature [73] can simulate the influence of shunt geometry and ingot speed on melt flow field. Building upon the thermal-fluid–solid coupling framework, the model proposed by [74] innovatively incorporates an ingot rupture prediction module and employs the discrete particle method (DPM) to address complex boundary conditions. This enhanced model offers a powerful tool for optimizing casting processes under extreme operational conditions.
Compared with traditional casting, DC casting has the advantage of a high degree of automation. However, due to the difficulty in achieving continuous production completely, it has gradually been phased out by the ERAR production process, which is currently technologically complete and mature, in contrast to the continuous casting and continuous rolling process.

4.3.2. Extrusion Process

The aluminum ingots obtained by DC casting need to be homogenized by annealing to eliminate dendrite segregation, then sawn and heated to a specific temperature for extrusion molding. In contrast to rolling, which shapes material through compressive deformation between rotating rolls, extrusion applies unidirectional pressure within a closed chamber to force the aluminum alloy through a die. This process allows full utilization of the alloy’s plasticity and enables the production of high-strength, complex-shaped profiles. As shown in Figure 23 [75], experimental results indicate that when the extrusion deformation reaches 70%, the coarse grains in the as-cast microstructure are fragmented and the grain boundaries become significantly denser. When deformation exceeds 80%, a fibrous structure becomes dominant, resulting in more than an 80% reduction in grain boundary defects and a substantial enhancement in material strength. However, limited by equipment, the extrusion process has segregation problems caused by a low degree of automation and uneven melt flow. The literature [76] confirms that the tensile strength of alloys treated with severe plastic deformation (SPD) technology increases by 40% (see Table 17), showing better mechanical properties. After SPD treatment, the grains are significantly refined and take on a fibrous shape, with a substantial increase in the density of low-angle grain boundaries and dislocations. A large number of dislocations cause the precipitated phase to break and dissolve, significantly enhancing the tensile strength. References [77,78] further found that SPD increased the volume fraction of strengthening phases (the phase composition is Solute clusters with extremely high number density) through severe plastic deformation and formed a submicron grain structure.
Equal Channel Angular Pressing (ECAP), the core technology of Severe Plastic Deformation (SPD), has become a hot topic in the research of ERAR extrusion process due to its unique shear deformation mechanism. Wang et al. [79] reported that, after heat treatment, nanoscale precipitates formed in ECAP-processed aluminum alloy contributed to the inhibition of grain growth. In addition, the increased dislocation density enhanced the diffusion of solute atoms, which led to a 25% reduction in the degree of segregation. Murashkin’s study further showed that the application of ECAP above 100 °C could increase the yield strength of alloy by 50% [80]. Traditional models are usually simple mathematical fittings, lacking connections with micro-mechanisms. Moreover, they can only accurately predict peak stress but fail to take strain into account to precisely predict the rheological stress under complex working conditions, optimize the parameters of thermal processing, provide methodological references for modeling new materials, and explain macroscopic behaviors through microscopic analysis. Eckert [81] established a constitutive equation for hot deformation of X153CrMoV12 steel through the Zener–Hollomon parameter framework and strain compensation technique, which can accurately predict the hot deformation behavior over a wide range with high precision. The core equations are:
ε ˙   =   A [ sin h ( α σ ) ] n   exp Q R T
Z = ε ˙ exp Q R T
where ε ˙ is the strain rate; Q is the deformation activation energy (the energy barrier reflecting material deformation); R and T are the gas constant and temperature, respectively; and α, n, A are the material constant.
This model can also be applied to the ERAR extrusion production process, providing a theoretical tool for it. The Z-parameter model can correlate temperature with the strain rate of aluminum alloys, thereby determining the optimal temperature range for recrystallization, achieving the purpose of refining grains and enhancing the homogeneity of ERAR. Meanwhile, the polynomial model for strain compensation in reference [81] can be extrapolated and extended to the large strain range of SPD to predict the work hardening behavior and guide the controllable preparation of ultrafine-grained structures.
Future technology development will focus on the continuous development of SPD devices, combined with ultrasonic assistance and intelligent temperature control systems, to break the traditional strain limits. Combined with the application of digital twin models, real-time optimization of process parameters can be achieved, promoting the upgrading of extrusion processes towards high precision and low energy consumption. It can be expected that the extrusion process combined with SPD technology and big data models will become a popular research direction in the future, and may even replace the existing mature continuous casting and continuous rolling process.

4.4. Continuous Casting and Extrusion Process

Aluminum alloy continuous casting and extrusion process is an advanced metal integration technology [82], integrating the advantages of casting and extrusion. As a short process technology—compared with the traditional semi-continuous casting—extrusion technology, continuous casting, and the extrusion process are an integrated continuous production; it has significant advantages, such as to shorten the processing cycle and significantly enhance the mechanical properties of the alloy [39,83]. Research findings [83] indicate that both Al-3Fe alloy and pure aluminum exhibit the uncommon phenomenon of simultaneous enhancement in strength and elongation when the rolling deformation exceeds 60%. At 90% deformation, the alloy demonstrates optimal comprehensive mechanical properties (see in Figure 24 [84]). The continuous casting and extrusion process flow chart is shown in Figure 25 [85]. In this process, the melt is injected into the extrusion mold under the condition of solid–liquid coexistence, and the plastic deformation is completed under the action of high pressure and shear force. The casting cooling, sawing, heating, and multiple extrusion processes in the traditional process are omitted, and the energy consumption is reduced by 40–60%. The wire grains obtained by continuous casting and extrusion are fine equiaxed crystals. It is pointed out in the literature [84] that Al-Mg-Si-Cu alloy, under the double action of rapid cooling and frictional shear deformation, forms an equiaxed grain with an average grain size of 43.3 μm after continuous casting and extrusion, which has excellent ductility and good mechanical strength.
With the progress of semi-solid forming technology and intelligent control, the future continuous casting and extrusion process will be combined with integrated digital twin technology to achieve adaptive optimization of process parameters, and promote the development of green and high-end aluminum alloy processing. As an efficient short-process casting and rolling technology, it further enhances the mechanical properties of ERAR on the basis of accelerating the production cycle, demonstrating significant research potential. Once the technology matures in the future, it can greatly reduce the construction cost of ERAR production lines. Compared with traditional continuous casting and continuous rolling lines, this process is more likely to be favored by small metallurgical enterprises.

5. The Prospects of AI-Driven Process Innovation

Innovations and optimizations in purification and continuous casting–rolling processes are clear. However, several limitations remain in the field.
From the perspective of purification, although methods such as flux treatment, master alloy addition, filtration, gas refining, and composite purification have yielded notable results, the experimental process still relies heavily on repetitive trials. Outcomes often exhibit a degree of randomness, and the optimization of process parameters—such as the formulation of new fluxes or master alloys and the control of gas injection rates—typically requires either complex mathematical modeling or lengthy, costly, high-precision experiments.
Similarly, in the casting–rolling process, improvements to equipment dimensions and the optimization of parameters such as cooling rate and rolling speed also demand substantial financial and time investments. Integrating artificial intelligence (AI)—particularly machine learning (ML) and deep learning (DL)–into ERAR design and manufacturing could significantly shorten development timelines, reduce experimental workloads, and lower labor costs.
AI, and especially ML/DL, is reshaping the landscape of functional materials research. As shown in Figure 26 [86], the influence of ML/DL in materials science—from superconducting compounds to high-entropy alloys—is rapidly expanding by revealing composition–structure–property relationships, thereby accelerating materials discovery Current research on electrical round aluminum rod (ERAR) production primarily focuses on [87]. In materials synthesis and processing, AI has evolved from predicting single-point performance to enabling integrated, closed-loop innovation that combines data-driven design, intelligent control, and performance prediction.
Although dedicated AI applications in ERAR remain limited, advances in related aluminum alloy systems provide transferable strategies that can be adapted for ERAR manufacturing.

5.1. AI Innovations in Purification Processes

The purification of ERAR aims to achieve grain refinement and precise compositional adjustment through physicochemical means. ML/DL approaches enable accurate compositional homogenization control, eliminating the need for prolonged computation and experimentation.
For example, in melt purification, Yang et al. [88] employed transfer learning to automate nanoparticle size distribution analysis, greatly improving the efficiency of traditional electron microscopy and paving the way for data-driven optimization of grain structures—an important factor in enhancing ERAR conductivity. Hao et al. [89] proposed a hybrid MTGPR-PSO model to optimize hydrogen removal efficiency under multivariable conditions (e.g., rotor speed, flux addition rate), reducing prediction error to 3.8% and offering a promising path for improving melt treatment consistency and quality.
Reference [90] proposed an explainable AI framework from a novel perspective, achieving precise mapping and mechanism analysis of components, processes, and performance, and breaking through the strong plasticity trade-off. It has broken the traditional 7xxx aluminum alloy design reliance on trial-and-error experiments and local empirical models. The research team conducted 225 sets of experiments, covering 25 alloys and 5 processes. And the strengthening phase and strengthening mechanism were confirmed by TEM. The evolution process of grain structure was analyzed using EBSD. Develop a new large model of artificial intelligence DNN, and at the same time use the LIME algorithm to explain the DNN prediction logic and quantify the contribution of each parameter to YS/TS/EL. The average error MAPE of the test set is 9.93%, and the coefficient of determination R2 is 0.926. This paper provides efficient design tools, promotes the application of high-strength aluminum alloys, and also offers novel ideas for the composition design of aluminum alloys in the ERAR field.
ML can extract valuable insights from large datasets and predict future trends. In the ERAR context, it could be used to forecast material properties, optimize microstructures, and design alloy and refining agent compositions. Such models can rapidly screen potential nanomaterials and predict their properties, significantly shortening R&D cycles and lowering experimental costs.

5.2. AI Applications in Casting-Rolling Processes

The properties of ERAR materials following the casting–rolling process are significantly influenced by factors like porosity [91,92,93]. Being able to accurately forecast these microstructural attributes based on chemical composition and processing conditions—such as cooling rate and initial hydrogen levels—is vital for enhancing process efficiency [94,95,96].
In the context of managing porosity and secondary dendrite arm spacing (SDAS), Chen et al. [97] proposed a cellular automaton (CA) framework to model dendritic solidification in both binary Al-Si and ternary Al-7Si-Mg alloys. The SDAS results obtained from simulated cross-sections were found to align closely with empirical estimates. Nevertheless, as castings grow larger and more complex, these simulation techniques demand greater computational resources and become increasingly expensive. Liu et al. [98] developed a U-Net convolutional neural network to dynamically monitor grain refinement in aluminum alloys, demonstrating that adding yttrium to AlTiB refiners significantly reduced SDAS by 38.9%. This highlights the potential of ML-based monitoring for real-time process control and microstructure optimization of ERAR conductors.
Hou [99] proposed an AI-based framework to predict porosity and SDAS in aluminum alloy castings (the experimental methodology is illustrated in Figure 27). Using 3D CA simulations, a dataset of 472 porosity samples was generated for A356 alloy, covering cooling rates from 0 to 10 /s and initial hydrogen contents from 0.15 to 0.40 mL/g. From these reconstructions, five porosity indicators were extracted. Additionally, a dataset of 310 SDAS samples was compiled from the literature, including alloy composition and cooling rate, with strict quality control.
All data were standardized and split into training and testing sets (80–90% for training). Seven AI models—XGBoost, CatBoost, gradient boosting regression (GBR), TabNet, Transformer, random forest regression (RFR), and support vector regression (SVR)—were evaluated. Hyperparameters were optimized via grid search with five-fold cross-validation. Model performance was assessed using the coefficient of determination (R2), root mean square error (RMSE), mean absolute error (MAE), and relative error analysis. The SHAP (Shapley additive explanation) method was used to interpret the influence of input features (e.g., cooling rate, hydrogen content, alloying elements).
Results showed that XGBoost achieved the highest prediction accuracy for most porosity indicators (R2 up to 0.89) and for SDAS (R2 = 0.95), with prediction errors within 10–15% in experimental validation. CatBoost also performed competitively for porosity percentage prediction. Overall, ensemble learning models outperformed deep learning and traditional regression models for small structured datasets, with inference times of only a few seconds, compared with several hours for CA simulations. SHAP analysis identified cooling rate and initial hydrogen content as the most influential factors for porosity, and cooling rate, strontium (Sr), and silicon (Si) content as most influential for SDAS.
The experimental results mentioned above provide a new perspective for optimizing the ERAR casting and rolling process. The application of artificial intelligence models addresses the most critical shortcomings of traditional trial-and-error experiments. Artificial intelligence can accurately identify intrinsic relationships from the provided data resources, thereby enabling the prediction and optimization of product performance. This significantly reduces human resource requirements, shortens experimental time, and lowers input costs.

5.3. AI Applications in ERAR Product Inspection

Tiwari [100] demonstrated the value of AI in alloy design by combining principal component analysis (PCA) and K-means clustering to classify 50 commercial 6xxx series aluminum alloys based on performance features. This interpretable analysis enables reverse derivation of optimal compositions based on precipitation-hardening theory. Since ERAR often uses related alloy grades, such classification methods can guide intelligent selection of alloying elements to balance conductivity and strength.
AI also offers inherent advantages in product quality inspection [101]. One study developed an automated inspection system based on a Faster R-CNN deep learning model to detect shape defects in aluminum smelter anode rods. The system replaces subjective manual inspection with an automated, near real-time, multi-class defect detection process.
Using SketchUp, the team generated a dataset of 600 images (720 × 1280 pixels) covering five defect classes and one normal class (100 images per class). Images were annotated with LabelImg, producing XML label files, converted into TFRecord format, and split into 80% training and 20% testing sets.
The TensorFlow Faster R-CNN Inception V2 model was configured for six classes with customized label mapping and training parameters [100]. Training began with an initial loss of around 1.6 and continued until the loss stabilized below 0.05 (at approximately 6265 steps).
The trained model achieved accurate multi-class defect detection with near real-time performance. Its modular architecture supports future feature expansion, while its Python (Python 3.7)-based implementation ensures flexibility for integration with other classifiers. This approach shows strong potential for reducing subjectivity and improving cost-effective decision-making in anode rod maintenance.
In summary, the AI research on ERAR will significantly drive the production transformation of ERAR. Opportunities exist in refiner optimization, melt purification, alloy composition control, and casting–rolling process management. Integrating ML models into these domains can enhance process precision, reduce trial-and-error costs, and accelerate the development of high-performance electrical conductor materials.

5.4. Summary and Prospects

Artificial intelligence (AI) has demonstrated transformative potential in overcoming longstanding bottlenecks in the manufacturing of aluminum conductors, particularly in the production of electrical round aluminum rods (ERAR). Conventional purification processes often rely on repetitive experimentation and are characterized by high variability in outcomes. The optimization of refining agent formulations, gas purging parameters, and other process variables typically demands complex modeling or costly experimental procedures. Similarly, controlling cooling rates and rolling speeds during the casting–rolling stage poses challenges related to high capital investment and extended development cycles.
AI technologies, leveraging machine learning (ML) and deep learning (DL), offer a fundamental shift in this landscape. In purification processes, AI enables intelligent and precise control to achieve grain refinement and compositional homogenization, thereby significantly enhancing the electrical conductivity of the resulting conductors. During casting and rolling, AI-based ensemble learning models can accurately predict microstructural features such as porosity in real time-reducing computation times from hours to mere seconds and guide the optimization of cooling rates and alloying strategies. In quality inspection, vision systems based on Faster R-CNN enable real-time automated detection of multiple defect types, replacing subjective manual assessments.
While AI applications specifically tailored to ERAR remain in their early stages, the integration of data-driven design, intelligent process control, and performance prediction has already shown promise in dramatically shortening development cycles, reducing experimental costs, and minimizing human labor. Collectively, AI is poised to revolutionize the entire ERAR production chain—from alloy design and melting purification to casting–rolling and quality inspection. The increasing adoption of ML and DL across these stages will further accelerate the development of high-performance, reliable aluminum conductor materials, enabling the industry to meet the growing technological demands of power transmission and other critical applications.

6. Conclusions

With increasing performance demands in the power transmission industry, driven by technological advancement, optimizing the production process of aluminum alloy rods has become essential for achieving high-quality products. As manufacturing technologies continue to evolve, various production methods and equipment have emerged. Therefore, it is crucial to analyze the advantages and limitations of different processing steps. This paper provides a comparative summary of two representative production processes.
As the most widely used purification method for aluminum alloys, the flux purification process is undergoing significant innovation. Traditional fluoride-based fluxes are being rapidly replaced by new composite flux systems. Currently, the mainstream high-efficiency fluxes are based on fluorine–chlorine systems enhanced with rare-earth additives. Researchers are actively exploring the incorporation of additional thermodynamic components to further improve purification efficiency. Future research will continue to focus on the quantitative optimization of flux compositions and interface reactions. In particular, more attention will be paid to understanding the release mechanisms of atomic-scale rare-earth elements and developing cost-effective industrial production routes. However, the usage of refining agents at this stage will inevitably lead to changes in ERAR’s composition and increase the cost burden. Therefore, in the future, it is necessary to explore more appropriate ways of adding refining agent and balance the relationship between refining effect and cost.
Master alloys have effectively addressed the challenge of directly incorporating refractory and easily oxidized elements during the aluminum casting process. With the usage of master alloy, they contribute to grain refinement and enhance the mechanical strength and toughness of aluminum alloy products. Master alloys also reduce elemental loss and improve the compositional uniformity of the molten aluminum. However, several technical challenges remain. The addition of master alloys may introduce unintended impurity elements, promote the agglomeration of TiB2 particles, and increase production costs by 10–20%. These drawbacks can reduce the electrical conductivity of the final product and lead to defects such as perforation or wire breakage. In addition, extensive industrial usage is still subject to economic restrictions, and is usually used together with refining agents.
Other refining technologies are expected to evolve toward environmentally friendly, energy-efficient, and multifunctional composite purification systems, particularly with advancements in gas purification, rotary gas injection, and filtration techniques. Other emerging techniques such as vacuum refining, ultrasonic treatment, and zone refining also show considerable potential. However, their widespread application is limited by high energy consumption and poor compatibility with continuous production processes. Therefore, it can be foreseen that the future purification strategy may involve hybrid methods and integrate a variety of technologies to achieve efficient impurity removal.
Within the realm of casting–rolling technologies, continuous casting and rolling have reached a highly mature stage and are widely applied throughout the industry. Looking forward, future advancements are expected to focus on integrating physical fields, intelligent control systems, and automated monitoring tools to enhance both product quality and production efficiency.
AI technology has been widely used to guide the production of iron and steel industry. For ERAR aluminum processing industry, the combination of AI can also promote the rapid development of the industry. By optimizing the composition of refining agent through AI design, we can obtain multifunctional and efficient composite refining agent to achieve deep impurity removal. AI can also automatically monitor the production process, simulate production through big data modeling, optimize process parameters, and prevent defects such as segregation.

Author Contributions

Conceptualization, X.L. and J.J.; methodology, X.L. and H.J.; formal analysis, H.J. and J.J.; investigation, X.L. and J.J.; data curation, X.L.; writing—original draft preparation, X.L. and J.J.; writing—review and editing, H.J.; visualization, J.J.; supervision, H.J.; project administration, H.J.; funding acquisition, H.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Science and Technology Program of Guizhou Province (LH[2024]018, [2024]123), the Science and Technology Plan of Guizhou Provincial Department of Education ([2024]001), and the Student Research Training of Guizhou University (2024SRT033).

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ERARElectrician Round Aluminum Rod
CRPCast-Rolling Process
IACSInternational Annealed Copper Standard
UHVUltra-High Voltage
SNIFSpinning Nozzle Inert Flotation
ALPURAluminum Purification
FILDFumeless In-Line Degassing
RDURefining and Degassing Unit
GBFGas Bubbling Flotation
MINTMelt In-Line Treatment System
Alcoa469Melt Purification by ALCOA469 Method
HeprojetHelium Projection
LARSTMLiquid Aluminum Refining System
SPDSevere Plastic Deformation
ESCPEqual Channel Angle Extrusion
DCDirect Water-Cooled Semicontinuous
MLMachine Learning
DLDeep Learning

References

  1. Shen, W.; Hu, A.; Liu, S.; Hu, H. Al-Mn alloys for electrical applications: A review. J. Alloys Metall. Syst. 2023, 2, 2949–9178. [Google Scholar] [CrossRef]
  2. Zhang, A.; Li, Y. Thermal conductivity of aluminum alloys—A review. Materials 2023, 16, 2972. [Google Scholar] [CrossRef] [PubMed]
  3. Zhang, J.; Peng, J. A review on aluminum alloy conductors influenced by alloying elements and thermomechanical treatments: Microstructure and properties. J. Mater. Res. 2023, 38, 1488–1509. [Google Scholar] [CrossRef]
  4. Sims, Z.C.; Rios, O.R.; Weiss, D.; Turchi, P.E.A.; Perron, A.; Lee, J.R.I.; Li, T.T.; Hammons, J.A.; Hansen, M.B.; Willey, T.M.; et al. High performance aluminum–cerium alloys for high-temperature applications. Mater. Horiz. 2017, 4, 1070–1078. [Google Scholar] [CrossRef]
  5. Haridas, R.S.; Agrawal, P.; Thapliyal, S.; Agrawal, P.; Dhal, A.; Shukla, S.; Zhou, L.; Sohn, Y.; Mishra, R.S. Synergy of tensile strength and high cycle fatigue properties in a novel additively manufactured Al-Ni-Ti-Zr alloy with a heterogeneous microstructure. Addit. Manuf. 2023, 62, 103380. [Google Scholar] [CrossRef]
  6. Czerwinski, F. Aluminum alloys for electrical engineering: A review. J. Mater. Sci. 2024, 59, 14847–14892. [Google Scholar] [CrossRef]
  7. Korotkova, N.O.; Belov, N.A.; Timofeev, V.N.; Motkov, M.M.; Cherkasov, S.O. Influence of heat treatment on the structure and properties of an Al–7% REM conductive aluminum alloy casted in an electromagnetic crystallizer. Phys. Met. Metallogr. 2020, 121, 173–179. [Google Scholar] [CrossRef]
  8. Olafsson, P.; Sandstrom, R.; Karlsson, Å. Comparison of experimental, calculated and observed values for electrical and thermal conductivity of aluminum alloys. J. Mater. Sci. 1997, 32, 4383–4390. [Google Scholar] [CrossRef]
  9. Yavaş, A.; Cilingir, C.; Turk, A.; Celik, E. Production and characterization of highly conductive aluminum metal for electric motor applications. J. Mater. Eng. Perform. 2025, 34, 1705–1716. [Google Scholar] [CrossRef]
  10. She, X.; Jiang, X.; Tan, X.; Guo, S. Status and prospect for aluminum industrial development in China. Chin. J. Nonferrous Met. 2020, 30, 709–718. [Google Scholar] [CrossRef]
  11. Levin, A.A.; Narykova, M.V.; Lihachev, A.I.; Kardashev, B.K.; Kadomtsev, A.G.; Brunkov, P.N.; Panfilov, A.G.; Prasolov, N.D.; Sultanov, M.M.; Kuryanov, V.N.; et al. Modification of the Structural, Microstructural, and Elastoplastic Properties of Aluminum Wires after Operation. Metals 2021, 11, 1955. [Google Scholar] [CrossRef]
  12. Levin, A.A.; Narykova, M.V.; Lihachev, A.I.; Kardashev, B.K.; Kadomtsev, A.G.; Panfilov, A.G.; Prasolov, N.D.; Sokolov, R.V.; Brunkov, P.N.; Sultanov, M.M.; et al. Structural, Microstructural, Elastic, and Microplastic Properties of Aluminum Wires (from AAAC (A50) Cables) after Fatigue Tests. Metals 2023, 13, 298. [Google Scholar] [CrossRef]
  13. Sebastian, M. Influence of Non-Uniform Temperature Distribution on Metallic Charge Length on Energy and Force Parameters During Groove-Rolling. J. Iron Steel Res. Int. 2012, 19, 17–24. [Google Scholar] [CrossRef]
  14. Nan, M.; Li, H.; Chen, S. Review of the effects of composition and processing on 8030 aluminum alloy cable conductors. J. Shanxi Univ. Technol. (Nat. Sci. Ed.) 2024, 40, 1–9. [Google Scholar] [CrossRef]
  15. Chen, Y.; Zhang, A. Application and market analysis of aluminum alloy cable. Rare Met. Cem. Carbide 2016, 44, 81–84. [Google Scholar] [CrossRef]
  16. Analysis of China’s Aluminum Alloy Industry Products Tend to Be High-End Under the Policy of Sufficient Supply Side Capacity. Available online: www.chinabaogao.com/market/202401/688881.html (accessed on 10 July 2025).
  17. Investment in Power Grid Construction of State Grid and China Southern Power Grid over the Years. Available online: https://bg.sgpjbg.com/hyshuju/2e56fe116d0325b4459a278b758387d5.html (accessed on 10 July 2025).
  18. SMM. Available online: https://price.smm.cn (accessed on 10 July 2025).
  19. GB/T 3954:2022; Aluminium and Aluminium Alloys Rod for Electrical Purpose. Nation Standardization Administration of the People’s Republic of China: Beijing, China, 2022.
  20. Li, J. Current situation and development of aluminum alloy refining technology. Low-Carbon World 2018, 6, 330–331. [Google Scholar] [CrossRef]
  21. Qiu, Y.; Chen, C.; Li, J.; Wang, L.; Lan, Y.; Zhao, F. Research status and trend of aluminum melt purification. China Metall. 2023, 33, 9–16. [Google Scholar] [CrossRef]
  22. Ma, L. Exploration of common impurity purification technology in aluminum alloy melting process. Shanxi Metall. 2024, 47, 93–94. [Google Scholar] [CrossRef]
  23. Wan, B.; Li, W.; Liu, F.; Lu, T.; Jin, S.; Wang, K.; Yi, A.; Tian, J.; Chen, W. Determination of fluoride component in the multifunctional refining flux used for recycling aluminum scrap. J. Mater. Res. Technol. 2020, 9, 3447–3459. [Google Scholar] [CrossRef]
  24. Cai, Y.; Song, D.; Yang, D.; Zhao, Y.; Ke, B.; Zhang, D.; Zhang, W. Effect of Na3AlF6 contents in refining flux on 3D characteristics of pore and mechanical properties of recycled Al-Mg-Si alloy. Mater. Today Commun. 2024, 39, 108797. [Google Scholar] [CrossRef]
  25. Kulikov, B.; Partyko, E.; Kosovich, A.; Yuryev, P.; Mansurov, Y.; Stepanenko, N.; Baykovskiy, Y.; Bozhko, D.; Durnopyanov, A.; Dombrovskiy, N.; et al. Analysis of composition, properties, and usage efficiency of different commercial salt fluxes for aluminum alloy refining. Metals 2025, 15, 448. [Google Scholar] [CrossRef]
  26. Shi, M.; Li, Y. Performance Improvement in Aluminum Alloy Treated by Salt Flux with Different Fluorides. J. Mater. Eng. Perform. 2023, 32, 3065–3072. [Google Scholar] [CrossRef]
  27. Chen, G.; Huang, S.; Fu, G.; Wu, J.; Chen, R.; Zheng, X. Optimization of impurity removal and purification flux components and their application in the production of A356 aluminum alloy. J. Phys. Conf. Ser. 2024, 2845, 012008. [Google Scholar] [CrossRef]
  28. Yuan, S.; Hao, F.; Zhou, F.; Wang, Z.; Jin, H.; Huang, R.; Gu, W. Current status and prospects of aluminum melt furnace refining process for electrical round aluminum rod. J. Netshape Form. Eng. 2024, 16, 71–78. [Google Scholar] [CrossRef]
  29. Zhao, P.; Jiang, X.; Zhao, Y.; Zhu, Y.; Mu, H.; Cui, T. Use of a new environment-friendly AJ401K smokeless particle refiner. Chongqing China 2021, 325–329. [Google Scholar] [CrossRef]
  30. Tkacheva, O.; Kataev, A.; Rudenko, A.; Dedyukhin, A.; Zaikov, Y. Novel molten salts media for production of functional materials. MATEC Web Conf. 2016, 67, 06044. [Google Scholar] [CrossRef]
  31. Dhaneswara, D.; Fatriansyah, J.F.; Firmansyah, M.R.; Prasetyo, Y. Removal of oxide inclusions in aluminum scrap casting process with sodium based fluxes. MATEC Web Conf. 2019, 269, 07002. [Google Scholar] [CrossRef]
  32. Chen, S.; Jin, Y.; Wei, H.; Li, Z. Study on the effect of rare earth elements La, Ce, Y and Sc doping on the mechanical properties of aluminum–lithium alloys. Appl. Surf. Sci. 2025, 680, 161334. [Google Scholar] [CrossRef]
  33. Zhang, G.; Lu, W.; Wu, X.; Wu, X.; Yang, B.; Tan, Y.; Xu, Z.; Tang, H.; Zeng, J.; Wang, J. A new strategy on designing fluxes for aluminum alloy melt refinement. Materials 2023, 16, 2322. [Google Scholar] [CrossRef]
  34. Yang, H.; Qian, Z.; Chen, H.; Zhao, X.; Han, G.; Du, W.; Nie, X.; Zhao, K.; Liu, G.; Sun, Q.; et al. A new insight into heterogeneous nucleation mechanism of Al by non-stoichiometric TiCx. Acta Mater. 2022, 233, 117977. [Google Scholar] [CrossRef]
  35. Chen, G.; Huang, S.; Fu, G.; Wu, J.; Chen, R. Study on the Effect of Al-Ti-B-RE refiner on the refining effect of A356 aluminum alloy after different melt treatments. J. Phys. Conf. Ser. 2024, 2845, 12039. [Google Scholar] [CrossRef]
  36. Ait, E.H.B.; Hamadellah, A.; Bouayad, A.; El, A.C. Review of grain refinement performance of aluminum cast alloys. Metall. Mater. Eng. 2023, 29, 1–15. [Google Scholar] [CrossRef]
  37. Cui, X.; Wu, Y.; Zhang, G.; Liu, Y.; Liu, X. Study on the improvement of electrical conductivity and mechanical properties of low alloying electrical aluminum alloys. Compos. Part B Eng. 2017, 110, 381–387. [Google Scholar] [CrossRef]
  38. Liu, G.; Ren, Y.; Ma, W.; Morita, K.; Lei, Y.; Zhan, S.; Lv, G.; Li, S.; Zeng, Y.; Li, R. Development process and future trends of chemical refining agents’ influence on grain refinement in aluminum alloys. J. Mater. Res. Technol. 2024, 29, 242–257. [Google Scholar] [CrossRef]
  39. Zeng, P.; Zhang, X.; Zeng, H.; Liu, Y. Influence of rare earth content on microstructure properties of 8030 aluminum alloy. Light Met. 2017, 2, 47–50. [Google Scholar] [CrossRef]
  40. Lin, G.; Li, L.; Guo, Z.; Jia, X.; Wang, X.; Yuan, Z.; Zhang, G.; Zhan, Y.; Shan, Q.; Li, Z. Influence of cerium and yttrium addition on strength and electrical conductivity of pure aluminum alloys. J. Rare Earths 2024, 42, 600–611. [Google Scholar] [CrossRef]
  41. Teng, D.; Zhang, G.; Zhang, S.; Li, J.; Guan, R. Response of mechanical properties of A356 alloy to continuous rheological extrusion Al-5Ti-0.6C-xCe alloy addition upon different Ce contents. J. Mater. Res. Technol. 2023, 27, 209–224. [Google Scholar] [CrossRef]
  42. Al-Alimi, S.; Yusuf, N.K.; Ghaleb, A.M.; Lajis, M.A.; Shamsudin, S.; Zhou, W.; Altharan, Y.M.; Abdulwahab, H.S.; Saif, Y.; Didane, D.H.; et al. Recycling aluminum for sustainable development: A review of different processing technologies in green manufacturing. Results Eng. 2024, 23, 102566. [Google Scholar] [CrossRef]
  43. Du, S.; Zhang, S.; Wang, J.; Lv, Z.; Xu, Z.; Liu, C.; Liu, J.; Liu, B. Sustainable recycling of aerospace-grade ultra-clean 7050 aluminum alloy melts through argon refining without secondary aluminum dross generation. J. Mater. Res. Technol. 2023, 27, 2102–2116. [Google Scholar] [CrossRef]
  44. Qin, Y.; Xu, G.; Li, Z.; Li, Y.; Long, M.; Wu, Q.; Li, J.; Wu, G.; Hu, W.; Chen, Z. Research and application of aluminum alloy refining Technology for automobile. Pop. Sci. Technol. 2023, 25, 58–61. [Google Scholar] [CrossRef]
  45. Wu, J.; Djavanroodi, F.; Gode, C.; Attarilar, S.; Ebrahimi, M. Melt refining and purification processes in Al alloys: A comprehensive study. Mater. Res. Express 2022, 9, 032001. [Google Scholar] [CrossRef]
  46. Hoppach, D.; Peuker, U.A. Increasing cleanliness of Al-melts by addition of ceramic fibers. In Multifunctional Ceramic Filter Systems for Metal Melt Filtration; Aneziris, C.G., Biermann, H., Eds.; Springer Series in Materials Science; Springer: Cham, Switzerland, 2024; Volume 337. [Google Scholar] [CrossRef]
  47. Wu, Y.; Yan, H.; Wang, J.; Zheng, J.; Na, X.; Wang, X. Research on Online Monitoring Technology and Filtration Process of Inclusions in Aluminum Melt. Sensors 2024, 24, 2757. [Google Scholar] [CrossRef]
  48. Schoß, J.P.; Baumann, B.; Keßler, A.; Szucki, M.; Wolf, G. Filtration efficiency in the recycling process of particle-reinforced aluminum alloys using different filter materials. Int. J. Met. 2023, 17, 1681–1696. [Google Scholar] [CrossRef]
  49. Liu, G.; Ren, Y.; Ma, W.; Morita, K.; Lei, Y.; Zhan, S.; Lv, G.; Li, S.; Wang, Z.; Li, R. Recent advances and future trend of aluminum alloy melt purification: A review. J. Mater. Res. Technol. 2024, 28, 2238–7854. [Google Scholar] [CrossRef]
  50. Curtolo, D.C.; Xiong, N.; Friedrich, S.; Friedrich, B. High- and ultra-high-purity aluminum, a review on technical production methodologies. Metals 2021, 11, 1407. [Google Scholar] [CrossRef]
  51. Zhou, W.; Li, J.; Bian, Y.; Han, X.; Guan, R. Exploration of deep purification of aluminum alloy in vacuum based on adsorption of oxides by hydrogen bubbles. Vacuum 2024, 229, 113594. [Google Scholar] [CrossRef]
  52. Jaime, R.F.; Puga, H.; Prokic, M.; Söderhjelm, C.; Apelian, D. Fundamentals of Ultrasonic Treatment of Aluminum Alloys. Int. J. Met. 2024, 18, 2783–2807. [Google Scholar] [CrossRef]
  53. Zhou, W.; Li, J.; Bian, Y.; Han, X.; Jiang, J.; Guan, R. The impact of frequency and power on the ultrasonic purification of aluminum alloy. Ultrason. Sonochem. 2024, 109, 107006. [Google Scholar] [CrossRef]
  54. Wan, H.; Xu, B.; Yang, B.; Zhao, J.; Dai, Y. The impurities distribution and purification efficiency of high-purity aluminum preparation by zone melting in vacuum. Vacuum 2020, 171, 108839. [Google Scholar] [CrossRef]
  55. Li, L.; Chen, Q.; Luo, F.; Wang, X.; Zhu, Q.; Zhao, Z. Structure of the moderate-strength aluminum alloy for overhead conductors during the rotary-wheel casting process. Light Alloy Fabr. Technol. 2015, 43, 28–31. [Google Scholar] [CrossRef]
  56. Li, J.; Wang, Z.; Bai, J.; Xue, H.; Tian, N.; Zhao, Z.; Qin, G. Research progress on microstructure tuning of heat-resistant cast aluminum alloys. J. Mater. Sci. 2024, 59, 10035–10057. [Google Scholar] [CrossRef]
  57. Liu, X.; Wang, C.; Li, X.; Zhu, B.; Li, R. Investigation of 7075 aluminum alloy TIG welding joint using 7075 aluminum alloy wire before and after heat treatment. Mater. Res. Express 2023, 10, 046512. [Google Scholar] [CrossRef]
  58. Jia, X.; Fan, Z.; Luo, Z.; Hu, G.; Zhang, H. Impact of Heat Treatment Parameters on the Plastic Properties of 6061 Aluminum Alloy. Materials 2025, 18, 1705. [Google Scholar] [CrossRef] [PubMed]
  59. Mao, Q.; Wang, L.; Nie, J.; Zhao, Y. Optimizing strength and electrical conductivity of 6201 aluminum alloy wire through rotary swaging and aging processes. J. Mater. Process. Technol. 2024, 331, 118497. [Google Scholar] [CrossRef]
  60. Md Ali, A.; Omar, M.Z.; Salleh, M.S.; Mohamed, I.F.; Hashim, H.; Wakhi Anuar, N.F.B. Effect of Heat Treatment on the Microstructures and Mechanical Properties of Thixoformed Graphene-reinforced Aluminium Alloy Composite. Int. J. Met. 2024, 18, 1695–1709. [Google Scholar] [CrossRef]
  61. Yu, G.; Wu, Z. Manufacturing technologies and quality controls of 8030 aluminum alloy rod. Electr. Wires Cables 2017, 2, 12–16+46. [Google Scholar] [CrossRef]
  62. Jia, T.; Xu, Z.; Tian, S.; Qian, A.; Chen, Y.; Chen, Z. Technological research status and development trend of horizontal direct casting of aluminum alloy. Non-Ferr. Met. Process. 2025, 54, 7–13. [Google Scholar] [CrossRef]
  63. Wang, Y.; Ma, X. Research on casting properties and relative parameters database of aluminum alloy. Cast 2024, 73, 514–519. [Google Scholar] [CrossRef]
  64. EN 754-2; Aluminium and Aluminium Alloys—Cold Drawn Rod/Bar and Tube. European Committee for Standardization: Brussels, Belgium, 2016.
  65. EN 755-2; Aluminium and Aluminium Alloys—Extruded Rod/Bar, Tube and Profiles. European Committee for Standardization: Brussels, Belgium, 2016.
  66. Sidelnikov, S.; Galiev, R.; Bersenev, A.; Voroshilov, D. Application and research twin roll casting-extruding process for production longish deformed semi-finished products from aluminum alloys. Mater. Sci. Forum 2018, 918, 13–20. [Google Scholar] [CrossRef]
  67. Wang, J.; Fu, J.; Cong, F.; Ren, W.; Wu, Y.; Li, G.; Zhang, X.; Wang, F.; Li, Y.; Liu, C. Progress and development trend of aluminum alloy twin-roll casting process. Light Alloy Fabr. Technol. 2024, 52, 1–10. [Google Scholar] [CrossRef]
  68. Cho, H.-G.; Kim, Y.D.; Kim, M.-S. Effect of in situ Al roll coating on Strip Surface quality in traditional twin-roll casting of aluminum alloys. Metals 2025, 15, 377. [Google Scholar] [CrossRef]
  69. Zhang, H.; Cheng, X.; Chen, C.; Lan, S.; Wang, R.; Fan, Y.; Zhang, Y.; Dang, P. Microstructure and properties of Al-0.65Mg-0.58Si-0.17Fe alloy prepared by horizontal continuous casting. Min. Metall. Eng. 2024, 44, 174–178. [Google Scholar] [CrossRef]
  70. Guo, X.; Zhao, Z. The current research status of aluminum alloy welding wire production. Light Alloy Fabr. Technol. 2023, 51, 1–6. [Google Scholar] [CrossRef]
  71. Engler, O. Microchemistry evolution during industrial DC casting and subsequent homogenization of aluminum alloy AlMn0.5Mg0.5 (AA 3105). Mater. Today Commun. 2023, 36, 106724. [Google Scholar] [CrossRef]
  72. Duan, W.; Li, J.; Pan, Y.; Han, F.; Zhu, X.; Mao, H.; Zhou, G.; Lang, Y.; Wang, S.; Zhang, Z. Development of semi-continuous casting technology under external physical field. Spec. Cast. Nonferrous Alloys 2024, 44, 745–752. [Google Scholar] [CrossRef]
  73. Baserinia, A.R.; Ng, H.; Weckman, D.C.; Weiis, M.A.; Barker, S.; Gallerneault, M. A simple model of the mold boundary condition in direct-chill (DC) casting of aluminum alloys. Metall. Mater. Trans. B 2012, 43, 887–901. [Google Scholar] [CrossRef]
  74. Tokunaga, H.; Motoyama, Y.; Okane, T. Particle method simulation of direct-chill casting process including breakout. Int. J. Adv. Manuf. Technol. 2023, 129, 4407–4418. [Google Scholar] [CrossRef]
  75. Zhou, X.; Tian, Z.; Liu, Y. Rolling deformation and microstructure properties on 7050 aluminum alloy under different rolling deformation amounts. Forg. Stamp. Technol. 2024, 49, 80–87. [Google Scholar] [CrossRef]
  76. Jin, H.; Guan, R.; Tie, D. Mechanical and conductive performance of aged 6xxx aluminum alloy during rotary swaging. Crystals 2022, 12, 530. [Google Scholar] [CrossRef]
  77. Sun, W.; Zhu, Y.; Marceau, R.; Wang, L.; Zhang, Q.; Gao, X.; Hutchinson, C. Precipitation strengthening of aluminum alloys by room-temperature cyclic plasticity. Science 2019, 363, 972–975. [Google Scholar] [CrossRef]
  78. Zhao, D.; Qin, J.; Lü, S.; Li, J.; Guo, W.; Wu, S.; Zhang, Y. Mechanisms of extrusion-ratio dependent ultrafine-grain fabrication in AZ31 magnesium alloy by continuous squeeze casting-extrusion. J. Mater. Process. Technol. 2022, 310, 117755. [Google Scholar] [CrossRef]
  79. Afifi, M.A.; Wang, Y.C.; Pereira, P.H.R.; Wang, Y.; Li, S.; Huang, Y.; Langdon, T.G. Characterization of precipitates in an Al-Zn-Mg alloy processed by ECAP and subsequent annealing. Mater. Sci. Eng. A 2018, 712, 146–156. [Google Scholar] [CrossRef]
  80. Murashkin, M.; Medvedev, A.; Kazykhanov, V.; Krokhin, A.; Raab, G.; Enikeev, N.; Valiev, R.Z. Enhanced mechanical properties and electrical conductivity in ultrafine-grained Al 6101 alloy processed via ECAP-conform. Metals 2015, 5, 2148–2164. [Google Scholar] [CrossRef]
  81. Krbaťa, M.; Eckert, M.; Križan, D.; Barényi, I.; Mikušová, I. Hot Deformation Process Analysis and Modelling of X153CrMoV12 Steel. Metals 2019, 9, 1125. [Google Scholar] [CrossRef]
  82. Wu, S. Study on Microstructure and Properties of Al-Zr Alloy with Gas Assisted Continuous Casting. Master’s Thesis, Kunming University of Science and Technology, Kunming, China, 2024. [Google Scholar] [CrossRef]
  83. Chen, J.; Liu, C.; Guan, R.; Wen, F.; Zhou, Q.; Zhao, H. Improving the comprehensive mechanical property of the rheo-extruded Al-Fe alloy by severe rolling deformation. J. Mater. Res. Technol. 2020, 9, 1768–1779. [Google Scholar] [CrossRef]
  84. Li, J.; Gao, M.; Yang, L.; Shen, C.; Guan, R. Effect of cold drawing on the microstructure and property of precipitation strengthening Al–Mg–Si–Cu alloy prepared by continuous casting-extrusion. Mater. Sci. Eng. A 2025, 922, 147669. [Google Scholar] [CrossRef]
  85. Zhou, T.; Zhang, X.; Chen, C. Experimental Study on continuous casting and extrusion forming technology of high electrical conductivity round aluminum rods. J. Shenyang Univ. (Nat. Sci. Ed.) 2012, 24, 31–33. [Google Scholar] [CrossRef]
  86. Han, X.; Wang, X.; Xu, M.; Feng, Z.; Yao, B.; Guo, P.; Gao, Z.; Lu, Z. AI-driven inverse design of materials: Past, present, and future. Chin. Phys. Lett. 2025, 42, 27403. [Google Scholar] [CrossRef]
  87. Meng, S.; Shi, Z.; Peng, M.; LI, G.; Zheng, H.; Liu, L.; Zhang, L. Landslide displacement prediction with step-like curve based on convolutional neural network coupled with bi-directional gated recurrent unit optimized by attention mechanism. Eng. Appl. Artif. Intell. 2024, 133, 108078. [Google Scholar] [CrossRef]
  88. Yang, L.; Wang, H.; Leng, D.; Fang, S.; Yang, Y.; Du, Y. Machine learning applications in nanomaterials: Recent advances and future perspectives. Chem. Eng. J. 2024, 500, 156687. [Google Scholar] [CrossRef]
  89. Hao, M.; Yu, Q.; Wei, C.; Chen, Y.; Chai, L.; Ge, Y. Statistical determination of Johnson-Cook model parameters for porous materials by machine learning and particle swarm optimization algorithm. J. Mater. Eng. Perform. 2022, 31, 7176–7190. [Google Scholar] [CrossRef]
  90. Park, S.; Kayani, S.; Euh, K.; Seo, E.; Kim, H.; Park, S.; Yadav, B.; Park, S.; Sung, H.; Jung, I. High strength aluminum alloys design via explainable artificial intelligence. J. Alloys Compd. 2022, 903, 163828. [Google Scholar] [CrossRef]
  91. Houria, M.; Nadot, Y.; Fathallah, R.; Roy, M.; Maijer, D. Influence of casting defect and SDAS on the multiaxial fatigue behaviour of A356-T6 alloy including mean stress effect. Int. J. Fatigue 2015, 80, 90–102. [Google Scholar] [CrossRef]
  92. Yang, B.; Chen, S.; Song, H.; Zhang, S.; Chang, H.; Xu, S.; Zhu, Z.; Li, C. Effects of microstructure coarsening and casting pores on the tensile and fatigue properties of cast A356-T6 aluminum alloy: A comparative investigation. Mater. Eng. A 2022, 857, 144106. [Google Scholar] [CrossRef]
  93. Han, J.; Dong, G.; Li, S.; Zheng, J.; Wang, J.; Li, H.; Mikhail, D.; Bi, J. Uneven distribution of cooling rate, microstructure and mechanical properties for A356-T6 wheels fabricated by low pressure die casting. J. Manuf. Process. 2024, 127, 196–210. [Google Scholar] [CrossRef]
  94. Li, X.; Yang, X.; Xue, C.; Wang, S.; Zhang, Y.; Li, Q.; Miao, Y.; Wang, J. Hydrogen microporosity evolution and dendrite growth during long solidification of Al-Cu-Li alloys: Modeling and experiment. J. Mater. Process. Technol. 2023, 321, 118135. [Google Scholar] [CrossRef]
  95. Zhang, A.; Guo, Z.; Jiang, B.; Du, J.; Wang, C.; Huang, G.; Zhang, D.; Liu, F.; Xiong, S.; Pan, F. Multiphase and multiphysics modeling of dendrite growth and gas porosity evolution during solidification. Acta Mater. 2021, 214, 117005. [Google Scholar] [CrossRef]
  96. Xue, C.; Mao, P.; Liu, C.; Guo, Y.; Qian, F.; Zhang, C.; Liu, K.; Zhang, M.; Tang, S.; Wang, J. Improving mechanical properties of wire arc additively manufactured AA2196 Al-Li alloy by controlling solidification defects. Addit. Manuf. 2021, 43, 102019. [Google Scholar] [CrossRef]
  97. Chen, R.; Xu, Q.; Liu, B. Cellular automaton simulation of three-dimensional dendrite growth in Al-7Si-Mg ternary aluminum alloys. Comput. Mater. Sci. 2015, 105, 90–100. [Google Scholar] [CrossRef]
  98. Liu, X.; Wang, B.; Li, Q.; Wang, J.; Xue, C.; Yang, X.; Tian, G.; Chang, X.; Liu, X. Quantifying the effects of grain refiners (AlTiB and Y) on microstructure and properties in W319 alloys. Mater. Today Commun. 2022, 33, 104671. [Google Scholar] [CrossRef]
  99. Hou, Q.; Wu, X.; Li, Z.; Feng, S.; Kong, D.; Wang, S.; Ma, X.; Miao, Y.; Qiao, H.; Li, X.; et al. Artificial intelligence enabled microstructure prediction in Al alloy castings. J. Mater. Sci. Technol. 2026, 241, 21–34. [Google Scholar] [CrossRef]
  100. Tiwari, T.; Jalalian, S.; Mendis, C.; Eskin, D. Classification of T6 Tempered 6XXX Series Aluminum Alloys Based on Machine Learning Principles. JOM 2023, 75, 4526–4537. [Google Scholar] [CrossRef]
  101. Chehri, H.; Chehri, A.; Kiss, L.; Zimmerman, A. Automatic Anode Rod Inspection in Aluminum Smelters using Deep-Learning Techniques: A Case Study. Procedia Comput. Sci. 2020, 176, 3536–3544. [Google Scholar] [CrossRef]
Figure 1. Chinese market size of aluminum alloy cable in the past six years Reprinted from ref. [16].
Figure 1. Chinese market size of aluminum alloy cable in the past six years Reprinted from ref. [16].
Metals 15 00981 g001
Figure 2. Recent Chinese state grid investment trends Reprinted from ref. [17].
Figure 2. Recent Chinese state grid investment trends Reprinted from ref. [17].
Metals 15 00981 g002
Figure 3. Average price of electrician round aluminum rods in the Chinese market Reprinted from ref. [18].
Figure 3. Average price of electrician round aluminum rods in the Chinese market Reprinted from ref. [18].
Metals 15 00981 g003
Figure 4. Schematic of inclusion removal by molten salt flux: (a) Molten salts are adsorbed onto oxide film; (b) Continuous oxide film is destroyed and stripped by fluxes; (c) Discrete oxide films are wetted by molten salt Reprinted from ref. [23].
Figure 4. Schematic of inclusion removal by molten salt flux: (a) Molten salts are adsorbed onto oxide film; (b) Continuous oxide film is destroyed and stripped by fluxes; (c) Discrete oxide films are wetted by molten salt Reprinted from ref. [23].
Metals 15 00981 g004
Figure 5. Schematic of inclusion removal by molten salt flux: (a) Refining flux adding; (b) Interaction among flux, hydrogen, and inclusion; (c) Agglomeration of hydrogen and inclusion; (d) Formation of pores Reprinted with permission from ref. [24]. 2024 Elsevier.
Figure 5. Schematic of inclusion removal by molten salt flux: (a) Refining flux adding; (b) Interaction among flux, hydrogen, and inclusion; (c) Agglomeration of hydrogen and inclusion; (d) Formation of pores Reprinted with permission from ref. [24]. 2024 Elsevier.
Metals 15 00981 g005
Figure 6. XRD patterns of the investigated salt fluxes: (a) SF1; (b) SF2; (c) SF3; (d) SF4; (e) SF5 Reprinted from ref. [25].
Figure 6. XRD patterns of the investigated salt fluxes: (a) SF1; (b) SF2; (c) SF3; (d) SF4; (e) SF5 Reprinted from ref. [25].
Metals 15 00981 g006
Figure 7. Phase composition diagram of the investigated salt fluxes (wt.%) Reprinted from ref. [25].
Figure 7. Phase composition diagram of the investigated salt fluxes (wt.%) Reprinted from ref. [25].
Metals 15 00981 g007
Figure 8. Thermograms of the investigated salt fluxes: (a) SF1; (b) SF2; (c) SF3; (d) SF4; (e) SF5 Reprinted from ref. [25].
Figure 8. Thermograms of the investigated salt fluxes: (a) SF1; (b) SF2; (c) SF3; (d) SF4; (e) SF5 Reprinted from ref. [25].
Metals 15 00981 g008
Figure 9. Variation of interfacial tension between flux and alumina at 740 °C as a function of additives in equimolar NaCl–KCl salt solution Reprinted from ref. [23].
Figure 9. Variation of interfacial tension between flux and alumina at 740 °C as a function of additives in equimolar NaCl–KCl salt solution Reprinted from ref. [23].
Metals 15 00981 g009
Figure 10. Coalescence of aluminum droplets in NaCl–KCl–0.3 mass% KAlF4 salt solution: (a) 1 min; (b) 10 min; (c) 60 min; (d) 80 min Reprinted from ref. [23].
Figure 10. Coalescence of aluminum droplets in NaCl–KCl–0.3 mass% KAlF4 salt solution: (a) 1 min; (b) 10 min; (c) 60 min; (d) 80 min Reprinted from ref. [23].
Metals 15 00981 g010
Figure 11. The average concentration distribution of inclusions of each size post-SNIF, post-CFF, and post-MCF Reprinted from ref. [47].
Figure 11. The average concentration distribution of inclusions of each size post-SNIF, post-CFF, and post-MCF Reprinted from ref. [47].
Metals 15 00981 g011
Figure 12. SNIF for melt treatment Reprinted from ref. [49].
Figure 12. SNIF for melt treatment Reprinted from ref. [49].
Metals 15 00981 g012
Figure 13. Swivel nozzle in the SNIF method Reprinted from ref. [49].
Figure 13. Swivel nozzle in the SNIF method Reprinted from ref. [49].
Metals 15 00981 g013
Figure 14. Sketch representation of the Alcoa segregation technique Reprinted from ref. [50].
Figure 14. Sketch representation of the Alcoa segregation technique Reprinted from ref. [50].
Metals 15 00981 g014
Figure 15. Schematic of the rotary degassing process Reprinted from ref. [49].
Figure 15. Schematic of the rotary degassing process Reprinted from ref. [49].
Metals 15 00981 g015
Figure 16. Movement trend of bubbles and oxides in vacuum melting stirring process Reprinted with permission from ref. [51]. 2024 Elsevier.
Figure 16. Movement trend of bubbles and oxides in vacuum melting stirring process Reprinted with permission from ref. [51]. 2024 Elsevier.
Metals 15 00981 g016
Figure 17. EBSD images of HAZ: (a) 160 A As-Welded; (b) 160 A Heat-Treated; (c) 180 A As-Welded; (d) 180 A Heat-Treated; (e) the average grain size variation trend under different currents Reprinted from ref. [57].
Figure 17. EBSD images of HAZ: (a) 160 A As-Welded; (b) 160 A Heat-Treated; (c) 180 A As-Welded; (d) 180 A Heat-Treated; (e) the average grain size variation trend under different currents Reprinted from ref. [57].
Metals 15 00981 g017
Figure 18. Microstructure recrystallization of HAZ: (a) 160 A As-Welded; (b) 160 A Heat-Treated; (c) 180 A As-Welded; (d) 180 A Heat-Treated; (e) recrystallization distribution trend diagram under different currents Reprinted from ref. [57].
Figure 18. Microstructure recrystallization of HAZ: (a) 160 A As-Welded; (b) 160 A Heat-Treated; (c) 180 A As-Welded; (d) 180 A Heat-Treated; (e) recrystallization distribution trend diagram under different currents Reprinted from ref. [57].
Metals 15 00981 g018
Figure 19. Schematic diagram of center segregation formation Reprinted from ref. [67].
Figure 19. Schematic diagram of center segregation formation Reprinted from ref. [67].
Metals 15 00981 g019
Figure 20. Appearance of strip surface Reprinted from ref. [68].
Figure 20. Appearance of strip surface Reprinted from ref. [68].
Metals 15 00981 g020
Figure 21. SEM images of the TRC strip surfaces: (a) no coating condition and (b) Al-coated condition Reprinted from ref. [68].
Figure 21. SEM images of the TRC strip surfaces: (a) no coating condition and (b) Al-coated condition Reprinted from ref. [68].
Metals 15 00981 g021
Figure 22. Schematic diagram of DC casting Reprinted from ref. [70].
Figure 22. Schematic diagram of DC casting Reprinted from ref. [70].
Metals 15 00981 g022
Figure 23. Tensile properties of 7050 aluminum alloy under different rolling deformation Reprinted from ref. [75].
Figure 23. Tensile properties of 7050 aluminum alloy under different rolling deformation Reprinted from ref. [75].
Metals 15 00981 g023
Figure 24. The mechanical property of the Al-3Fe alloy and the pure Al with different deformation reductions: (a) tensile strength; (b) the elongation Reprinted with permission from ref. [84]. 2025 Elsevier.
Figure 24. The mechanical property of the Al-3Fe alloy and the pure Al with different deformation reductions: (a) tensile strength; (b) the elongation Reprinted with permission from ref. [84]. 2025 Elsevier.
Metals 15 00981 g024
Figure 25. Casting extrusion molding diagram Reprinted from ref. [85]: 1—Melting furnace (aluminum melt); 2—holding furnace; 3—casting system; 4—casting extrusion wheel; 5—squeeze boots; 6—stop block; 7—Mold; 8—cooling system; 9—aluminum rods; 10—coiler.
Figure 25. Casting extrusion molding diagram Reprinted from ref. [85]: 1—Melting furnace (aluminum melt); 2—holding furnace; 3—casting system; 4—casting extrusion wheel; 5—squeeze boots; 6—stop block; 7—Mold; 8—cooling system; 9—aluminum rods; 10—coiler.
Metals 15 00981 g025
Figure 26. AI-driven inverse design of materials Reprinted from ref. [86].
Figure 26. AI-driven inverse design of materials Reprinted from ref. [86].
Metals 15 00981 g026
Figure 27. Computational flow of the constructed prediction model: (a) The dataset for CA simulation and literature data collection; (b) Preprocess and optimize the prepared data; (c) Carry out model training and comparison for seven AI algorithms; (d) Model validation and interpretation Reprinted with permission from ref. [99]. 2026 Elsevier.
Figure 27. Computational flow of the constructed prediction model: (a) The dataset for CA simulation and literature data collection; (b) Preprocess and optimize the prepared data; (c) Carry out model training and comparison for seven AI algorithms; (d) Model validation and interpretation Reprinted with permission from ref. [99]. 2026 Elsevier.
Metals 15 00981 g027
Table 1. Performance comparison of different electrical conductors Reprinted from refs. [14,15].
Table 1. Performance comparison of different electrical conductors Reprinted from refs. [14,15].
Type of Electrical ConductorDensity
/(g/mm3)
Melting Point
/°C
Conductivity
/(%IACS)
RTC/(°C−1)TS/MPaYS/MPaE/%
8030 2.766061.84.03113.853.930
Cu8.8910831003.93220–27060–8030–45
Al2.7660624.0370–11020–3023–25
Note: RTC = Resistance temperature coefficient; TS = Tensile strength; YS = Yield strength; E = Elongation; 8030 = 8030 aluminum alloy electrical conductor; Cu = Copper electrical conductor; Al = Aluminum electrical conductor.
Table 2. Mechanical and electrical properties required by GB/T 3954-2022 Reprinted from ref. [19].
Table 2. Mechanical and electrical properties required by GB/T 3954-2022 Reprinted from ref. [19].
Brand NumberStatusT/MPaE/%Resistivity/nΩ·m
1B90, 1B93, 1B95, 1B97O35~653527.15
H1460~901527.25
1B85H14105~130527.95
1370O60~952527.90
H1285~1151128.01
H13105~135828.03
H14115~150628.05
H16130~160528.08
1A60, 1R50O60~902527.55
H1280~1101327.85
H1395~1151128.01
H14110~130828.01
H16120~150628.01
1R60O95~130828.03
H15100~135728.34
1350O60~952527.90
H1285~1151228.03
H14105~1351028.08
H16120~150828.12
6101T4150~2001034.50
6201T4160~2201034.50
8A07H1595~135728.64
H17120~160631.25
8030H14105~1551029.73
8E76O60~1002528.00
H1285~1251328.45
8R76H1395~1201128.78
H14105~140828.89
H16115~150629.00
Note: T = Tensile strength under room temperature testing; E = Elongation after breaking under room temperature testing; GB/T 3954-2022 = the latest version of the National Standard of the People’s Republic of China, “Electrical Aluminum alloy Rod”.
Table 3. Comparison of common aluminum conductor standards in Europe, America.
Table 3. Comparison of common aluminum conductor standards in Europe, America.
Compare DimensionsEuropean Aluminum Conductor GradesAmerican Aluminum Conductor Grades
Main ingredientsBasically consistentBasically consistent
Applicable marketEuropean projectWorldwide
Testing key pointDimensional accuracy of wireAnnealing Process
Table 4. Typical impurities in aluminum conductors Reprinted from refs. [21,22].
Table 4. Typical impurities in aluminum conductors Reprinted from refs. [21,22].
Impurity TypeIngredientsMorphologyDensity/(g/cm3)Size/(μm)Mass Fraction Sources
GasH2Atom 0.00002~0.00006Raw materials, melting environment, chemical reactions
OxidesAl2O3Granules, membranes3.97Granules:
0.2~30
Membrane: 10–5000
0.0006~0.0016Oxidation reaction
High-temperature melting
Casting
Refractory
MgOGranules, membranes3.58Pellets:
0.1~5.0
Membrane: 10~5000
MgAl2O4Granules, membranes3.60Pellets:
0.1~5.0
Membrane: 10~5000
SiO2Pellets2.660.5~5.0
SaltFluoride, chlorideParticle1.98–2.160.1~5.0<0.0001Chloride, cryolite electrolyte, etc.
CarbidesAl4C3Pellets2.360.5~25.00.0002~0.0012High-temperature reaction
SiCPellets3.22 High-temperature reaction
NitridesAlNGranules, membranes3.2510~500.0003~0.0012High-temperature reaction
BoridesTiB2Particle4.501~30<0.0001Grain refiner
AlB2Pellets3.190.1~3.00.0001~0.0100Grain refiner used incorrectly
Intermetallic compoundsAl (FeMnCr) SiGranules>4.001~50 Melting process tool
Table 5. Comparison of different types of aluminum alloy refining agents Reprinted from refs. [24,26,27,29,30,33].
Table 5. Comparison of different types of aluminum alloy refining agents Reprinted from refs. [24,26,27,29,30,33].
Refining Agent TypeHydrogen Removal Capacity/(mL/100 g Al)Impurity Removal CapacityCost EstimationEnvironmental FriendlyBasis
Traditional chlorinated salt refining agent0.15~0.2040%~60%LowLarge amounts of smoke and slagNaCl, KCl, KF, KAlF4
New type of compound refining agent≤0.1270%~90%MediumThe amount of flue gas and slag is significantly reducedK2SO4, NaNO3, KF, K2CO3, RE
Nano-composite refining agent≤0.10>85%HighLow slag output, high energy consumptionC2Cl6, Na2CO3, nano-Al2O3
Table 6. Electrical conductivity of the alloys with boron treatment (%IACS) Reprinted with permission from ref. [37]. 2017 Elsevier.
Table 6. Electrical conductivity of the alloys with boron treatment (%IACS) Reprinted with permission from ref. [37]. 2017 Elsevier.
SampleB Addition/%
00.030.060.090.120.24
Al-0.5Mg-0.35Si52.257.054.353.955.058.7
Al-0.5Fe-0.2Si53.054.156.656.958.357.4
Al-0.8Fe-0.2Cu56.257.058.058.359.858.7
Table 7. Electrical conductivity, ultimate tensile strength, yield strength, and elongation of different chemical compositions in the as-annealed state and comparison of their properties with those of other works Reprinted with permission from ref. [40]. 2024 Elsevier.
Table 7. Electrical conductivity, ultimate tensile strength, yield strength, and elongation of different chemical compositions in the as-annealed state and comparison of their properties with those of other works Reprinted with permission from ref. [40]. 2024 Elsevier.
Alloy Composition/wt.%Electrical Conductivity/%IACSUltimate Tensile Strength/MPaYield Strength/MPaElongation/%
Pure Al59.9–60.391.485.413.6
Al-0.2Ce60.3–61.2100.394.714.2
Al-0.2Ce-0.1Y62.9–63.3106.4101.314.3
Al-0.25Zr-0.03Y~57.2
Al-0.2Y-0.05Sc60.8144 ± 2 12.6 ± 0.2
Al-0.2Y-0.2Sc60.2194 ± 1 12.2 ± 0.4
Al-0.2Y-0.2Sc-0.3Er59.7207 14
Al-0.2Ce-0.2Sc-0.1Y61.77 ± 0.11198 ± 2 8.5 ± 0.2
Al-0.2Y-0.2Sc-0.3Yb54.9228 ± 2 11.5 ± 0.8
Table 8. Typical obtained aluminum purity grade and its impurity content—in ppm, data from Reprinted from ref. [50].
Table 8. Typical obtained aluminum purity grade and its impurity content—in ppm, data from Reprinted from ref. [50].
AlSiFeCuMnMgCrNiZnTiVBGaZr
3N61351651013421010101067010
3N8696039423561344210
4N32218251105022113
4N77430110111121
5N50.70.20.30.10.50.10.0050.050.050.010.10.0050.01
Table 9. Comparison of different casting—rolling processes.
Table 9. Comparison of different casting—rolling processes.
Equipment InvestmentEnergy ConsumptionAdvantage
continuous casting and rolling processmediummediumcontinuous production, high efficiency
two-roll cast-rolling processlowlowshort process
DEPhighhighbetter mechanical properties
Continuous Casting and Extrusion Processmediummediumhigh flexibility
Note: DEP = Direct water-cooled semi-continuous casting and extrusion process.
Table 10. Chemical composition of base metal and welding wire (wt.%) Reprinted from ref. [57].
Table 10. Chemical composition of base metal and welding wire (wt.%) Reprinted from ref. [57].
ElementSiFeCuMnMgCrZnTiOther ElementsAl
7075/T6 Base Metal0.080.201.410.062.550.205.600.02≤0.05Bal.
7075Wire<0.010.032.180.161.96<0.015.80<0.01≤0.05Bal.
Table 11. Chemical composition table of 6061 aluminum alloy (wt.%) Reprinted from ref. [57].
Table 11. Chemical composition table of 6061 aluminum alloy (wt.%) Reprinted from ref. [57].
MgSiMnFeCuCrZnTiAl
0.970.130.120.560.270.120.080.023Bal.
Table 12. Mechanical and electrical performance parameters of 6061 aluminum alloy tensile test specimens Reprinted from ref. [57].
Table 12. Mechanical and electrical performance parameters of 6061 aluminum alloy tensile test specimens Reprinted from ref. [57].
Strain Rate/(S−1)DirectionYield Strength/MPaTensile Strength/MPaElongation Rate/%
0.05RD3473959.0
ST3063557.9
LT3213789.8
0.01RD34639510.9
ST3123577.4
LT33437810.5
0.001RD34539110.4
ST3453557.4
LT3103678.2
Table 13. Tensile properties of 6201 Al alloy in different processing condition. Error values are standard deviations Reprinted from ref. [58].
Table 13. Tensile properties of 6201 Al alloy in different processing condition. Error values are standard deviations Reprinted from ref. [58].
SampleYield Strength/MPaTensile Strength/MPaUniform Elongation/%Elongation at Failure/%Electrical Conductivity/%IACS
ST121 ± 11210 ± 921 ± 332 ± 445.2 ± 0.2
UA153 ± 10235 ± 1416 ± 229 ± 248.6 ± 0.6
PA226 ± 17274 ± 1411 ± 119 ± 250.5 ± 0.2
ST + RS328 ± 13347 ± 162.0 ± 0.36.8 ± 0.543.9 ± 0.3
UA + RS367 ± 10371 ± 91.7 ± 0.16.5 ± 0.345.5 ± 0.4
PA + RS384 ± 8387 ± 140.7 ± 0.16.1 ± 0.250.3 ± 0.7
ST + RS + RA160341 ± 21360 ± 165.1 ± 0.212.8 ± 0.347.0 ± 0.2
UA + RS + RA160350 ± 15363 ± 144.2 ± 0.311.8 ± 0.450.6 ± 0.3
PA + RS + RA160348 ± 17352 ± 134.0 ± 0.211.5 ± 0.351.7 ± 0.2
ST + RS + RA180319 ± 14330 ± 104.2 ± 0.111.3 ± 0.349.8 ± 0.2
UA + RS + RA180313 ± 10320 ± 93.5 ± 0.210.6 ± 0.551.5 ± 0.4
PA + RS + RA180308 ± 7312 ± 102.3 ± 0.110.2 ± 0.453.0 ± 0.5
ST + RS + RA200258 ± 11278 ± 82.5 ± 0.111.0 ± 0.252.8 ± 0.3
UA + RS + RA200269 ± 13277 ± 112.8 ± 0.110.8 ± 0.652.9 ± 0.2
PA + RS + RA200268 ± 9275 ± 83.0 ± 0.111.4 ± 0.453.5 ± 0.3
Note: ST = solution treatment; UA = under-aged; PA = peak-aged; RS = rotary swaging; RA = re-aging.
Table 14. Mechanical properties of rods with diameter 9 mm from alloy 6082 Reprinted with permission from ref. [66]. 2018 Materials Science Forum.
Table 14. Mechanical properties of rods with diameter 9 mm from alloy 6082 Reprinted with permission from ref. [66]. 2018 Materials Science Forum.
ConditionBar Length Selection Zone/mmUltimate Tensile Strength/MPaYield Strength/MPaElongation to Failure/%
hot-extrudedoutput28516522.2
middle21313323.8
in the end21012625.3
heat-treatedoutput39729415.6
36925814.0
middle38632615.6
37330515.8
In the end37130710.7
37831512.4
Requirements for EN 755-2 for extruded rods to 25 mm from alloy 6082 in condition T6≥295≥250≥8
Table 15. Mechanical properties of a drawn rod with a diameter of 8 mm from alloy 6082 Reprinted with permission from ref. [66]. 2018 Materials Science Forum.
Table 15. Mechanical properties of a drawn rod with a diameter of 8 mm from alloy 6082 Reprinted with permission from ref. [66]. 2018 Materials Science Forum.
ConditionUltimate Tensile Strength/MPaYield Strength/MPaElongation to Failure/%
hot-extruded29026812.0
28426012.3
26925112.0
heat-treated33126816.0
32726418.5
33527616.5
34328719.5
34328017.5
33127618.7
Requirements for EN 755-2 for extruded rods to 25 mm from alloy 6082 in condition T6≥310≥255≥10
Table 16. Comparative technical and economic indicators of traditional and new technology of twin roll casting–extruding Reprinted with permission from ref. [66]. 2018 Materials Science Forum.
Table 16. Comparative technical and economic indicators of traditional and new technology of twin roll casting–extruding Reprinted with permission from ref. [66]. 2018 Materials Science Forum.
IndicatorsTraditional Extruding TechnologyTRCR Technology
The yield/%melting7796
extruding6397
Electricity/(kW·h on 1 ton)melting83295
extruding53033073
Cost price at variable costs/(RUR/ton)193,754163,692
Reducing the cost price/(RUR/ton) 30,062
Note: TRCR = Two-Roll Cast-Rolling.
Table 17. Performance evolution of 6xxx series aluminum alloys under varying strain conditions Reprinted from ref. [76].
Table 17. Performance evolution of 6xxx series aluminum alloys under varying strain conditions Reprinted from ref. [76].
IndicatorsInitial Peak Aging StateRate of Area Reduction (19%)Rate of Area Reduction (91%)
average grain size/μm184.4781.228.80
yield strength/MPa289.61351.00410.72
Elongation/% 9.446.71
Conductivity/%IACS52.85 52.78
dislocation density/m−2 4.4 × 1014
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

Liu, X.; Jin, H.; Jiang, J. Development Status of Production Purification and Casting and Rolling Technology of Electrical Aluminum Rod. Metals 2025, 15, 981. https://doi.org/10.3390/met15090981

AMA Style

Liu X, Jin H, Jiang J. Development Status of Production Purification and Casting and Rolling Technology of Electrical Aluminum Rod. Metals. 2025; 15(9):981. https://doi.org/10.3390/met15090981

Chicago/Turabian Style

Liu, Xiaoyu, Huixin Jin, and Jiajun Jiang. 2025. "Development Status of Production Purification and Casting and Rolling Technology of Electrical Aluminum Rod" Metals 15, no. 9: 981. https://doi.org/10.3390/met15090981

APA Style

Liu, X., Jin, H., & Jiang, J. (2025). Development Status of Production Purification and Casting and Rolling Technology of Electrical Aluminum Rod. Metals, 15(9), 981. https://doi.org/10.3390/met15090981

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