Next Article in Journal
Impact of Metal Salt Oxidants and Preparation Technology on Efficacy of Bacterial Cellulose/Polypyrrole Flexible Conductive Fiber Membranes
Next Article in Special Issue
Analysis of Slag-Containing Steamed Concrete’s Composition Efficiency
Previous Article in Journal
Corrosion Behavior of Homogenized and Extruded 1100 Aluminum Alloy in Acidic Salt Spray
Previous Article in Special Issue
Development of Solid Waste-Based Composite Calcium Ferrite Flux and Its Application in Hot Metal Pre-Dephosphorization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Corrosion Behavior of 10 ppi TAD3D/5A05Al Composite in a Chloride Environment

1
Key Laboratory for Ecological Metallurgy of Multimetallic Mineral, Ministry of Education, Northeastern University, Shenyang 110819, China
2
School of Metallurgy, Northeastern University, Shenyang 110819, China
3
China Minmetals Nonferrous Metals Co., Ltd., Beijing 100044, China
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(6), 1280; https://doi.org/10.3390/ma17061280
Submission received: 23 January 2024 / Revised: 27 February 2024 / Accepted: 6 March 2024 / Published: 10 March 2024
(This article belongs to the Special Issue Recycling and Sustainability of Industrial Solid Waste)

Abstract

:
This study utilizes desalted and denitrated treated aluminum dross (TAD) as a raw material, along with kaolin and 10 ppi (pores per inch) polyurethane foam as a template. The slurry is converted into an aluminum dross green body with a three-dimensional network structure using the impregnation method. A three-dimensional network aluminum dross ceramic framework (TAD3D) is created at a sintering temperature of 1350 °C. The liquid 5A05 aluminum alloy at a temperature of 950 °C infiltrates into the voids of TAD3D through pressureless infiltration, resulting in TAD3D/5A05Al composite material with an interpenetrating phase composite (IPC) structure. The corrosion behavior of TAD3D/5A05 composite material in sodium chloride solution was examined using the salt spray test (NSS) method. The study shows that the pores of the TAD3D framework, produced by sintering aluminum dross as raw material, are approximately 10 ppi. The bonding between TAD3D and 5A05Al interfaces is dense, with strong interfacial adhesion. The NSS corrosion time ranged from 24 h to 360 h, during which the composite material underwent pitting corrosion, crevice corrosion and self-healing processes. Results from Potentiodynamic Polarization (PDP) and Electrochemical Impedance Spectroscopy (EIS) indicate that, as corrosion progresses, the Ecorr of TAD3D/5A05Al decreases from −0.718 V to −0.786 V, and Icorr decreases from 0.398 μA·cm−2 to 0.141 μA·cm−2. A dense oxide film forms on the surface of the composite material, increasing the anodic Tafel slope and decreasing the cathodic Tafel slope, thus slowing down the rates of cathodic and anodic reactions. Factors such as lower interface corrosion resistance or a relatively weak passivation film at the interface do not significantly diminish the corrosion resistance of TAD3D and 5A05Al. The corrosion resistance of the composite material initially decreases and then increases.

1. Introduction

The 5A05Al aluminum alloy belongs to the Al-Mg series, exhibiting high strength, excellent corrosion resistance, formability, and weldability. It finds extensive applications in transportation and mechanical engineering fields [1]. To enhance the wear resistance of 5A05 aluminum alloy, researchers have developed ceramic particles, ceramic whiskers, and ceramic frameworks to reinforce 5A05Al composite materials [2]. Such composite materials boast advantages including high temperature strength, elevated wear resistance, and high hardness [3]. Aluminum oxide (Al2O3) possesses excellent properties such as a high modulus, wear resistance, strength, and low density. Introducing aluminum oxide particles or whiskers into 5A05 aluminum alloy can significantly enhance its high-temperature and wear-resistant performance [4]. However, uneven dispersion of Al2O3 particles and whiskers in the composite material leads to a significant reduction in the material’s mechanical properties, wear resistance, and corrosion resistance [5]. Al2O3 is shaped into a three-dimensional network ceramic framework (Al2O33D). Subsequently, an aluminum alloy solution is infiltrated into the pores of the ceramic framework to prepare a three-dimensional network Al2O3 ceramic framework-reinforced aluminum-based composite material, Al2O33D/Al. In this material, Al2O33D and Al interpenetrate in three-dimensional space, exhibiting isotropy and a unique structure [6]. Such materials are referred to as interpenetration phase composites (IPC) [7]. When external forces are applied to the Al2O33D/Al composite material with an IPC structure, the interaction forces between Al2O33D ceramic and 5A05Al metal are effectively dispersed, absorbed, and dissipated. Consequently, the composite material can simultaneously leverage the high hardness, wear resistance, and thermal resistance of Al2O3 ceramic in three-dimensional space, as well as the high toughness and excellent thermal conductivity of 5A05 aluminum alloy. Research indicates that Al2O33D reinforcement significantly enhances the high-temperature resistance and wear resistance of the friction surface of aluminum-based composite materials [8,9]. Additionally, it maintains the low density, high thermal conductivity, and high specific strength of the composite material. Therefore, Al2O33D/5A05Al composite materials, such as for brake discs, exhibit tremendous potential [10]. Recently, researchers have utilized industrial solid waste aluminum dross as raw material for alumina-based ceramic material production [11]. Aluminum dross is an industrial residue generated during aluminum smelting, typically with an Al2O3 content exceeding 70%. If aluminum dross can be used to produce porous aluminum dross ceramic materials (TAD3D) instead of Al2O33D, it would bring significant benefits to the high-value utilization of solid waste and cost reduction in ceramic-reinforced aluminum-based composite materials [12].
Typically, methods for preparing IPC include pressureless infiltration, squeeze casting, and vacuum pressure infiltration, among others [13]. Squeeze casting and vacuum pressure infiltration processes are more complex, while the pressureless infiltration method for preparing IPC is relatively simple and enables low-cost industrial applications [14]. However, if the pressureless infiltration method is used to prepare TAD3D/5A05Al, and the wetting property between TAD3D and 5A05Al is poor, numerous defects may occur at the interface of TAD3D and 5A05Al, significantly reducing the overall performance of the composite material [15]. Moreover, when TAD3D/5A05Al IPC is used as a friction material, such as brake discs, it often operates in environments with water mist, salt spray, etc., for extended periods. Interface defects typically reduce the material’s corrosion resistance, leading to issues like pitting corrosion, crevice corrosion, and intergranular corrosion (IGC) [16]. TAD3D/5A05Al exhibits diverse phases, making its corrosion behavior intricate. Studies indicate that moderate levels of Mn, Mg, and Zn are advantageous for improving the corrosion resistance of aluminum alloys. Conversely, Fe, Cu, and Ni tend to decrease the corrosion resistance, while the impact of Si and Ti on aluminum alloy corrosion resistance varies depending on the alloy type and environment [17,18,19].
The presence of Mg significantly enhances the wettability of aluminum oxide ceramic with aluminum alloy solutions [20]. As 5A05 contains Mg, it markedly improves the wetting property between TAD3D and 5A05Al, enabling the possibility of preparing TAD3D/5A05Al using the pressureless infiltration method. However, there is limited literature on the pressureless infiltration method for preparing TAD3D/5A05Al IPC, making the study of the pressureless infiltration preparation of TAD3D/5A05Al and its corrosion-resistance behavior of great significance. This study employs aluminum dross as the aluminum source, kaolin as the silicon source, and utilizes a 10 ppi polyurethane foam as the template to prepare TAD3D using the sacrificial template method. The 5A05Al alloy solution is pressureless infiltrated into TAD3D to produce TAD3D/5A05Al. The microstructure of TAD3D/5A05Al and its neutral salt spray (NSS) corrosion performance at different time intervals are investigated using an electrochemical workstation, optical microscope, scanning electron microscope (SEM). The composite material prepared by pressureless infiltration exhibits minimal casting defects, and the interface between TAD3D and 5A05Al is tightly bonded, ensuring excellent corrosion resistance. This approach realizes the high-value utilization of aluminum dross solid waste, providing a crucial theoretical foundation for the application of aluminum dross ceramic-reinforced aluminum-based composite materials in the fields of long-lasting wear resistance and corrosion resistance.

2. Materials and Methods

2.1. Experimental Materials

This study utilizes treated aluminum dross as the main raw material. The TAD3D ceramic scaffold is prepared using the polyurethane foam template method, and the composite material is fabricated by pressureless casting of TAD3D with 5A05Al. The composite material samples undergo corrosion through the NSS method. This study explores the corrosion behavior of the composite material in a chloride environment by observing surface morphology, analyzing surface composition, and conducting electrochemical measurements on samples at different corrosion times.
The treated aluminum dross powder (Inner Mongolia Hengsheng Environmental Protection Technology Co., Ltd., Tongliao, China) and kaolin powder (Lingshou County Huayao Mineral Products Processing Factory, Shijiazhuang, China) serve as the primary raw materials. Their compositions are shown in Table 1. Dolapix CE-64 (Zschimmer-Schwarz (China) Ltd., Foshan, China) serves as a dispersant, Polyvinyl alcohol (PVA, Changchun Chemical (Jiangsu) Co., Ltd., Changshu, China) functions as a binder, and carboxymethyl cellulose (CMC, Shanghai Shenguang Edible Chemicals Co., Ltd., Shanghai, China) acts as a thickener. The treated aluminum dross, kaolin, Dolapix CE-64, CMC, and PVA were mixed in weight percentages of 80 wt.%, 17 wt.%, 2 wt.%, 0.5 wt.%, and 0.5 wt.%, respectively. The mixture was added to 100 mL distilled water, poured into a nylon jar equipped with 10 mm diameter ZrO2 balls, ball-milled for 12 h, resulting in a slurry with a solid content of approximately 70%. Adjust the pH of the slurry to 9.1–9.8 using ammonia water. Immerse a 10 ppi (10 pores per inch) polyurethane open-cell sponge template with a diameter Φ = 100 mm and thickness h = 30 mm into the slurry, and remove excess slurry using a double-roll extrusion. Compressed air is blown over the polyurethane sponge covering the slurry to eliminate remaining closed pores. After drying in a microwave for 15–20 min, a porous network body with good strength is obtained. The polyurethane sponge is visibly coated with a light gray 0.8–1 mm thick layer of aluminum dross. The body is placed in a high-temperature box resistance furnace (KSL1700X, Hefei Kejing Material Technology Co., Ltd., Hefei, China), heated from 25 °C to 1350 °C at a rate of 5 °C/min, and sintered at 1350 °C for 3 h to obtain TAD3D.
The TAD3D/5A05Al composite material was prepared using the pressureless casting method [21]. The 5A05Al alloy (Guangxi Nannan Aluminum Processing Co., Ltd., Nanning, China) with its chemical composition is shown in Table 2. The 5A05Al aluminum alloy block was placed in a corundum crucible, and the crucible along with the TAD3D block was positioned in a crucible resistance furnace (SG-2, Henan Eng Furnace Machinery Equipment Co., Ltd., Zhengzhou, China), with the temperature raised to 950 °C. The residue from the surface of the 5A05 Al in the crucible was removed, and the aluminum melt was stirred uniformly. TAD3D was placed into the aluminum melt, the crucible was tightly pressed with refractory materials and maintained at 950 °C for 30 min. The 5A05Al melt fully infiltrated into TAD3D, and after cooling, TAD3D/5A05Al was obtained. TAD3D/5A05Al specimens with dimensions of 10 × 10 × 6 mm3 were cut using a metallographic cutting machine (Laizhou Weiyi Machinery Manufacturing Co., Ltd., Yantai, China) for the study of the microstructure, electrochemical properties, and NSS corrosion performance of the composite material.

2.2. Neutral Salt Spray Corrosion

The NSS was conducted using an automated neutral salt spray chamber (ZK-60K, Dongguan Zhenke Testing Equipment Co., Ltd., Dongguan, China). The corrosive solution consisted of 5 wt.% with a pH range of 6.5 to 7.2 in a neutral NaCl solution. Experimental durations were set at 24 h, 72 h, 144 h, 240 h, and 360 h, with a test temperature of (35 ± 1) °C.

2.3. Surface Morphology of the Corroded Composite

After NSS corrosion, the corrosion microstructures on the sample surface were observed using an electron microscope equipped with an energy-dispersive X-ray spectrometer (EDS) (HITACHI S4800, Tokyo, Japan), and the composition of the corrosion products was analyzed.

2.4. Electrochemical Measurement

The electrochemical properties of TAD3D/5A05Al samples corroded under NSS for various durations were investigated using an electrochemical workstation (CHI 790E, Shanghai Chenhua Instrument Co., Ltd., Shanghai, China). Electrochemical testing comprised polarization curves (PDP) and electrochemical impedance spectroscopy (EIS), conducted using a three-electrode system. A standard calomel electrode (saturated potassium chloride solution) served as the reference electrode, with a platinum foil employed as the auxiliary electrode [19]. The TAD3D/5A05Al was connected to copper wire, sealed with epoxy resin, and used as the working electrode. The area of the test sample was 1 cm2. At room temperature, the sample was immersed in a 3.5% NaCl solution for 30 min. After stabilizing the open-circuit potential (OCP), polarization curves (PDP) and electrochemical impedance spectra (EIS) were measured. A scan rate of 0.25 mV/s was used for the cathodic-to-anodic scan, covering the range of OCP ± 600 mV for PDP measurements. During EIS measurements, a frequency range of 10−2–106 Hz was employed, with a sinusoidal perturbation signal of 10 mV added. The data were fitted using ZView software (Ver. 2.70) after the tests.

3. Results and Discussion

3.1. Microstructure of TAD3D/5A05Al

Figure 1 depicts the raw materials, process schematic, and sample photos for the preparation of TAD3D/5A05Al composite material. In Figure 1a, an SEM image of the treated aluminum dross is shown, revealing its spherical shape. The sintered TAD3D framework is shown in Figure 1b, with a shrinkage rate of approximately 18% after sintering the green body. The framework appears yellowish-white, indicating successful sintering into ceramics; it exhibits a measured compressive strength of around 2–5 MPa, and a pore size of approximately 10 ppi. Figure 1c illustrates the schematic of the pressureless infiltration process, where the TAD3D framework guides the molten 5A05Al alloy, effectively expelling fine air bubbles between the framework and the 5A05Al solution, preventing defects in the internal structure of the composite material and reducing internal stresses during the preparation process. The samples of TAD3D/5A05Al composite material exhibit a reflective silver-white luster and produce a solid tapping sound when cut to obtain test specimens, as shown in Figure 1d. On the smooth surface, gray streaks represent the framework, while the silver-white metallic luster corresponds to the 5A05Al alloy. The TAD3D framework has good affinity and wetting properties with the 5A05Al alloy, resulting in a dense interface and strong interfacial bonding. The samples of TAD3D/5A05Al composite material exhibit a reflective silver-white luster and produce a solid tapping sound when cut to obtain test specimens, as shown in Figure 1d. On the smooth surface, gray streaks represent the framework, while the silver-white metallic luster corresponds to the 5A05Al alloy. The TAD3D framework exhibits excellent wetting properties with the 5A05Al alloy solution, and a dense bond is formed at the interface between TAD3D and 5A05Al, resulting in high bonding strength.
Figure 2 illustrates the microstructure of the non-corroded TAD3D/5A05Al composite material. In Figure 2a-1, the microstructure of TAD3D in the composite material is shown. Figure 2a-2 is an enlarged view of Figure 2a-1, revealing that the pores in the ceramic phase are completely filled by the 5A05 aluminum alloy. TAD3D, manufactured by the polyurethane foam template method, contains triangular voids left by the combustion of polyurethane foam. In this study, the pressureless infiltration method was used to completely fill the triangular voids in TAD3D with the 5A05Al solution, achieving a dense IPC structure between the 5A05Al alloy and TAD3D, significantly enhancing the mechanical and corrosion resistance properties of TAD3D/5A05Al. In Figure 2a-2, the crystal boundaries and intra-crystalline regions of 5A05Al within the triangular voids are clearly visible, showing numerous second-phase particles, including the β-phase Mg5Al8, Mg2Si, and (FeMn)Al6. In 5A05 aluminum alloy, Si and Mg react to form Mg2Si according to Equation (1), while Al and Mg react to produce the β-phase Mg5Al8 according to Equation (2). The coarse needle-like or plate-like AlMnMgSi phase, generated by the elements Al, Mn, Mg, and Si, represents a non-equilibrium solidification microstructure [22].
Si + 2Mg → Mg2Si
8Al+ 5Mg → Mg5Al8
Figure 2b-1,b-2 depicts the microstructure of 5A05Al, revealing that the α-Al grains of 5A05Al contain relatively fewer second-phase particles such as Mg2Si, β-phase Mg5Al8, and (FeMn)Al6. The Mg2Si grains are small, approximately 3 μm, while Mg5Al8 is around 3 μm, Mg2Si is approximately 30 μm, and (FeMn)Al6 is about 15 μm.
Figure 2c-1,c-2 shows the microstructure of the interface between TAD3D and 5A05Al alloy. The interface between the 5A05Al alloy and TAD3D is clear, with a tight bond. At the interface, the quantity of second-phase particles, including Mg2Si, Mg5Al8, and (FeMn)Al6, increases, and the Mg2Si precipitates become coarser, forming a Chinese character pattern. Due to the different solubility of alloy elements in solid and liquid phases, elements such as Si, Mg, Fe, Mn in the 5A05 aluminum alloy tend to segregate into the liquid around the dendritic arms. Segregation alters the local thermodynamics of the alloy, providing a driving force for the formation of a second phase in the interdendritic region. Therefore, the second-phase particles, including Mg2Si, Mg5Al8, and (FeMn)Al6, increase in the interface between the alloy and the TAD3D.
By complementing the TAD3D ceramic network and 5A05A with their own advantages, the TAD3D/5A05Al composite can provide the desirable mechanical properties including high specific stiffness, high plastic flow strength, creep resistance, good oxidation, and corrosion resistance.
Figure 3 depicts SEM and EDS surface scan images of the non-corroded composite material. The dense interface between the 5A05Al alloy and TAD3D in the composite material is clearly visible. An interface layer, corroded by the etching solution, can be observed between TAD3D and the 5A05Al matrix, as indicated by the white dashed line in Figure 3a, with a thickness of approximately 4 μm. The EDS analysis of Figure 3b-1–b-5 indicates that the 5A05Al alloy is mainly composed of Al, Mn, Mg, and Si elements. Mg diffuses from the 5A05Al matrix to TAD3D and accumulates in TAD3D. The distribution of O elements in the 5A05Al matrix is uniform, while in TAD3D, it exhibits enrichment. Mg and Si elements form granular Mg2Si phases and excess Si particles. Si elements in the 5A05Al matrix precipitate in small amounts, exhibiting a punctate aggregation. Mg and Al elements form Mg5Al8 phases, and Fe, Mn, Al, and other elements form (FeMn)Al6 phases in TAD3D. TAD3D/5A05Al exhibits a slender network distribution of precipitated Mg2Si [23].

3.2. Microstructure of the Composite Material with NSS Corrosion Products

Figure 4 presents SEM images of the corroded surface of TAD3D/5A05 composite materials after NSS corrosion for 24 h, 72 h, 144 h, 240 h, and 360 h, respectively.
Figure 4a-1–a-3 reveal that after 24 h of NSS corrosion, there are relatively few pitting corrosion pits on the metal surface at the interface between TAD3D and 5A05Al. Pitting corrosion in the 5A05Al matrix generates, expands, and connects with each other, forming larger pitting corrosion pits. The corrosion products of 5A05 aluminum alloy increase, forming sheet-like accumulations on the corroded surface. Figure 4b-1–b-3 illustrates that after 72 h of NSS corrosion, corrosion products are generated on the surface of the 5A05Al matrix, accompanied by the formation of a passive film. Cracks are observed on the passive film, exhibiting a relatively uniform pattern with a length of approximately 50 μm and a width of about 5 μm. The cracks create an oxygen-deficient zone, where Cl accumulates in the solution, leading to a different metal surface state inside the cracks compared to the exterior. The solution inside the cracks transitions from neutral to acidic, resulting in crevice corrosion in these areas. Figure 4c-1–c-3 illustrates the condition of pitting corrosion on the 5A05Al matrix after 144 h of NSS corrosion. The diameters of these pitting pits have expanded laterally, increasing from 1 μm to 5 μm, with simultaneous growth in depth and area. The increased adsorption of Cl on the 5A05Al matrix results in the formation of more pitting pits on the surface. These pits expand, leading to the merging of multiple small pitting pits into larger ones. Some large pitting pits retain a small amount of residual corrosion products. Figure 4d-1–d-3 show that after 240 h of NSS corrosion, a secondary reaction occurs on the 5A05Al matrix, leading to the formation of an aluminum oxide film [24], which enhances the passivation performance of the metal, resulting in improved corrosion resistance. Destructive separation behavior due to crevice corrosion is prevented. Figure 4e-1–e-3 reveal that after 360 h of NSS corrosion, the larger pitting corrosion sites on the corroded sample surface are covered by corrosion products, leading to self-healing of the corroded surface and a significant enhancement in corrosion resistance [25]. We preliminarily believe that the reason for self-healing is that TAD3D sintered from aluminum dross has strong hydration reaction ability. TAD3D reacts with water to form aluminum hydroxide gel with high strength. At the same time, MgAl2O4, Al(OH)3, and Al2O3 are filled in the gaps between TAD3D and 5A05, as well as in the pitting pits on the surface, and together with the TAD3D skeleton embedded in 5A05Al, form a dense and sturdy passivation film. This results in a self-healing phenomenon at the interface of composite materials, enhancing their corrosion resistance.
Figure 5 presents the EDS analysis results of TAD3D/5A05 after NSS corrosion for 24 h, 72 h, 144 h, 240 h, and 360 h. EDS analysis results after 24 h of NSS corrosion presented in Figure 5a indicate that the main components of the corrosion products are O, Si, Mg, Al, Na, and Cl, suggesting that the corrosion product is Al(OH)3 [23]. The uniformly distributed O element on the 5A05Al matrix reacts with H and Al according to Equation (3) to produce Al(OH)3.
Al + 3O + 3H → Al(OH)3
Figure 5b presents the EDS analysis results after 72 h of NSS corrosion, indicating that the corrosion products are Al(OH)3 or corrosion products containing Mg. The morphology of the corrosion products is crack-like, and this feature is formed by the continuous generation of corrosion products in the active corrosion zone, gradually spreading and diffusing towards the surrounding area [26]. EDS analysis of the corrosion products indicates that the main elemental components in the corrosion products are Al, O, Cl, Mg, and Si. Due to the presence of interface reaction products and defects at the interface of TAD3D and the 5A05Al matrix, it becomes an active site. The Al matrix, as the anodic region, undergoes dissolution reactions first, forming Al3+ and releasing electrons, while the cathodic region undergoes oxygen absorption reactions forming OH. The electrochemical reactions are shown in Equations (4) and (5).
Al → Al3+ + 3e
O2 + 2H2O + 4e → 4OH
Some of the Al3+ in the solution reacts with OH to generate Al(OH)3, and it can further react to form Al2O3 through Equation (6).
2Al(OH)3 → Al2O3 + 3H2O
The formation of Al(OH)3 and Al2O3 through reactions has elevated the corrosion potential of the composite material, reducing the corrosion rate.
Figure 5c shows the EDS analysis results after 144 h, indicating the adsorption of Na+ and Cl ions at the interface, and some pitting corrosion extending to the interface between TAD3D and 5A05Al. The pitting corrosion is significantly deepened compared to NSS corrosion at 72 h, and the pits on the metal matrix develop into larger cracks, causing damage to the interface. At this point, corrosion intensifies [27].
Figure 5d shows the NSS corrosion for 240 h. EDS analysis of the corrosion products inside the pitting reveals that the predominant components are O and a small amount of Al, with overlapping states of the corrosion product Al(OH)3. Compared to 144 h, the number of cracks on the metal matrix surface decreases, and the corrosion behavior at the interface is mainly characterized by the healing of small cracks on the passivation film [28]. With the extension of NSS exposure, the oxide film on the metal surface gradually thins. Cl reacts directly with 5A05Al through tiny pores in the oxide film, forming a gray-white corrosion product AlCl3. Therefore, it can be inferred that the corrosion products mainly consist of Al(OH)3, Al2O3, and AlCl3.
Figure 5e shows NSS corrosion after 360 h, where some pits on the surface of the 5A05Al matrix are covered by corrosion products, improving the corrosion resistance of the material. EDS analysis of the corrosion products at the interface indicates the presence of O, Al, and Mg, suggesting the existence of MgAl2O4 spinel in the corrosion products. The corrosion products MgAl2O4, Al(OH)3, Al2O3, and AlCl3 adsorb at the interface, generating an oxide film along with the TAD3D ceramic phase. This oxide film serves as an insulating barrier, restricting the inward diffusion of corrosive substances (such as ions or moisture) into the Al alloy matrix. It demonstrates a synergistic effect resulting from the combination of TAD3D and 5A05Al in IPC composite materials, forming a stronger and more resistant passive layer that enhances the corrosion resistance of the composite material.

3.3. Microstructure of the Composite Material after Removal of NSS Corrosion Products

SEM images in Figure 6 reveal the microstructure of the TAD3D/5A05Al interface after corrosion removal. To observe fine details, the interface is magnified. Figure 6a-1–a-3 depicts the microstructure after 24 h of NSS corrosion, showing pitting on the surface of the 5A05Al alloy, with pit sizes ranging from 2 μm to 8 μm, and minor corrosion. Figure 6b-1–b-3 displays the microstructure after 72 h of NSS corrosion. Compared to Figure 6a-1–a-3, the pits deepen, cracks with lengths of 10 μm to 20 μm appear inside the pits, cracks extend outward from the pits, and a small amount of aluminum oxide film peels off. Figure 6c-1–c-3 shows the microstructure after 144 h of NSS corrosion, revealing widespread pitting on the material surface, with fine cracks of lengths ranging from 20 μm to 80 μm on the pits. Some areas of the 5A05Al matrix are covered by corrosion products, while others are exposed again, indicating a decrease in corrosion resistance with extended NSS time. Figure 6d-1–d-3 reveals the microstructure after 240 h of NSS corrosion, showing a small number of pits on the metal surface covered by aluminum oxide. Compared to 144 h, some corrosion healing occurs in the passive film. Figure 6e-1–e-3 shows the microstructure after 360 h of NSS corrosion, where the 5A05Al matrix generates a large amount of MgAl2O4, Al(OH)3, and Al2O3. These aluminum oxides, along with the TAD3D framework, form a dense corrosion product film, exhibiting a self-healing phenomenon in the composite material and enhancing corrosion resistance.
Figure 7 shows the EDS surface scan results of the TAD3D/5A05Al interface after corrosion removal, allowing the determination of the chemical composition of the corroded surface of the sample.
Figure 7a displays the EDS results after 24 h of NSS corrosion, indicating that the corrosion product is primarily composed of O, Si, Mg, Na, and Al elements. This corresponds to the corrosion-generated Al(OH)3 in the pits at the crack locations, causing minor damage to the passive film. The precipitated corrosion product is Al(OH)3 or corrosion-generated phases containing Al.
Figure 7b presents the EDS results after 72 h of NSS corrosion, indicating a higher content of Mg elements on the exposed matrix, suggesting the breakdown of the passive film. Compared to 5A05Al, the anodic oxidation of the β-phase Mg5Al8 leads to a higher sensitivity to the nucleation of early pitting micropores. Therefore, pits are formed in 5A05Al, and residual Na and Cl elements aggregate within these micropores.
Figure 7c presents the EDS results after 144 h of NSS corrosion, indicating that the surface product Al(OH)3 decomposes into aluminum oxides. Precipitated phases like Mg2Si and Si particles preferentially precipitate at grain boundaries, leading to the depletion of solute atoms in the 5A05Al matrix near the grain boundaries, forming precipitate-free zones (PFZs) on both sides along the grain boundaries. Since atoms like Si and Mn usually have higher potentials, they act as cathodes in the corrosion microcells. The precipitated phases exhibit a continuous distribution along the grain boundaries, forming a continuous cathode, while the PFZs formed near the precipitated phases are approximated as Al. Aluminum, with generally lower potential, acts as an anode in the corrosion microcell. Consequently, the PFZs can be rapidly dissolved in this corrosion microcell, resulting in corrosion [29]. The areas where the oxide film is damaged act as anodic sites in the active state, while those undamaged maintain a passive state as cathodic sites, forming an active–passive corrosion cell. The oxidation-reduction reactions cause metal dissolution within the pores to maintain internal electrical neutrality. Cl migrates into the pores, lowering the pH. Under the influence of H+ and Cl, the metal is in an active state, leading to the formation of an active (inside the pore)–passive (outside the pore) corrosion cell. The increased migration of Cl and the decreased pH promote corrosion.
Figure 7d shows the EDS results after 240 h of NSS corrosion, indicating that the thin layer of interface reaction products formed at the interface of TAD3D and 5A05.
Figure 7e shows the EDS results after 360 h of NSS corrosion, indicating that the corrosion products at the interface are primarily composed of O and Al. The corrosion products of the TAD3D/5A05Al composite material are mainly aluminum metal corrosion products. The corrosion products exhibit a point-like distribution, adsorbing at the interface between the TAD skeleton and the Al matrix, forming a dense surface oxide layer. The Al2O3 in the passive layer is largely an insulator for electron conduction, hindering localized corrosion and preventing the formation and expansion of corrosion-related defects such as pitting and cracks. The interpenetrating phase structure of IPC composite materials enhances the stability of the microstructure and improves overall corrosion resistance by forming a protective oxide layer.
Therefore, the NSS corrosion from 24 h to 360 h experienced processes of pitting corrosion and crevice corrosion. Self-healing of the passivation film became apparent at 240 h of corrosion, and at 360 h, the self-healing phenomenon was pronounced. This aligns with the conclusions drawn from SEM and EDS observations in Figure 4 and Figure 5.

3.4. Electrochemical Test Results of the Composite Material at Different NSS Corrosion Durations

Figure 8 depicts the PDP curves of TAD3D/5A05Al at different times after NSS. It can be observed that the PDP trends of TAD3D/5A05Al are consistent at different times after NSS. The Ecorr and Icorr values can be obtained from the PDP curve for Table 3. The Ecorr value serves as a thermodynamic criterion for the corrosion resistance of the composite material; a higher Ecorr value indicates stronger corrosion resistance, while a lower Ecorr value suggests weaker corrosion resistance [30]. From Table 3, it can be seen that as the NSS corrosion time increases from 0 to 144 h, the Ecorr value decreases. After 144 h of NSS corrosion, Ecorr decreases to the minimum value of −0.971 V, indicating that the passive film is damaged, the corrosion voltage decreases, and the corrosion resistance reaches its lowest point. From 240 h to 360 h of NSS corrosion, the Ecorr value increases. After 360 h of NSS corrosion, Ecorr increases to −0.786 V, slightly lower than the Ecorr value of the uncorroded sample (−0.718 V). This indicates that after 240 h of NSS, a large amount of corrosion products is formed on the surface of the composite material, protecting it. The corrosion voltage increases, the corrosion rate decreases, and the corrosion resistance improves. When NSS reaches 360 h, a complete corrosion product film forms on the surface. The corrosion voltage continues to increase, playing a suppressive role in corrosion, and enhancing corrosion resistance.
Icorr can serve as a dynamic criterion for the corrosion resistance of composite materials. A higher Icorr value indicates weaker corrosion resistance, while a lower Icorr value indicates stronger corrosion resistance. The trend of Icorr changes in Table 3 is also used to assess the corrosion resistance of the composite material, and the pattern is similar to that of Ecorr. As the NSS corrosion time increases from 0 to 144 h, the Icorr value increases. After 144 h of NSS corrosion, Icorr increases to the maximum value of 0.692 µA·cm−2, indicating that the passive film is damaged, the corrosion current is at its maximum, the corrosion rate increases, and the corrosion resistance reaches its lowest point. From 240 to 360 h of NSS corrosion, the Icorr value decreases. After 240 h of NSS corrosion, Icorr decreases to 0.178 µA·cm−2, which is already lower than the Icorr value of the uncorroded sample (0.398 µA·cm−2). After 360 h of NSS corrosion, Icorr decreases to 0.141 µA·cm−2, lower than the Icorr value of the uncorroded sample. After 240 h of NSS, the composite material’s surface begins to generate a large amount of corrosion products, providing protection. The corrosion current decreases, corrosion rate decreases, and corrosion resistance enhances. When NSS reaches 360 h, a complete corrosion product film forms on the surface, further reducing the corrosion current, suppressing corrosion, decreasing the corrosion rate, and improving corrosion resistance. Additionally, after NSS treatment, a dense oxide film forms on the surface of the TAD3D/5A05Al composite, increasing the anodic Tafel slope and decreasing the cathodic Tafel slope. This indicates that the presence of the oxide film alters the corrosion reaction kinetics of the aluminum alloy, reducing the rates of cathodic and anodic reactions.
Figure 9 shows the fitted results of EIS tests on TAD3D/5A05Al with corroded products after different NSS durations. Figure 9a shows the Nyquist plot of TAD3D/5A05Al, indicating a trend of decreasing and then increasing high-frequency arc radius with increasing NSS time. Since the arc radius is proportional to corrosion resistance, the corrosion resistance decreases and then increases with increasing NSS time. This is because the presence of O2 in the solution promotes the formation of the passive film on the surface of the 5A05Al alloy, but the film formation rate is slow in the early stages. Once a complete corrosion product film is formed, corrosion resistance begins to increase. Figure 9b,c depict the Bode plots of TAD3D/5A05Al. The impedance values of samples at different NSS times in the Bode plots exhibit a trend of initially decreasing and then increasing, consistent with the results from PDP. After NSS for 24 h, TAD3D/5A05Al is under the protection of a passive film, exhibiting excellent corrosion resistance. Between NSS 72 h and 144 h, with the increasing NSS time, the arc radius decreases, the peak value of the phase angle in the low-frequency region decreases, the height of the phase angle peak in the high-frequency region decreases, and its width narrows. Additionally, the slope of log f versus |Z| modulus decreases, and the impedance modulus value decreases [31]. All these indicate that the electrode is in the anodic dissolution stage, and the corrosion activity on the electrode surface becomes more intense. This suggests that the passive film of TAD3D/5A05Al is damaged, leading to a reduction in corrosion resistance. The sample corroded for 144 h. It is possible that, although the corrosion current is still high and the corrosion potential is still low, the passivation film of the composite material has begun to form; moreover, EIS shows an increase in the obstruction of the composite. After NSS for 240 h, there are more corrosion products on the 5A05Al alloy. Corrosion products such as MgAl2O4, Al(OH)3, Al2O3, and AlCl3 adsorb at the interface and, together with TAD3D ceramics, form a protective oxide film, reducing the unevenness of the microelectrochemical environment and the pitting sensitivity on the surface of the composite material. The corrosion process of the composite material involves the inhibition of both anodic dissolution and cathodic depolarization reactions. Due to the abundant corrosion products and their rapid generation, approaching or exceeding the diffusion rate, the arc radius increases. The peak value of the phase angle in the low-frequency region increases, and its width widens. The slope of log f versus |Z| modulus increases, and the impedance modulus value increases. All these indicate that the corrosion resistance begins to enhance. By NSS 360 h, the low-frequency impedance modulus value increases, the arc radius begins to increase, and the peak phase angle region widens. This indicates a steady improvement in the corrosion resistance of TAD3D/5A05Al, as the surface corrosion products are most abundant and dense, and the surface passivation film is fully formed, achieving optimal corrosion resistance. Figure 9d illustrates the equivalent circuit diagram used during the fitting process. By analyzing the EIS spectra, it can be concluded that all the curves exhibited two capacitive loops. It can be assumed that overall response of the system is associated with (R1CPE1) loop owing to the formation of film on the alloy surface and (R2CPE2) loop due to the double layer formed at metal–solution interface. Among them, Rs is the resistance of the electrolyte, CPE1 is the oxide film capacitance, R1 is the oxide film resistance, CPE2 is the double-layer capacitance, and R2 is the charge transfer resistance. The 1st loop (R1CPE1) encompasses all the information related to the surface layer and the possible defects that may be present within it. The EIS results were analyzed with the ZView program using the equivalent circuit shown in Figure 9d, and the values of the parameters obtained are tabulated in Table 4.
From Table 4, it can be observed that the solution resistance (Rs) remained constant. The results showed that the increases in impedance values were strongly linked by the values of R1 and R2. Further, a slight deviation from typical capacitive behavior at high frequencies, possibly due to uneven oxide film formation on the metal matrix. With different times of NSS corrosion treatment, the R2 value of the TAD3D/5A05Al increases. Without NSS, the R2 value of the sample is the lowest, at 7.30 × 103 Ω·cm2. After NSS 360 h, the R2 value of the sample is the highest, at 2.40 × 104 Ω·cm2. R1 and R2 play a decisive role in the impedance value of the sample, and compared to the values of R1 and R2, the Rs value can be basically ignored.
It is worth noting that the presence of the second phase/interface phase typically disrupts the continuity of the passivation film, and the vicinity of the second phase/interface phase often becomes a weak point in the passivation film. The presence of corrosive ions, especially Cl ions, can lead to the rupture of the passivation film, thereby triggering galvanic corrosion/pitting initiation. Specifically, the presence of Mg in 5A05Al can reduce the surface tension of the Al liquid and the surface energy between the Al liquid and TAD3D, improving the wetting properties between them. This enables the preparation of composite materials with lower porosity and higher density, indirectly suppressing the direct interface reaction between TAD3D and Al liquid and the formation of harmful interface reaction phases. Meanwhile, electrochemical tests indicate that factors such as lower interface corrosion resistance or a relatively weak passivation film at the interface do not significantly reduce the corrosion resistance of TAD3D and 5A05Al. This is because the TAD3D/5A05Al composite material has a unique IPC structure. The Mg in 5A05Al improves the wetting properties between the Al liquid and TAD3D and simultaneously forms a macroscopic interface composite with TAD3D. When immersed in a corrosive medium containing ions, before the formation of a chemical oxidation film, oxygen and electrons freely diffuse and migrate at the interface between the solution and the metal. Therefore, the anodic and cathodic corrosion rates of TAD3D/5A05Al are relatively high. After the formation of the chemical oxidation film, this uniformly dense film hinders the diffusion and migration of O2 and electrons between the solution and the metal interface, suppressing both anodic and cathodic reactions. As a result, the corrosion resistance of the aluminum alloy is improved, ensuring that the overall corrosion resistance of the composite material is not reduced. This is consistent with the microscopic morphology of the oxide film observed by SEM and EDS chemical composition analysis.

4. Conclusions

In this investigation, the corrosion behavior of TAD3D/5A05Al composite material in a chloride environment was investigated through various methods, including SEM, EDS, and electrochemical tests. The conclusions are as follows:
  • Processed aluminum dross is used as the raw material, and the sintered TAD3D framework has a pore size of about 10 ppi. During the pressureless melt infiltration process, the TAD3D framework guides the molten 5A05Al alloy, effectively expelling fine air bubbles between the framework and the 5A05Al solution, preventing defects inside the material. The interface between TAD3D and 5A05Al is tightly bonded, exhibiting strong interface adhesion.
  • SEM and EDS indicate that during the NSS corrosion from 24 h to 360 h, the composite material undergoes pitting corrosion, crevice corrosion and self-healing processes. Self-healing of the passivation film became apparent at 240 h of corrosion, and at 360 h, the self-healing phenomenon was pronounced. This is attributed to the unique IPC structure of the TAD3D/5A05Al composite. The macroscopic interface between 5A05Al and TAD3D facilitates the formation of corrosion products such as MgAl2O4, Al(OH)3, Al2O3, and AlCl3 adsorbed at the interface. Together with TAD3D ceramics, they form an oxide film, reducing the unevenness of the microelectrochemical environment and sensitivity to pitting.
  • PDP tests indicate that the corrosion current density of TAD3D/5A05Al increases and then decreases after NSS 0 h, 24 h, 72 h, 144 h, 240 h, and 360 h, suggesting that the corrosion resistance initially decreases and then increases. Ecorr decreases from −0.718 V to −0.786 V, and Icorr decreases from 0.398 μA·cm−2 to 0.141 μA·cm−2. After NSS treatment, a dense oxide film forms on the surface of the TAD3D/5A05Al composite, enhancing the anodic Tafel slope while reducing the cathodic Tafel slope. This indicates that the presence of the oxide film alters the corrosion reaction kinetics of the aluminum alloy, reducing the rates of cathodic and anodic reactions.
  • EIS test results indicate that with an increase in NSS time, the high-frequency arc radius shows a trend of initially decreasing and then increasing. The impedance values in the Bode plot also exhibit a trend of initially decreasing and then increasing. This suggests that factors such as lower interface corrosion resistance or a relatively weak passivation film at the interface do not significantly reduce the corrosion resistance of TAD3D and 5A05Al. Instead, the corrosion resistance of the composite material decreases initially and then increases.
The study has demonstrated the potential use of TAD3D/5A05Al in a chloride environment. In future work, it is advisable to extend the duration of NSS corrosion testing to observe whether the material’s corrosion resistance can be further enhanced. Attempts to protect the surface of composite materials using methods such as micro-arc oxidation. Additionally, further investigation into the passivation self-healing mechanism and corrosion resistance mechanisms of the composite material is warranted.

Author Contributions

Conceptualization, G.F., L.J., S.W. and Z.L.; writing original draft preparation, Z.L., S.W. and Y.C.; writing—review and editing, Z.L., S.W., Y.C., G.F. and L.J.; supervision, G.F. and L.J. project administration, Z.L., G.F. and L.J.; funding acquisition, G.F.; All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (U20A20239).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article

Conflicts of Interest

Author Shengpu Wang was employed by the company China Minmetals Nonferrous Metals Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Li, J.; Li, W.; Zhao, C.; Xing, Y.; Lang, F.; Hou, X. Experimental Investigation of Crack Propagation and Strain Fields Evolution around a Crack Tip in 5A05 Aluminum Alloy. Metals 2018, 8, 685. [Google Scholar] [CrossRef]
  2. Huang, Q.; He, R.; Wang, C.; Tang, X. Microstructure, Corrosion and Mechanical Properties of TiC Particles/Al-5Mg Composite Fillers for Tungsten Arc Welding of 5083 Aluminum Alloy. Materials 2019, 12, 3029. [Google Scholar] [CrossRef] [PubMed]
  3. Aktar Zahid Sohag, M.; Gupta, P.; Kondal, N.; Kumar, D.; Singh, N.; Jamwal, A. Effect of ceramic reinforcement on the microstructural, mechanical and tribological behavior of Al-Cu alloy metal matrix composite. Mater. Today Proc. 2020, 21, 1407–1411. [Google Scholar] [CrossRef]
  4. Ravikumar, M.; Reddappa, H.N.; Suresh, R.; Babu, E.R.; Nagaraja, C.R. Study on Micro–Nano Sized Al2O3 Particles on Mechanical, Wear and Fracture Behavior of Al7075 Metal Matrix Composites. Frat. Ed Integrità Strutt. 2021, 15, 166–178. [Google Scholar] [CrossRef]
  5. Potoczek, M.; Śliwa, R. Microstructure and Physical Properties of AlMg/Al2O3 Interpenetrating Composites Fabricated by Metal Infiltration into Ceramic Foams. Arch. Metall. Mater. 2011, 56, 1265–1269. [Google Scholar] [CrossRef]
  6. Ji, C.; Huang, H.; Wang, T.; Huang, Q. Recent advances and future trends in processing methods and characterization technologies of aluminum foam composite structures: A review. J. Manuf. Process. 2023, 93, 116–152. [Google Scholar] [CrossRef]
  7. Kota, N.; Charan, M.S.; Laha, T.; Roy, S. Review on development of metal/ceramic interpenetrating phase composites and critical analysis of their properties. Ceram. Int. 2022, 48, 1451–1483. [Google Scholar] [CrossRef]
  8. Jiang, L.; Jiang, Y.; Yu, L.; Yang, H.; Li, Z.; Ding, Y. Thermo-Mechanical Coupling Analyses for Al Alloy Brake Discs with Al2O3-SiC(3D)/Al Alloy Composite Wear-Resisting Surface Layer for High-Speed Trains. Materials 2019, 12, 3155. [Google Scholar] [CrossRef]
  9. Dolata, A.J. Tribological Properties of AlSi12-Al2O3 Interpenetrating Composite Layers in Comparison with Unreinforced Matrix Alloy. Materials 2017, 10, 1045. [Google Scholar] [CrossRef]
  10. Yu, L.; Hao, S.; Nong, X.; Cao, X.; Zhang, C.; Liu, Y.; Yan, Y.; Jiang, Y. Comparative Study on the Corrosion Resistance of 6061Al and SiC3D/6061Al Composite in a Chloride Environment. Materials 2021, 14, 7730. [Google Scholar] [CrossRef]
  11. Yang, H.L.; Li, Z.S.; Ding, Y.D.; Ge, Q.Q.; Jiang, L. Hydrolysis Behavior and Kinetics of AlN in Aluminum Dross during the Hydrometallurgical Process. Materials 2022, 15, 5499. [Google Scholar] [CrossRef] [PubMed]
  12. Cinarli, U.; Turan, A. Investigation of Alumina-Based Ceramic Production from Aluminum Black Dross. Min. Metall. Explor. 2021, 38, 257–267. [Google Scholar] [CrossRef]
  13. Kozera, P.; Boczkowska, A.; Perkowski, K.; Malek, M.; Kluczynski, J. Influence of Fabrication Method and Surface Modification of Alumina Ceramic on the Microstructure and Mechanical Properties of Ceramic-Elastomer Interpenetrating Phase Composites (IPCs). Materials 2022, 15, 7824. [Google Scholar] [CrossRef]
  14. Jiang, Y.; Xu, P.; Zhang, C.; Jin, F.; Li, Y.; Cao, X.; Yu, L. Simulation and Experimental of Infiltration and Solidification Process for Al2O3(3D)/5083Al Interpenetrating Phase Composite for High Speed Train Prepared by Low-Pressure Infiltration. Materials 2023, 16, 6634. [Google Scholar] [CrossRef]
  15. Tolulope Loto, R.; Babalola, P. Evaluation of the influence of alumina nano-particle size and weight composition on the corrosion resistance of monolithic AA1070 aluminium. Mater. Today Proc. 2022, 65, 2138–2143. [Google Scholar] [CrossRef]
  16. Ji, Y.-y.; Xu, Y.-z.; Zhang, B.-b.; Behnamian, Y.; Xia, D.-h.; Hu, W.-b. Review of micro-scale and atomic-scale corrosion mechanisms of second phases in aluminum alloys. Trans. Nonferrous Met. Soc. China 2021, 31, 3205–3227. [Google Scholar] [CrossRef]
  17. Lorenzi, S.; Carrozza, A.; Cabrini, M.; Nani, L.; Andreatta, F.; Virtanen, E.; Tirelli, T.; Pastore, T. Corrosion behavior assessment of an Al-Cu alloy manufactured via laser powder bed fusion. Corros. Sci. 2024, 227, 111698. [Google Scholar] [CrossRef]
  18. Jones, R.H.; Baer, D.R.; Danielson, M.J.; Vetrano, J.S. Role of Mg in the stress corrosion cracking of an Al-Mg alloy. Metall. Mater. Trans. a-Phys. Metall. Mater. Sci. 2001, 32, 1699–1711. [Google Scholar] [CrossRef]
  19. Ostash, O.P.; Polyvoda, S.Y.; Narivskyi, A.V.; Chepil, R.V.; Podhurska, V.Y.; Kulyk, V.V. Influence of Chemical Composition on the Structure, Mechanical, and Corrosion Properties of Cast Alloys of the Al-Mg-Sc System. Mater. Sci. 2021, 56, 570–576. [Google Scholar] [CrossRef]
  20. Pourbakhshi, M.; Sedghi, A.; Farhadinia, F. Effect of Zn and Mg additives on the fabrication of SiCp/Al2O3 + Al composite by Directed Metal Oxidation (DIMOX) of aluminum. Ceram. Int. 2020, 46, 22307–22312. [Google Scholar] [CrossRef]
  21. Yu, L.; Zhang, C.; Liu, Y.; Yan, Y.; Xu, P.; Jiang, Y.; Cao, X. Comparing the Corrosion Resistance of 5083 Al and Al(2)O(3)3D/5083 Al Composite in a Chloride Environment. Materials 2022, 16, 86. [Google Scholar] [CrossRef]
  22. Vysotskii, I.; Malopheyev, S.; Mironov, S.; Kaibyshev, R. Deformation behavior of friction-stir welded Al-Mg-Mn alloy with ultrafine-grained structure. Mater. Charact. 2022, 185, 111758. [Google Scholar] [CrossRef]
  23. Son, Y.G.; Jung, S.S.; Park, Y.H.; Lee, Y.C. Effect of Semi-Solid Processing on the Microstructure and Mechanical Properties of Aluminum Alloy Chips with Eutectic Mg2Si Intermetallics. Metals 2021, 11, 1414. [Google Scholar] [CrossRef]
  24. Karabulut, H.; Karacif, K.; Çitak, R.; Çinici, H. Corrosion behavior of particle reinforced aluminum composites. Mater. Test. 2021, 63, 1157–1163. [Google Scholar] [CrossRef]
  25. Loto, R.T.; Babalola, P. Electrochemical analysis of SiC composite additions at 7.5% weight content on the corrosion resistance of monolithic aluminium alloy in sulphate–chloride solution. J. Mater. Res. Technol. 2019, 8, 2517–2527. [Google Scholar] [CrossRef]
  26. Zhao, K.; Han, G.; Gao, T.; Yang, H.; Qian, Z.; Hu, K.; Liu, G.; Nie, J.; Liu, X. Interface precipitation and corrosion mechanisms in a model Al–Zn–Mg–Cu alloy strengthened by TiC particles. Corros. Sci. 2022, 206, 110533. [Google Scholar] [CrossRef]
  27. Ozturk, K.; Gecu, R.; Karaaslan, A. Microstructure, wear and corrosion characteristics of multiple-reinforced (SiC–B4C–Al2O3) Al matrix composites produced by liquid metal infiltration. Ceram. Int. 2021, 47, 18274–18285. [Google Scholar] [CrossRef]
  28. Shinozaki, I.; Sakakibara, Y.; Nakayama, G.; Tada, E.; Ooi, A.; Nishikata, A. Effects of Fe3+ on the Corrosion Behavior of High-Purity Aluminum in Neutral Solutions Containing Cl. Mater. Trans. 2021, 62, 492–497. [Google Scholar] [CrossRef]
  29. Renk, O.; Weissensteiner, I.; Cihova, M.; Steyskal, E.M.; Sommer, N.G.; Tkadletz, M.; Pogatscher, S.; Schmutz, P.; Eckert, J.; Uggowitzer, P.J.; et al. Mitigating the detrimental effects of galvanic corrosion by nanoscale composite architecture design. Npj Mater. Degrad. 2022, 6, 47. [Google Scholar] [CrossRef]
  30. Kumar, N.; Sharma, A.; Manoj, M.K. Influence of Different Aqueous Media on the Corrosion Behavior of B4C-Modified Lightweight Al-Mg-Si Matrix Composites. Materials 2022, 15, 8531. [Google Scholar] [CrossRef] [PubMed]
  31. Peng, Y.-y.; Yin, Z.-m.; Li, S.-h. Evaluation of pitting corrosion of Al-Mg-Mn-Sc-Zr alloy in EXCO solution by EIS. Trans. Nonferrous Met. Soc. China 2005, 15, 1226. [Google Scholar]
Figure 1. The raw materials, process, and samples for preparing TAD3D/5A05Al composite materials: (a) SEM of treated aluminum dross; (b) TAD3D; (c) schematic of the pressureless infiltration process; (d) TAD3D/5A05Al.
Figure 1. The raw materials, process, and samples for preparing TAD3D/5A05Al composite materials: (a) SEM of treated aluminum dross; (b) TAD3D; (c) schematic of the pressureless infiltration process; (d) TAD3D/5A05Al.
Materials 17 01280 g001
Figure 2. Optical micrograph (OM) image of TAD3D/5A05Al: (a-1,a-2) TAD3D; (b-1,b-2) 5A05Al; (c-1,c-2) composite interface.
Figure 2. Optical micrograph (OM) image of TAD3D/5A05Al: (a-1,a-2) TAD3D; (b-1,b-2) 5A05Al; (c-1,c-2) composite interface.
Materials 17 01280 g002
Figure 3. SEM and EDS of TAD3D/5A05Al: (a) SEM of composite interface; (b-1) Al; (b-2) Mg; (b-3) Mn; (b-4) O; (b-5) Si.
Figure 3. SEM and EDS of TAD3D/5A05Al: (a) SEM of composite interface; (b-1) Al; (b-2) Mg; (b-3) Mn; (b-4) O; (b-5) Si.
Materials 17 01280 g003
Figure 4. SEM images of TAD3D/5A05Al with corrosion products after NSS corrosion for different durations: (a-1a-3) 24 h; (b-1b-3) 72 h; (c-1c-3) 144 h; (d-1d-3) 240 h; (e-1e-3) 360 h.
Figure 4. SEM images of TAD3D/5A05Al with corrosion products after NSS corrosion for different durations: (a-1a-3) 24 h; (b-1b-3) 72 h; (c-1c-3) 144 h; (d-1d-3) 240 h; (e-1e-3) 360 h.
Materials 17 01280 g004
Figure 5. EDS of TAD3D/5A05Al with corrosion products after NSS corrosion for different durations: (a) 24 h; (b) 72 h; (c) 144 h; (d) 240 h; (e) 360 h.
Figure 5. EDS of TAD3D/5A05Al with corrosion products after NSS corrosion for different durations: (a) 24 h; (b) 72 h; (c) 144 h; (d) 240 h; (e) 360 h.
Materials 17 01280 g005
Figure 6. SEM images of TAD3D/5A05Al after removing corrosion products at different NSS corrosion durations: (a-1a-3) 24 h; (b-1b-3) 72 h; (c-1c-3) 144 h; (d-1d-3) 240 h; (e-1e-3) 360 h.
Figure 6. SEM images of TAD3D/5A05Al after removing corrosion products at different NSS corrosion durations: (a-1a-3) 24 h; (b-1b-3) 72 h; (c-1c-3) 144 h; (d-1d-3) 240 h; (e-1e-3) 360 h.
Materials 17 01280 g006
Figure 7. EDS of TAD3D/5A05Al after removing corrosion products at different NSS corrosion durations: (a) 24 h; (b) 72 h; (c) 144 h; (d) 240 h; (e) 360 h.
Figure 7. EDS of TAD3D/5A05Al after removing corrosion products at different NSS corrosion durations: (a) 24 h; (b) 72 h; (c) 144 h; (d) 240 h; (e) 360 h.
Materials 17 01280 g007
Figure 8. PDP of TAD3D/5A05Al with corrosion products after different NSS durations.
Figure 8. PDP of TAD3D/5A05Al with corrosion products after different NSS durations.
Materials 17 01280 g008
Figure 9. EIS of TAD3D/5A05Al with corrosion products after different NSS durations: (a) Nyquist diagram; (b) Bode diagram (|Z| − F); (c) Bode diagram (θF); (d) equivalent circuit.
Figure 9. EIS of TAD3D/5A05Al with corrosion products after different NSS durations: (a) Nyquist diagram; (b) Bode diagram (|Z| − F); (c) Bode diagram (θF); (d) equivalent circuit.
Materials 17 01280 g009
Table 1. Composition of treated aluminum dross (TAD) and kaolin (mass fraction).
Table 1. Composition of treated aluminum dross (TAD) and kaolin (mass fraction).
CompositionAl2O3SiO2MgOCaOTiO2
TAD87.543.065.151.080.15
kaolin42.2356.520.550.490.21
Table 2. Composition of 5A05 Al alloy (mass fraction).
Table 2. Composition of 5A05 Al alloy (mass fraction).
ElementsMgSiZnMnCuFeAl
wt.%5.020.100.040.430.050.21Balance
Table 3. Corrosion potential (Ecorr) and corrosion current density (Icorr) of TAD3D/5A05Al under different NSS corrosion durations.
Table 3. Corrosion potential (Ecorr) and corrosion current density (Icorr) of TAD3D/5A05Al under different NSS corrosion durations.
NSS Corrosion DurationsEcorr (V vs. SCE)Icorr (µA·cm−2)
0 h−0.7180.398
24 h−0.8850.537
72 h−0.9570.646
144 h−0.9710.692
240 h−0.8390.178
360 h−0.7860.141
Table 4. Electrochemical impedance parameters of TAD3D/5A05Al under different NSS corrosion durations.
Table 4. Electrochemical impedance parameters of TAD3D/5A05Al under different NSS corrosion durations.
NSS Corrosion DurationsRs
(Ω·cm2)
CPE1
(F·cm−2)
n1R1
(Ω·cm2)
CPE2
(F·cm−2)
n2R2
(Ω·cm2)
0 h12.967.96 × 10−40.637.84 × 1018.57 × 10−51.137.30 × 103
24 h23.071.25 × 10−40.582.36 × 1035.22 × 10−41.052.06 × 103
72 h4.563.08 × 10−50.751.28 × 1023.35 × 10−40.503.67 × 103
144 h9.621.40 × 10−91.342.17 × 1007.81 × 10−50.661.58 × 104
240 h8.417.21 × 10−50.684.69 × 1039.87 × 10−60.911.43 × 104
360 h8.731.51 × 10−50.551.46 × 1012.67 × 10−50.762.40 × 104
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

Li, Z.; Wang, S.; Chen, Y.; Fu, G.; Jiang, L. Corrosion Behavior of 10 ppi TAD3D/5A05Al Composite in a Chloride Environment. Materials 2024, 17, 1280. https://doi.org/10.3390/ma17061280

AMA Style

Li Z, Wang S, Chen Y, Fu G, Jiang L. Corrosion Behavior of 10 ppi TAD3D/5A05Al Composite in a Chloride Environment. Materials. 2024; 17(6):1280. https://doi.org/10.3390/ma17061280

Chicago/Turabian Style

Li, Zishen, Shengpu Wang, Yuxin Chen, Gaofeng Fu, and Lan Jiang. 2024. "Corrosion Behavior of 10 ppi TAD3D/5A05Al Composite in a Chloride Environment" Materials 17, no. 6: 1280. https://doi.org/10.3390/ma17061280

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