Next Article in Journal
Joint Quality Assessment of Ultrasonic Metal Welded Parts by Fracture Surface Evaluation
Previous Article in Journal
Trimodal Grain Structured Aluminum Matrix Composites Regulated by Transitional Hetero-Domains
Previous Article in Special Issue
Ultrafine Grain 316L Stainless Steel Manufactured by Ball Milling and Spark Plasma Sintering: Consequences on the Corrosion Resistance in Chloride Media
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study on Microstructure and High-Temperature Mechanical Properties of Al-Mg-Sc-Zr Alloy Processed by LPBF

1
Faculty of Mechanical Engineering and Mechanics, Ningbo University, Ningbo 315211, China
2
Ningbo Institute of Materials Technology & Engineering, Chinese Academy of Sciences, Ningbo 315201, China
3
China SciLong Lightweight Technology Co., Ltd., Ningbo 315336, China
*
Authors to whom correspondence should be addressed.
Metals 2024, 14(8), 890; https://doi.org/10.3390/met14080890
Submission received: 9 July 2024 / Revised: 27 July 2024 / Accepted: 2 August 2024 / Published: 4 August 2024

Abstract

:
Al-Mg-Sc-Zr alloy processed via laser powder bed fusion (LPBF) is poised for significant application in aerospace, where its high-temperature capabilities are paramount for the safety and longevity of engineered structures. This study offers a systematic examination of the alloy’s high-temperature tensile properties in relation to its microstructure and precipitate phases, utilizing experimental approaches. The LPBF-processed Al-Mg-Sc-Zr alloy features a bimodal microstructure, with columnar grains in the melt pool’s interior and equiaxed grains along its boundary, conferring exceptional properties. The application of well-calibrated processing parameters has yielded an alloy with an impressive relative density of 99.8%, nearly fully dense. Following a thermal treatment of 350 °C for 4 h, the specimens were subjected to tensile tests at both room and elevated temperatures. The data reveal that the specimens exhibit a tensile strength of 560.6 MPa and an elongation of 11.1% at room temperature. A predictable decline in tensile strength with rising temperature is observed: at 100 °C, 150 °C, 200 °C, and 250 °C; the respective strengths and elongations are 435.1 MPa and 25.8%, 269.4 MPa and 20.1%, 102.8 MPa and 47.9%, 54.0 MPa and 72.2%. These findings underpin the technical rationale for employing LPBF-processed Al-Mg-Sc-Zr alloy in aerospace applications.

1. Introduction

Laser powder bed fusion (LPBF) is an innovative additive manufacturing technique that integrates 3D modeling, material design, and automated control from various disciplines. This technology enables the rapid formation of parts from metal powders, eliminating the need for molds, machining, and assembly processes traditionally required in manufacturing. LPBF is a highly material-efficient process. By melting powder layer by layer, it minimizes material waste, which is particularly crucial for expensive and rare alloy materials. Moreover, LPBF technology significantly shortens production cycles and reduces tooling costs, making prototype design and small-batch production more economically feasible. In the aerospace sector, this facilitates rapid iteration and design optimization, accelerating the implementation of new technologies and materials. The characteristic of forming parts in a single step allows for weight reduction through integrated structural design, as demonstrated in recent studies [1,2]. LPBF-processed aluminum alloys, known for their high specific strength, excellent machinability, and superior corrosion resistance, are gaining significant traction in industries such as aerospace and automotive manufacturing due to their broad application prospects [3,4,5,6].
In recent years, an advanced Al-Mg-Sc-Zr alloy has emerged as an ideal candidate for LPBF. This alloy stands out as a multifunctional material, demonstrating exceptional repeatable printability and superior mechanical properties. Derived from an Al-Mg-based alloy modified with the addition of Sc and Zr elements, it has successfully been commercialized in aerospace and automotive industries, exemplified by Scalmalloy®—a material with the composition Al-4.5Mg-0.75Sc-0.35Zr-0.45Mn [7,8,9,10]. Current research indicates that LPBF-produced Al-Mg-Sc-Zr alloys can achieve a yield strength (YS) of up to 500 MPa at room temperature (RT), while maintaining an elongation at fracture (εf) of over 10%. The high strength of the Al-Mg-Sc-Zr alloy is attributed to its microstructure, which features a bimodal grain structure, known as a dual-phase structure. This consists of fine grains (FG) distributed along the melt pool boundaries and coarse grains (CG) within the melt pool interior. The fine grain regions are formed by the precipitation of Al3(Sc, Zr) phases during solidification at the melt pool boundaries. These precipitates have lattice parameters similar to those of α-Al and a low lattice mismatch, making them ideal nucleation sites for Al grains, thus promoting heterogeneous nucleation [11]. This approach significantly mitigates the issue of cracking, providing effective nucleation sites for the crystallization of α-Al grains, thereby reducing the formation of columnar crystals [12,13]. The heterogeneous nucleation of columnar grains within the melt pool is attributed to the rapid cooling from temperatures above 800 degrees Celsius, which generates a substantial temperature gradient [14,15,16]. Existing research indicates that optimizing LPBF process parameters—such as laser power, scan speed, layer thickness, hatch spacing, powder layer thickness, and scan path—can significantly enhance the alloy’s density and microstructure, thereby improving its mechanical properties [17,18,19,20,21,22,23]. The research demonstrates that Al-Mg-Sc-Zr alloys processed by the LPBF process possess exceptional mechanical performance and robust corrosion resistance under room-temperature conditions. Despite the extensive studies on LPBF-processed Al-Mg-Sc-Zr alloys, further research is needed to understand their tensile properties in high-temperature environments, particularly the microstructural evolution and changes in mechanical properties under varying temperature conditions.
Researching the high-temperature tensile properties of Al-Mg-Sc-Zr alloy processed by LPBF holds significant practical importance and potential industry impact. High-temperature environments are common in the aerospace sector, making it crucial to study the tensile properties of this alloy at elevated temperatures for designing and manufacturing highly reliable aerospace components. Furthermore, understanding the microstructural evolution and mechanical property changes of the alloy under different temperature conditions aids in optimizing LPBF process parameters, further enhancing material performance and manufacturing quality. This study systematically investigates the tensile property anisotropy of LPBF Al-Mg-Sc-Zr alloy in the as-built (AB) state and after annealing at 350 °C for 4 h (as-aged, AA). The 4 h annealing process is also considered as a heat treatment to alleviate residual stresses, which are a known issue for LPBF components. In light of this, the research characterizes the microstructure of the LPBF-formed Al-Mg-Sc-Zr alloy, performs tensile experiments at room and elevated temperatures. It examines the tensile properties and fracture morphology, establishing a relationship between microstructure, anisotropy, and high-temperature mechanical performance, which is significant for the application and advancement of LPBF-processed Al-Mg-Sc-Zr alloys.

2. Materials and Methods

2.1. Powder and LPBF Parameters

The Al-Mg-Sc-Zr powder utilized in this study was graciously supplied by Baohang Advanced Material Co., Ltd. (Nanchang, China) This powder was meticulously produced through gas atomization technology, a method specifically optimized for use with LPBF machinery in 3D printing applications. The scanning strategy employed adheres to the machine’s default checkerboard pattern. A detailed composition of the material is presented in Table 1, which reveals that the measured content of each constituent within the Al-Mg-Sc-Zr powder falls well within the prescribed standard range. For the processing of the Al-Mg-Sc-Zr powder, the HBD-450 3D printing equipment, provided by Hanbang Laser Technology Co., Ltd. (Zhongshan, China), was chosen. The specific printing parameters are delineated in Table 2. Additionally, the volume energy density (VED) is determined through the application of the formula presented below [24]:
V E D = P V × h × t
P is the laser power, V is the laser scanning speed, h is the hatching distance, and t is the layer thickness.
Table 1. Chemical composition of Al-Mg-Sc-Zr powder.
Table 1. Chemical composition of Al-Mg-Sc-Zr powder.
ElementsMgScZrMnSiAl
wt%7.800.590.310.470.66Bal
Table 2. Printing parameters of Al-Mg-Sc-Zr powder.
Table 2. Printing parameters of Al-Mg-Sc-Zr powder.
Laser PowerScanning SpeedLayer ThicknessHatching DistanceVED
400 W1250 mm/s50 μm120 μm53.33 J/mm3
The particle size distribution of the powder ranges from 12 to 72 μm, with the specific diameters at 10%, 50%, and 90% cumulative distribution (d10, d50, d90) being 19.53 μm, 35.18 μm, and 57.68 μm, respectively. The morphology of the powder is depicted in Figure 1, revealing a high sphericity and smooth surface overall. However, minor instances of particle agglomeration and a small number of irregularly shaped particles are observed. Such irregularities can impair the powder’s flowability, leading to uneven spreading and consequently reducing the density of the formed parts. Samples were annealed in a muffle furnace at 350 °C for 4 h, as illustrated in Figure 2. This heat treatment results in the supersaturation of Mn within the matrix and the formation of nanoscale Al3Sc precipitates, which contribute to solution strengthening, grain refinement, and precipitation strengthening.

2.2. Microstructure Characterization

The density of each sample was measured using the Archimedes drainage method. The formed samples were wire-cut and subsequently polished using progressively finer sandpapers to obtain standard metallographic specimens. These specimens were etched for 15 s with Keller’s reagent (1 mL HF, 1.5 mL HCl, 2.5 mL HNO3, and 95 mL H2O) for surface morphology and microstructure observation using optical microscopy (OM, Axio Observer, Carl Zeiss AG, Oberkochen, Germany) and scanning electron microscopy (SEM, GEMINI 300, Carl Zeiss AG, Oberkochen, Germany). For electron backscatter diffraction (EBSD) analysis, the samples were electrolytically polished at −50 °C using an electrolyte solution composed of 5% perchloric acid and 95% ethanol. The grain size and crystallographic orientation data were extracted using AZtecCrystal software (v2.0, Oxford Instruments NanoAnalysis, High Wycombe, UK). In this study, high resolution transmission electron microscopy (HRTEM, TF20, FEI Company, Hillsboro, OR, USA) was used to examine the second-phase particles in the alloy’s microstructure. Samples were prepared by wire-cutting 10 mm × 10 mm × 0.3 mm sections, which were then ground to a thickness of 150 μm using sandpaper. Subsequently, a 3 mm diameter disc was punched from the thinned section. The final stage of sample preparation involved electrolytic twin-jet polishing, using an electrolyte consisting of 30% nitric acid (HNO3) and 70% methanol (CH3OH), cooled to −30 °C. This process was carried out at an applied voltage of 20 V and a current of 120 mA. The samples were immediately rinsed with anhydrous ethanol after polishing. The prepared specimens were then subjected to HRTEM analysis.

2.3. Experimental Method of Mechanics

To test the tensile anisotropy of the Al-Mg-Sc-Zr alloy processed by LPBF, tensile specimens were categorized into H direction and Z direction based on the building direction, as shown in Figure 3a. The tensile tests were conducted according to the GB/T 228.2-2015 standard, with specimen dimensions illustrated in Figure 3b,c, applying an engineering strain rate of 7 × 10−5 s−1. During high-temperature tensile tests, the specimens were preconditioned at the specified temperature for 30 min before loading. To ensure data reproducibility, each tensile test was performed three times, and the average value was taken.

3. Results and Discussion

3.1. Microstructure Observations

Figure 4 displays the optical micrographs (OM) of samples from the Al-Mg-Sc-Zr alloy. The relative density of the samples, determined by the Archimedes method, reached 99.8%. The microstructure showed no visible cracks, confirming the successful fabrication of high-density samples using LPBF technology. Figure 5 provides OM images of the etched Al-Mg-Sc-Zr alloy samples. In the XOY and XOZ planes, a distinct dark stripe was observed in the brighter regions, indicating the boundary of the melt pool, while the brighter regions represent the interior of the melt pool. The shape of the melt pool was unevenly distributed in the XOY and XOZ planes. In the XOY plane, the melt pool appeared as irregular ellipses. In the YOZ plane, the melt pool exhibited a fan shape, characterized by a regular, fish-scale-like arrangement. Measurements of the observed surface indicated that the size of the melt pools ranged from 20 μm to 120 μm.
Figure 6a displays the X-ray diffraction (XRD, D8 Advance, Bruker Corporation, Karlsruhe, Germany) patterns of the pristine powder and the sample produced by the laser powder bed fusion (LPBF) technique. The XRD analysis confirms that the α-Al phase constitutes the principal crystalline phase in both the powder and the LPBF-processed samples. Notably, no Mg-containing phases or other phases resulting from processing were identified in the XRD patterns, possibly due to their low abundance below the detection capabilities of XRD or the rapid scanning speed of the XRD equipment, which could hinder the detection of these phases. A significant observation is the increased intensity of the diffraction peak on the (111) crystal plane in the LPBF-processed sample when compared to the original powder. This suggests a pronounced preferred orientation on the (111) plane, likely a consequence of the thermal processes and crystallographic growth during LPBF, which can significantly affect the resulting microstructure and material properties. Figure 6b offers a detailed perspective of the XRD pattern between 37.5° and 39°. It is evident that the (111) diffraction peak of the α-Al phase in the sample prepared by LPBF has undergone a leftward shift relative to the standard α-Al peak at 2θ = 38.446°. This shift results in the (111) diffraction peaks aligning with lower 2θ values [25]. As dictated by Bragg’s law, such a displacement to the left suggests an increase in the lattice parameters. The rapid solidification process characteristic of LPBF does not allow for timely diffusion of the majority of Mg atoms. These atoms are instead quickly assimilated into the α-Al matrix as solutes, causing substantial lattice distortion. This distortion is responsible for the observed leftward shift of the (111) peak.
To further investigate the microstructural properties of the samples processed by LPBF, such as grain size, shape, and texture, EBSD analysis was conducted. Figure 7 shows the EBSD inverse pole figures and pole figures of the XOY and XOZ planes of the alloy. As shown in Figure 7a, the ultrafine equiaxed grains of the Al-Mg-Sc-Zr alloy are distributed around the periphery of the melt pool on the XOZ plane, while elongated grains are mainly found within the melt pool. Figure 7b reveals that the columnar crystals in the melt pool exhibit a preferred orientation along <100>. Additionally, the melt pool depth is approximately 150 μm, significantly exceeding the alloy’s 50 μm layer thickness. This is attributed to the excessively high melt pool temperature during the LPBF process, which causes remelting of previously solidified material during subsequent layer scans [26,27].
The SEM images in Figure 8 offer insights into the microstructure of the Al-Mg-Sc-Zr alloy crafted by LPBF. Figure 8a displays large columnar crystals in the heart of the molten pool, with several black corrosion pits evident at the pool’s edges. A high-magnification SEM image, Figure 8b, captures nanoscale white precipitated phase particles within and encircling the corrosion pits at the pool’s edge. These particles are postulated to be nano-Al3(Sc, Zr), as per the analysis in Reference [28]. The temperature gradients within the molten pool, under suitable process parameters, promote the precipitation of these thermally stable Al3(Sc, Zr) particles [29]. As such, these particles precipitate stably and accumulate at the periphery of the molten pool during the repeated cycles of melting and solidification.
The HRTEM-BF-S images in Figure 9, complemented by EDS spectra, provide a detailed examination of the fine grain zone in the 3D printed samples, highlighting the concentration of Sc and Zr elements around this area. The precipitated secondary phase particles are identified as Al3(Sc, Zr) based on HRTEM and SAED analyses. The LPBF process involves a laser scanning a powder bed, with each layer being elevated and melted in sequence to achieve a layered scanning and melting effect. The formation of a molten pool upon laser interaction with Al powder results in the distinctive fish-scale pattern observable in the alloy’s microstructure (as shown in Figure 5). The cooling rate of these molten pools is exceptionally high, between 103 and 106 K/s [30]. As the molten pool cools, solidification initiates at the edges due to uneven heat dissipation. The hypereutectic elements Sc and Zr form a fine, dispersed L12 phase with Al, known as the Al3(Sc, Zr) precipitate. This phase is coherent with the Al matrix [31] and is prevalent at the fine grain zone’s grain boundaries, where it serves to pin the boundaries, restrict dislocation movement, and also acts as an effective nucleation site for grain refinement [32]. The solidification process at the molten pool boundary leads to the consumption of the Sc element, thereby eliminating potential nucleation points within the molten pool. Additionally, the nth layer’s molten pool is subject to the influence of neighboring molten pools (as shown in Figure 7, where there is a preference for columnar crystal orientation) and the subsequent n + 1th layer, which collectively restricts solidification within the molten pool. With temperature gradients diminishing from a high to a lower range, solidification within the molten pool initiates at its boundary. The depletion of Sc at the boundary, coupled with a substantial temperature gradient, results in solidification occurring counter to the direction of heat transfer, thus giving rise to columnar crystals. This explains the bimodal microstructure characteristic of the LPBF-processed Al-Mg-Sc-Zr alloy. As shown in Figure 10, the dispersion of nano-sized Al3(Sc, Zr) precipitates in the alloy’s fine-grained region enhances its strength. The Hall–Petch strengthening model [33] dictates that the strength increment from grain boundaries is related to the inverse square root of the grain size, formulated as σGB ∝ d^(−1/2). Based on reports in the existing body of literature [34], the alloy’s strengthening mechanisms are considered to involve dispersion strengthening and the refinement of grain size.

3.2. Mechanical Properties

Tensile testing was performed on specimens AB in their as-built state and AA in their as-aged state, under both ambient and elevated temperature conditions. The resulting stress–strain curves are displayed in Figure 11, with ‘H’ indicating the specimens oriented perpendicular to the build direction and ‘Z’ indicating those oriented parallel. The mechanical properties, including yield strength (YS), ultimate tensile strength (UTS), and elongation, are summarized in Table 3. The Al-Mg-Sc-Zr alloy fabricated using the LPBF technique demonstrates a higher tensile strength compared to the commercial aluminum alloy sheets (5B70), with a measured value of 421 MPa [35]. This enhanced mechanical performance can be attributed to the crack-free nature and the pronounced grain refinement observed in the microstructure of the LPBF-processed material. In Table 3 and Figure 11, we observe that the elongation of AA-H at 150 °C is lower than that at 100 °C. This may be due to the fact that at 150 °C, internal microvoids and cracks in the material are more likely to propagate, leading to a significant decrease in elongation. At 200 °C, despite the higher temperature, the material likely undergoes more plastic deformation, and the propagation mechanism of microvoids and cracks differs from that at 150 °C.
Fracture usually occurs in the weakest or most stressed part of the material, and the fracture morphology records the relevant information of the fracture process. By analyzing the deformation degree, position, color, precipitates, particle size, and pattern of the fracture, the nature of the fracture can be determined, the cause of the failure can be analyzed, and the mechanism of the fracture can be further studied [36]. In order to better understand the fracture mechanism, the fracture surface of Al-Mg-Sc-Zr samples prepared by LPBF after tensile was observed by SEM. From the macroscopic view, it can be found that the dimples of different sizes can be found in the fracture of the samples, so it can be determined that the tensile fracture of the alloy is a typical ductile fracture. With the increase in tensile temperature, the fracture of the sample is more uneven, which is related to the increase in elongation. As depicted in Figure 12a, the fracture surface morphology of the AA specimen exhibits characteristics akin to columnar grain structures at the crack origin site (a). Such features are exclusively observed on the fracture crack surfaces of specimens following heat treatment. This can be attributed to two primary factors: The material’s bimodal microstructure, characterized by a varied grain size distribution, results in an inhomogeneous stress distribution. The equiaxed grain areas, reinforced by the pinning effect of Al3Sc particles at the grain boundaries, demonstrate greater strength than the columnar grain regions. Consequently, crack initiation is more prone to occur in the columnar grain areas. The existence of larger grains within the bimodal microstructure contributes to a reduced crack propagation velocity, which in turn, augments the material’s fracture toughness. Figure 12c,d, representing the fracture surfaces at temperatures of 100 degrees and 200 degrees, respectively, corroborate the presence of this phenomenon.
Ductile fractures are generally characterized by three distinct regions: the fiber zone, which is also known as the goose down zone, the radiation zone, and the shear lip zone. These zones are the key components of fracture morphology [37]. As depicted in Figure 13a,c, the fiber zone of the tensile fracture is examined across a range of temperatures, showing an increase in irregularity with rising temperature. The crack propagation within this zone is slower in comparison to other areas. The fracture surface features a jagged pattern punctuated by holes, which at a microscopic level, is composed of many fibrous ‘mini peaks’, each with a 45° inclination to the tensile axis. This inclination indicates that the fiber zone is a result of the progressive enlargement and merging of microcracks during plastic deformation [38]. The dimples on the fracture surface are a consequence of dislocation slip, with the second-phase particles acting as the main impetus. Figure 13b,d show the dimple morphology on the fracture surfaces at 100 °C and 200 °C, respectively. The dense distribution of dimples indicates that the material exhibits high ductility. The dimples at 200 °C are larger than those at 100 °C, suggesting that the specimens underwent greater plastic deformation before fracture. This observation is consistent with the measured elongation results. The size and shape of these dimples are correlated with the material’s properties, including its strength, the dimensions, form, and dispersion of the second-phase particles, as well as the stress conditions, encompassing both the nature and magnitude of the stress applied [39]. Figure 9 displays the secondary phase particles located at the grain boundaries in the equiaxed zone, as detected by EDS. The elemental analysis confirms their enrichment with Scandium (Sc) and Zirconium (Zr). Drawing from Wang’s characterization [40], these particles are identified as the primary Al3(Sc, Zr) phase, which is directly responsible for the dimple formation observed in the tensile fracture surfaces of Al-Mg-Sc-Zr alloy systems. In alignment with dislocation theory [41], it is around these Al3(Sc, Zr) particles that dislocation loops are formed, contributing to the material’s ductile fracture behavior. As depicted in Figure 14, in the absence of external forces, dislocation loops are in equilibrium, held in check by the repulsion from Al3(Sc, Zr) particles and the stress of dislocation stacking. When an external force is sufficiently large, dislocations around the Al3(Sc, Zr) particles are set into motion once more, moving toward the particles. The formation of microvoids occurs when the elastic strain energy that has built up in front of the dislocation loop reaches a level that can overcome the interfacial bonding forces between the Al3(Sc, Zr) particles and the aluminum matrix. The formation of microvoids significantly alleviates the repulsive forces on dislocations, leading to a surge of dislocations toward the microvoids under external force, which in turn, drives their enlargement. At this juncture, the dislocation sources that were quiescent behind the dislocation loop become re-energized as the constraint from dislocation accumulation is lifted, initiating new dislocation loops that migrate incessantly toward the microvoids. This results in a swift and unstable enlargement and coalescence of the microvoids. Dislocations, capable of amassing on diverse slip planes, may induce the formation of microvoids through their movement along one or several planes. Concurrently, dislocations on other slip planes are drawn to the microvoids, further promoting their expansion [42].

4. Conclusions

The microstructure, room-temperature and high-temperature tensile properties and fracture morphology of Al-Mg-Sc-Zr samples prepared by laser powder bed fusion were studied. The main results are as follows:
  • The microstructure is characterized by neatly aligned melt pools, with the boundaries marked by fine equiaxed grains and the interiors by larger equiaxed grains, creating a distinct bimodal distribution. EBSD analysis indicates an even dislocation distribution across the alloy, devoid of any textural preference.
  • Room-temperature tensile tests demonstrated that the heat-treated specimens achieved ultimate tensile strengths up to 560.6 MPa, coupled with an elongation of 11.1%. Tensile strengths for both untreated and treated samples were modestly higher in the H direction than in the Z direction, with the discrepancy being less than 5%. A progressive decrease in strength with increasing test temperature was observed, with values at 100 °C being 435.1 MPa and 25.8%, at 150 °C being 269.4 MPa and 20.1%, at 200 °C being 102.8 MPa and 47.9%, and at 250 °C being 54.0 MPa and 72.2%, respectively. This trend signifies a trade-off between strength and ductility with rising temperature, with the material exhibiting greater plasticity at elevated temperatures.
  • Fracture surface analyses from tensile tests across the temperature range disclosed an abundance of dimples, indicative of ductile fracture behavior. The dimples are mainly attributed to the presence of the second-phase Al3(Sc, Zr) particles. Upon the application of substantial external forces, dislocations surrounding these particles become active again, leading to fracture when the forces surpass the interfacial bond strength between the particles and the aluminum matrix.

Author Contributions

Conceptualization: Z.R. and H.X. Methodology: H.Z. and Y.D. Software: S.C. and Y.Y. Validation: S.C. Formal Analysis: H.Z. and L.L. Investigation: X.S. and Y.Z. Resources: Y.Y. Data Curation: L.L. and Y.D. Writing—Original Draft Preparation: Z.R. and Y.Z. Writing—Review and Editing: Z.R. and Y.Z. Visualization: Z.Q. Supervision: H.X., and X.S. Project Administration: Z.Q. Funding Acquisition: H.X. and X.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by “Pioneer” and “Leading Goose” R&D Program of Zhejiang, China (grant number: 2024C01121); Yongjiang talent introduction programmer (2021A-112-G;2022A-202-G); “3315 Project” innovation team (2020A-27-C); international partnership program of the Chinese Academy of Sciences, grant no. 181GJHZ2023132MI.

Data Availability Statement

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

Conflicts of Interest

Author Yaoyao Ding was employed by the company China SciLong Lightweight Technology 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, X.; Li, G.; Chen, H.; Zhang, M.; Xi, L. Novel approach to additively manufacture high-strength Al alloys by laser powder bed fusion through addition of hybrid grain refiners. Addit. Manuf. 2021, 48, 102400. [Google Scholar] [CrossRef]
  2. Han, D.; Xiao, Y.; Chen, H.; Bian, Z.; Wu, Y.; Yang, H.; Wang, H.; Wang, H. Design, microstructure and thermal stability of a novel heat-resistant Al-Fe-Ni alloy manufactured by selective laser melting. J. Alloys Compd. 2021, 885, 160949. [Google Scholar]
  3. Li, Q.; Li, G.; Lin, X.; Zhu, D.; Jiang, J.; Shi, S.; Liu, F.; Huang, W.; Vanmeensel, K. Development of a high strength Zr/Sc/Hf-modified Al-Mn-Mg alloy using laser powder Bed fusion: Design of a heterogeneous microstructure incorporating synergistic multiple strengthening mechanisms. Addit. Manuf. 2022, 57, 102967. [Google Scholar] [CrossRef]
  4. Kotadia, H.R.; Gibbons, G.; Das, A.; Howes, P.D. A review of laser powder Bed fusion additive manufacturing of aluminum alloys: Microstructure and properties. Addit. Manuf. 2021, 46, 102155. [Google Scholar]
  5. Lim, C.H.; Li, H.; Krishnan, M.; Chen, K.; Li, J. Novel method of residual stress reduction for AlSi10Mg manufactured using selective laser melting without compromise of mechanical strength. Virtual Phys. Prototyp. 2023, 18, 4. [Google Scholar] [CrossRef]
  6. Tang, Z.; Wei, Q.; Gao, Z.; Yang, H.; Wang, A.; Wan, L.; Luo, C.; Wu, Y.; Wang, H.; Wang, H. 2000W blue laser directed energy deposition of AlSi7Mg: Process parameters, molten pool characteristics, and appearance defects. Virtual Phys. Prototyp. 2023, 18, 3. [Google Scholar] [CrossRef]
  7. Palm, F.; Leuschner, R.; Schubert, T.; Kieback, B. Scalmalloy®= A unique high strength ALMgSc type material concept processed by innovative technologies for aerospace applications. In Proceedings of the World Powder Metallurgy Congress and Exhibition, World PM 2010, Florence, Italy, 10–14 October2010. [Google Scholar]
  8. Kuo, C.N.; Peng, P.C.; Liu, D.H.; Chao, C.Y. Microstructure evolution and mechanical property response of 3D-Printed Scalmalloy with different heat-treatment times at 325 °C. Metals 2021, 11, 555. [Google Scholar] [CrossRef]
  9. Schmidtke, K.; Palm, F.; Hawkins, A.; Emmelmann, C. Process and mechanical properties: Applicability of a scandium modified Al-alloy for laser additive manufacturing. Phys. Procedia 2011, 12, 369–374. [Google Scholar] [CrossRef]
  10. Martucci, A.; Aversa, A.; Manfredi, D.; Bondioli, F.; Biamino, S.; Ugues, D.; Lombardi, M.; Fino, P. Low-power laser powder bed fusion processing of Scalmalloy®. Materials 2022, 15, 3123. [Google Scholar] [CrossRef]
  11. Li, D.; Wu, Y.; Geng, Z.; Zhang, J.; Chen, C.; Liu, X.; Liu, Y.; Zhou, K. High strength Al−Mg−Sc−Zr alloy with heterogeneous grain structure and intragranular precipitation produced by laser powder bed fusion. J. Alloys Compd. 2023, 939, 168722. [Google Scholar] [CrossRef]
  12. Jägle, E.A. Alloy Design for Additive Manufacturing. In DARE Annual Workshop 2016; University of Sheffield: Sheffield, UK, 2016. [Google Scholar]
  13. Carluccio, D.; Bermingham, M.J.; Zhang, Y.; StJohn, D.H.; Yang, K.; Rometsch, P.A.; Wu, X.; Dargusch, M.S. Grain refinement of laser remelted Al-7Si and 6061 aluminium alloys with Tibor® and scandium additions. J. Manuf. Process. 2018, 35, 715–720. [Google Scholar] [CrossRef]
  14. Spierings, A.B.; Dawson, K.; Heeling, T.; Uggowitzer, P.J.; Schäublin, R.; Palm, F.; Wegener, K. Microstructural features of Sc- and Zr-modified Al-Mg alloys processed by selective laser melting. Mater. Des. 2017, 115, 52–63. [Google Scholar] [CrossRef]
  15. Yang, K.V.; Shi, Y.; Palm, F.; Wu, X.; Rometsch, P. Columnar to equiaxed transition in Al-Mg(-Sc)-Zr alloys produced by selective laser melting. Scr. Mater. 2018, 145, 113–117. [Google Scholar] [CrossRef]
  16. Spierings, A.B.; Dawson, K.; Kern, K.; Palm, F.; Wegener, K. SLM-processed Sc- and Zr- modified Al-Mg alloy: Mechanical properties and microstructural effects of heat treatment. Mater. Sci. Eng. A 2017, 701, 264–273. [Google Scholar] [CrossRef]
  17. Ekubaru, Y.; Gokcekaya, O.; Ishimoto, T.; Sato, K.; Manabe, K.; Wang, P.; Nakano, T. Excellent strength–ductility balance of Sc-Zr-modified Al–Mg alloy by tuning bimodal microstructure via hatch spacing in laser powder bed fusion. Mater. Des. 2022, 221, 110976. [Google Scholar] [CrossRef]
  18. Cordova, L.; Bor, T.; de Smit, M.; Carmignato, S.; Campos, M.; Tinga, T. Effects of powder reuse on the microstructure and mechanical behaviour of Al–Mg–Sc–Zr alloy processed by laser powder bed fusion (LPBF). Addit. Manuf. 2020, 36, 101625. [Google Scholar] [CrossRef]
  19. Wang, Z.; Lin, X.; Wang, J.; Kang, N.; Hu, Y.; Wang, D.; Li, H.; Huang, W.; Pan, F. Remarkable strength-impact toughness conflict in high-strength Al-Mg-Sc-Zr alloy fabricated via laser powder bed fusion additive manufacturing. Addit. Manuf. 2022, 59, 103093. [Google Scholar] [CrossRef]
  20. Zhang, H.; Zhang, L.C.; Liu, H.; Niu, X.; Lam, M.C.; Zhang, W.; Jin, X.; Chu, F.; Wu, X.; Cao, S. Strong and ductile Al–Mn–Mg–Sc–Zr alloy achieved in fabrication-rate enhanced laser powder bed fusion. Virtual Phys. Prototyp. 2023, 18, e2250769. [Google Scholar] [CrossRef]
  21. Zhou, Z.; Chen, C.; Guan, J.; Yuan, K. Study on the Effect of Remelt Scanning and Heat Treatment Processes on the Structure and Properties of Laser Powder Bed Fusion Formed Al-Mg-Sc-Zr Alloys. J. Phys. 2023, 2566, 012085. [Google Scholar] [CrossRef]
  22. Zhang, Z.; Sun, J.; Wu, J.; Chen, Z.; Zhang, B.; Li, Z.; Li, X.; Zhang, B. Anisotropic response in corrosion behavior of laser powder bed fusion Al-Mn-Mg-Sc-Zr alloy. Corros. Sci. 2022, 208, 110634. [Google Scholar] [CrossRef]
  23. Schliephake, D.; Lopes, C.; Eggeler, Y.M.; Chen, H.; Freudenberger, J.; Bayoumy, D.; Huang, A.J.; Kauffmann, A. Improved work hardening capability and ductility of an additively manufactured and deformed Al-Mn-Mg-Sc-Zr alloy. J. Alloys Compd. 2022, 924, 166499. [Google Scholar] [CrossRef]
  24. Read, N.; Wang, W.; Essa, K.; Attallah, M.M. Selective laser melting of AlSi10Mg alloy: Process optimisation and mechanical properties development. Mater. Des. 2015, 65, 417–424. [Google Scholar] [CrossRef]
  25. Zou, Y.; Tan, C.; Qiu, Z.; Ma, W.; Kuang, M.; Zeng, D. Additively manufactured SiC-reinforced stainless steel with excellent strength and wear resistance. Addit. Manuf. 2021, 41, 101971. [Google Scholar] [CrossRef]
  26. Zou, T.; Mei, S.; Chen, M. Microstructure and Electrochemical Corrosion Properties of AlMgScZr Alloys Fabricated Using Selective Laser Melting. Zhongguo Jiguang 2023, 50, 85–94. [Google Scholar]
  27. Spierings, A.B.; Dawson, K.; Uggowitzer, P.J.; Wegener, K. Influence of SLM scan-speed on microstructure, precipitation of Al3Sc particles and mechanical properties in Sc-and Zr-modified Al-Mg alloys. Mater. Des. 2018, 140, 134–143. [Google Scholar] [CrossRef]
  28. Zhang, H.; Gu, D.; Yang, J.; Dai, D.; Zhao, T.; Hong, C.; Gasser, A.; Poprawe, R. Selective laser melting of rare earth element Sc modified aluminum alloy: Thermodynamics of precipitation behavior and its influence on mechanical properties. Addit. Manuf. 2018, 23, 1–12. [Google Scholar] [CrossRef]
  29. Deng, Y.; Peng, B.; Xu, G.; Pan, Q.; Yin, Z.; Ye, R.; Wang, Y.; Lu, L. Effects of Sc and Zr on mechanical property and microstructure of tungsten inert gas and friction stir welded aerospace high strength Al–Zn–Mg alloys. Mater. Sci. Eng. A 2015, 639, 500–513. [Google Scholar] [CrossRef]
  30. Sing, S.; Wiria, F.; Yeong, W. Selective laser melting of titanium alloy with 50 wt% tantalum: Effect of laser process parameters on part quality. Int. J. Refract. Met. Hard Mater. 2018, 77, 120–127. [Google Scholar] [CrossRef]
  31. Rometsch, P.; Zhong, H.; Nairn, K.M.; Jarvis, T.; Wu, X. Characterization of a laser-fabricated hypereutectic Al–Sc alloy bar. Scr. Mater. 2014, 87, 13–16. [Google Scholar] [CrossRef]
  32. Li, G.; Zhao, N.; Liu, T.; Li, J.; He, C.; Shi, C.; Liu, E.; Sha, J. Effect of Sc/Zr ratio on the microstructure and mechanical properties of new type of Al–Zn–Mg–Sc–Zr alloys. Mater. Sci. Eng. A 2014, 617, 219–227. [Google Scholar] [CrossRef]
  33. Hansen, N. Hall–Petch relation and boundary strengthening. Scr. Mater. 2004, 51, 801–806. [Google Scholar] [CrossRef]
  34. Li, R.; Wang, M.; Li, Z.; Cao, P.; Yuan, T.; Zhu, H. Developing a high-strength Al-Mg-Si-Sc-Zr alloy for selective laser melting: Crack-inhibiting and multiple strengthening mechanisms. Acta Mater. 2020, 193, 83–98. [Google Scholar] [CrossRef]
  35. Shen, J.; Chen, B.; Wan, J.; Shen, J.; Li, J. Effect of annealing on microstructure and mechanical properties of an Al–Mg-Sc-Zr alloy. Mater. Sci. Eng. A 2022, 838, 142821. [Google Scholar] [CrossRef]
  36. Bayoumy, D.; Schliephake, D.; Dietrich, S.; Wu, X.H.; Zhu, Y.M.; Huang, A.J. Intensive processing optimization for achieving strong and ductile Al-Mn-Mg-Sc-Zr alloy produced by selective laser melting. Mater. Des. 2021, 198, 109317. [Google Scholar] [CrossRef]
  37. Bathias, C. Application of fracture mechanics to aluminum alloys selection. Eng. Fract. Mech. 1978, 10, 267–282. [Google Scholar] [CrossRef]
  38. Kuchariková, L.; Tillová, E.; Chalupová, M.; Uhríčik, M.; Pastierovičová, L.; Belan, J. The Fractographic Analysis of Tensile and Fatigue Fracture Surfaces in Secondary A356 Aluminum Alloy with a Higher Concentration of Iron. Defect. Diffus. Forum 2023, 422, 15–20. [Google Scholar] [CrossRef]
  39. Zhong, Q.; Zhao, Z.; Zhang, Z. Development of “Fractography” and Research of Fracture Micromechanism. Jixie Qiangdu 2005, 27, 358–370. [Google Scholar]
  40. Wang, Z.; Lin, X.; Kang, N.; Chen, J.; Tan, H.; Feng, Z.; Qin, Z.; Yang, H.; Huang, W. Laser powder bed fusion of high-strength Sc/Zr-modified Al–Mg alloy: Phase selection, microstructural/mechanical heterogeneity, and tensile deformation behavior. J. Mater. Sci. Technol. 2021, 95, 40–56. [Google Scholar] [CrossRef]
  41. Cini, A.; Irving, P.E. Development of fatigue cracks from mechanically machined scratches on 2024-T351 aluminium alloy—Part I: Experimentation and fractographic analysis. Fatigue Fract. Eng. Mater. Struct. 2017, 40, 776–789. [Google Scholar] [CrossRef]
  42. Zhong, Q.; Zhao, Z. Fractography, 1st ed.; Higher Education Press: Beijing, China, 2006; pp. 146–153. [Google Scholar]
Figure 1. (a) The morphology of Al-Mg-Sc-Zr powder was observed by SEM; (b) the particle size of the powder was tested by a laser particle size analyzer.
Figure 1. (a) The morphology of Al-Mg-Sc-Zr powder was observed by SEM; (b) the particle size of the powder was tested by a laser particle size analyzer.
Metals 14 00890 g001
Figure 2. Heat treatment method of Al-Mg-Sc-Zr alloy.
Figure 2. Heat treatment method of Al-Mg-Sc-Zr alloy.
Metals 14 00890 g002
Figure 3. (a) Al-Mg-Sc-Zr tensile bar processing direction diagram; (b) drawings of Al-Mg-Sc-Zr tensile test rods; (c) Al-Mg-Sc-Zr tensile test bar.
Figure 3. (a) Al-Mg-Sc-Zr tensile bar processing direction diagram; (b) drawings of Al-Mg-Sc-Zr tensile test rods; (c) Al-Mg-Sc-Zr tensile test bar.
Metals 14 00890 g003
Figure 4. Morphology of Al-Mg-Sc-Zr alloy by OM. (a) 100× magnification; (b) 200× magnification.
Figure 4. Morphology of Al-Mg-Sc-Zr alloy by OM. (a) 100× magnification; (b) 200× magnification.
Metals 14 00890 g004
Figure 5. (a) Morphology of YZ plane after corrosion by OM; (b) morphology of XY plane after corrosion by OM; (c) 3d metallographic diagram of test block; (d) Al-Mg-Sc-Zr test block.
Figure 5. (a) Morphology of YZ plane after corrosion by OM; (b) morphology of XY plane after corrosion by OM; (c) 3d metallographic diagram of test block; (d) Al-Mg-Sc-Zr test block.
Metals 14 00890 g005
Figure 6. (a) X-ray diffraction patterns of the original powder and LPBF alloy block; (b) 2θ amplification region from 37.5° to 39.0°.
Figure 6. (a) X-ray diffraction patterns of the original powder and LPBF alloy block; (b) 2θ amplification region from 37.5° to 39.0°.
Metals 14 00890 g006
Figure 7. EBSD results of Al-Mg-Sc-Zr alloy. (a) IPF graph; (b) pole figure.
Figure 7. EBSD results of Al-Mg-Sc-Zr alloy. (a) IPF graph; (b) pole figure.
Metals 14 00890 g007
Figure 8. SEM images of Al-Mg-Sc-Zr alloy processed by LPBF. (a) Entire melt pool; (b) melt pool boundary.
Figure 8. SEM images of Al-Mg-Sc-Zr alloy processed by LPBF. (a) Entire melt pool; (b) melt pool boundary.
Metals 14 00890 g008
Figure 9. (a) HRTEM-BF-S images and selected-area electron diffraction pattern corresponding to the image; (bd) corresponding EDS elemental distribution maps.
Figure 9. (a) HRTEM-BF-S images and selected-area electron diffraction pattern corresponding to the image; (bd) corresponding EDS elemental distribution maps.
Metals 14 00890 g009
Figure 10. The fine grain zone at the boundary of the molten pool was observed under HRTEM.
Figure 10. The fine grain zone at the boundary of the molten pool was observed under HRTEM.
Metals 14 00890 g010
Figure 11. Al-Mg-Sc-Zr stress–strain curve. (a) Tensile test of AA specimens and AB specimens at room temperature; (b) tensile test of AA specimens at high temperature.
Figure 11. Al-Mg-Sc-Zr stress–strain curve. (a) Tensile test of AA specimens and AB specimens at room temperature; (b) tensile test of AA specimens at high temperature.
Metals 14 00890 g011
Figure 12. Tensile fracture of test rod. (a) AA-RT; (b) regional enlargement of (a); (c) 100 °C; (d) 200 °C.
Figure 12. Tensile fracture of test rod. (a) AA-RT; (b) regional enlargement of (a); (c) 100 °C; (d) 200 °C.
Metals 14 00890 g012
Figure 13. Tensile fracture of AA-H test rod. (a) The fiber zone of fracture at 100 °C; (b) dimple morphology of fracture at 100 °C; (c) the fiber zone of fracture at 200 °C; (d) dimple morphology of fracture at 200 °C.
Figure 13. Tensile fracture of AA-H test rod. (a) The fiber zone of fracture at 100 °C; (b) dimple morphology of fracture at 100 °C; (c) the fiber zone of fracture at 200 °C; (d) dimple morphology of fracture at 200 °C.
Metals 14 00890 g013
Figure 14. Schematic diagram of straight-shaped dimples processed by Al3(Sc, Zr) particles.
Figure 14. Schematic diagram of straight-shaped dimples processed by Al3(Sc, Zr) particles.
Metals 14 00890 g014
Table 3. Al-Mg-Sc-Zr tensile test data.
Table 3. Al-Mg-Sc-Zr tensile test data.
YS (MPa)UTS (MPa)ELO (%)E (GPa)
AB-Z-RT32140421.369.9
AB-H-RT35241223.672.2
AA-Z-RT52755111.270.7
AA-H-RT53556011.165.5
AA-H-100 °C41743525.863.6
AA-H-150 °C21426920.162.1
AA-H-200 °C7710247.938.5
AA-H-250 °C525472.236.2
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

Ren, Z.; Zhang, H.; Shu, X.; Xu, H.; Chen, S.; Ding, Y.; Liang, L.; Qiu, Z.; Yang, Y.; Zheng, Y. Study on Microstructure and High-Temperature Mechanical Properties of Al-Mg-Sc-Zr Alloy Processed by LPBF. Metals 2024, 14, 890. https://doi.org/10.3390/met14080890

AMA Style

Ren Z, Zhang H, Shu X, Xu H, Chen S, Ding Y, Liang L, Qiu Z, Yang Y, Zheng Y. Study on Microstructure and High-Temperature Mechanical Properties of Al-Mg-Sc-Zr Alloy Processed by LPBF. Metals. 2024; 14(8):890. https://doi.org/10.3390/met14080890

Chicago/Turabian Style

Ren, Zhihao, Hao Zhang, Xuedao Shu, Haijie Xu, Siyuan Chen, Yaoyao Ding, Liwen Liang, Zixiang Qiu, Yang Yang, and Yongjian Zheng. 2024. "Study on Microstructure and High-Temperature Mechanical Properties of Al-Mg-Sc-Zr Alloy Processed by LPBF" Metals 14, no. 8: 890. https://doi.org/10.3390/met14080890

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