Next Article in Journal
Microstructural and Mechanical Characterization of Low-Alloy Fire- and Seismic-Resistant H-Section Steel
Previous Article in Journal
Effect of the Fe/Mn Ratio on the Microstructural Evolution of the AA6063 Alloy with Homogenization Heat Treatment Interruption
Previous Article in Special Issue
The Microstructure and Magnetic Properties of a Soft Magnetic Fe-12Al Alloy Additively Manufactured via Laser Powder Bed Fusion (L-PBF)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Aging Temperature on Microstructure, Mechanical, and Wear Properties of 18Ni-300 Maraging Steel Produced by Powder Bed Fusion

1
Department of Advanced Materials Engineering, Dong-Eui University, Busan 47340, Republic of Korea
2
Center for Brain Busan 21 Plus Program, Dong-Eui University, Busan 47340, Republic of Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Metals 2024, 14(4), 375; https://doi.org/10.3390/met14040375
Submission received: 29 February 2024 / Revised: 19 March 2024 / Accepted: 22 March 2024 / Published: 23 March 2024

Abstract

:
Additive manufacturing technologies for metallic materials based on powder bed fusion have enormous industrial potential. In this study, we manufactured 18Ni-300 maraging steel using the powder bed fusion (PBF) process and investigated the effects of annealing temperatures of 430 °C, 490 °C, and 550 °C for 3 h on its microstructure, tensile fracture mechanism, and wear properties compared with the as-built specimen. The results show that annealing heat treatment effectively improved the dry sliding friction, wear properties, and room temperature tensile properties compared to the as-built specimen. Compared to other aging-treated samples, specimens that underwent heat treatment in optimal settings had superior properties. With optimal heat treatment, while melt pool boundaries remained, the cellular and columnar structures became finer compared to the un-treated specimens, and the number of dimples decreased. Consequently, the hardness and tensile strength improved by approximately 56.17% and 40.63%, respectively. The 18Ni-300 maraging steel sample that underwent heat treatment at optimal settings exhibited a coefficient of friction approximately 33.33% lower than the as-built alloy.

1. Introduction

18Ni-300 maraging steel is an iron-nickel-based alloy that contains 18% nickel. The term “maraging” was coined from a combination of “martensite” and “age hardening”. Its mechanical properties, including yield and tensile strength, ductility, toughness, fatigue limit, compressive strength, hardness, and wear resistance, are excellent. Maraging steel is used in aerospace components, extrusion, plastics injection molds, and metal casting dies where abrasion resistance is required. Currently, 18Ni-300 maraging steel parts are mainly produced through casting, forging, and welding processes, which have drawbacks such as prolonged production times and difficulties in fabricating complex-shaped parts [1,2,3,4]. The production of small, custom-made high-precision parts directly from metallic powder through additive manufacturing (AM) has become a primary focus of current research [5]. AM technologies offer engineers a new array of possibilities for designing and fabricating products that would not be feasible or cost-effective with conventional methods such as machining, injection molding, or casting [6]. Among the various additive manufacturing techniques, powder bed fusion (PBF) is one of the most commonly used methods in the industry. The powder bed fusion process has experienced significant growth in recent years due to its cost-effectiveness and high-quality output. Maraging steel powder is very compatible to be processed by the PBF process due to its low reflectivity and high weldability. Parts can be manufactured with nearly full density (99.5%) and exhibit mechanical properties similar to those of conventionally produced metals.
During the PBF process, metal parts are built layer by layer. The orientation of grain texture is influenced by the various building directions of specimens, resulting in anisotropy in the microstructure. For instance, the grains are elongated and columnar in the plane along the building direction, while they are typically equiaxed in the plane perpendicular to the building direction. Therefore, the performance of the parts manufactured using PBF depends on their building direction [7]. Shamsdini et al. [8] examined the microstructure and mechanical properties of 18Ni-300 maraging steel produced by LPBF, and they investigated the effects of building direction (vertical and horizontal). They found that the porosity decreased after a 6 h aging treatment at 490 °C compared to the as-built state. Additionally, the horizontal directions exhibited higher strength and ductility compared to the vertical orientations. They also claimed that heat treatment affects the fracture mechanism, but the building direction does not affect fracture in both the as-built and heat-treated conditions. Song, Tang, et al. [9] fabricated 18Ni-300 maraging steel using PBF in both horizontal and vertical orientations. They conducted microhardness and tensile tests to analyze the mechanical properties and anisotropic behavior of the material. Under the same heat treatment conditions, no significant anisotropic properties were observed. However, the yield strength and ultimate tensile strength of the horizontally fabricated specimens were higher than those of the vertically fabricated ones. This was attributed to the formation of columnar grains along the build direction during PBF.
It is reported that in the 18Ni-300 maraging steel, the Ni3Ti phase (or more generally, Ni3X where X = Ti, Mo, V, and W) readily forms during short-term aging at low temperatures (400–450 °C), followed by the precipitation of Fe2Mo or Fe7Mo6. Aging at temperatures above 500 °C facilitates the simultaneous formation of austenite through a diffusion-controlled reaction. This process is also facilitated by the release of nickel into the matrix due to the decomposition of the Ni3Ti phase [10]. In the PBF process, a previously deposited material may undergo cyclic reheating with a gradually decreasing laser intensity during the deposition of neighboring tracks and subsequent layers. This indicates that the material deposited in one layer undergoes in situ heat treatment by the subsequent tracks and layers, allowing the nuclei of the precipitates to form easily during the PBF process [11]. As a result, precipitation can occur easily without undergoing solution heat treatments. In this regard, some researchers have already considered omitting the solution heat treatment for the maraging steel produced by PBF [12]. Through aging treatment, the internal stress caused by quenching can be released, and evenly distributed nano precipitates would form in the substrate, thereby enhancing both strength and toughness [13]. Therefore, optimizing the aging temperature and time is crucial for achieving the desired properties. Aging treatment at an appropriate temperature leads to the hardening of the matrix as a result of the precipitation of Ni-Mo, Fe-Mo, and Fe-Ni intermetallic compounds within the martensitic structure. Casati et al. [10] conducted a study on the microstructural changes, hardness, and tensile properties of PBFed 18Ni-300 steel at different aging temperatures and durations. Using differential scanning calorimetry (DSC), isothermal aging curves were measured at temperatures of 460 °C, 490 °C, 540 °C, and 600 °C for durations ranging from 10 min to 14 days. During the aging treatment, it was discovered that austenite reversion initially occurred at the cell boundary, with austenite appearing at the intercellular level during over-aging. Kim. D, Kim. T, et al. [12] investigated the effect of heat treatment on the mechanical anisotropy of PBFed 18Ni-300 maraging steel. When heat-treated at 450 °C, similar hardness values to those of fully hardened martensitic steel were observed. As a result, it was claimed that the alloy can be hardened only through aging treatment, without the need for solution treatment. The specimens aged at 450 °C exhibited the highest yield strength when the building direction was horizontal. Yan et al. [14] studied the microstructure and mechanical properties of additively manufactured 18Ni-300 maraging steel through solution treatment and aging treatment. The aging-treated specimens had higher hardness than the solubilization-treated specimens. They also found that the hardness increased with the increase in aging time but decreased after exceeding 3 h due to coarsening of precipitates.
Several studies have been conducted on the microstructural evolution resulting from heat treatment. However, there is a lack of analysis regarding the tensile properties and wear behavior of 18Ni-300 maraging steel in relation to different heat treatment temperatures. This study investigates the effects of building direction and aging heat treatment conditions on the microstructure, tensile properties, fracture mechanisms, and wear properties of 18Ni-300 maraging steel. In particular, wear reduces the lifespan of components that operate at high speeds and high loads and requires in-depth analysis of wear behavior for industrial applications. The 18Ni-300 maraging steel produced by the PBF process underwent thermal treatment, including aging treatment at various temperatures followed by air cooling. Additionally, tensile tests, hardness evaluations, wear tests, and microstructure inspections were conducted on the thermally treated specimens to find the most optimal aging conditions.

2. Experimental Procedure

18Ni-300 maraging steel powder with a particle size of 10–90 µm was used, and the chemical composition is shown in Table 1. The Concept Laser M1 was utilized to manufacture wear and tensile specimens, as depicted in Figure 1, in an argon atmosphere. The machine specifications include a laser power of 180 W, scan speed of 1000 mm/s, hatch distance of 0.105 mm, beam diameter of 0.05 mm, layer thickness of 0.035 mm, focus move of −3 mm, rotating angle of 90°, and a laser pattern of continuous exposure. The specimens were aged at 430 °C, 490 °C, and 550 °C for 3 h without solution treatment and then cooled in air, as shown in Figure 1c.
After heat treatment, the microstructure was observed by grinding with SiC paper (80–600 grit), polishing with a 3 µm diamond suspension, and further polishing with colloidal silica down to 0.04 µm before wear testing. The etching solution was Fry’s reagent (25 mL of HNO3 + 50 mL of HCl + 1 g of CuCl2 + 150 mL of H2O). Measurements were taken 12 times for each specimen with a 0.1 kgf load using a Vickers hardness tester (HM-210, Mitutoyo, Tokyo, Japan). XRD analysis (SC-XRD, D8 VENTURE, Bruker, Tokyo, Japan) was conducted to examine the phases at various aging temperatures. It was measured at 40 kV and 40 mA within a 2θ range of 20° to 90° using CuKα radiation at a scan rate of 2°/min. The sample was then subjected to Rietveld analysis. The tribological properties were measured using the ball-on-disc method with an abrasion tester (RB-102PD, R&B Co., Ltd., Seoul, Republic of Korea) at room temperature. The test was performed for 99.645 s at 50 N, with a rotational speed of 1000 RPM and a sliding distance of 120 m using a tungsten carbide (WC) ball, which has higher hardness compared to the specimen. Tensile testing was conducted at room temperature using a tensile testing machine (KSU-20M, Kyoungsung, Seoul, Republic of Korea) with a load of 50 N and a deformation rate of 5 mm/min. The field emission scanning electron microscope (FE-SEM, Quanta 200 FEG, FEI Company, Amsterdam, The Netherlands) at the Converging Materials Core Facility of Dong-Eui University was used to observe the microstructure, wear tracks, mechanisms, and tensile fracture surface.

3. Results and Discussion

Figure 2 displays scanning electron microscope (SEM) observations of microstructure changes based on aging temperature. Figure 2a shows the microstructure of the as-built specimen without heat treatment. Meltpool boundaries are observed, and columnar and cellular structures appear. These fine cellular microstructures are unique to the PBF process. They form in response to the instantaneous melting and rapid solidification with an extremely high cooling rate of the powder alloy during laser irradiation. The refined cellular structure is a common microstructure of PBFed steels, which can enhance the hardness and strength of PBF-produced steels compared to conventionally manufactured steels [15]. Figure 2b–d shows the microstructure after annealing heat treatment. As shown in Figure 2b, the morphology was similar to the as-built specimen when annealed at 430 °C. Meltpool boundaries were observed, and the cellular and columnar structure was dissolved by the heat treatment and not clearly visible. The black dots in Figure 2b,c are estimated to be pores, and the number of pores was higher when heat treated at 490 °C. For heat treatments above 490 °C, the meltpool boundaries disappeared, as shown in Figure 2c,d. Cellular and columnar structures disappeared, and an irregular, island-like microstructure was observed. Such sub-grain cells may disappear through diffusion during the post-heat treatment process. In addition, the thickness of the cells increased as the aging temperature increased, indicating an over-aging phenomenon. The microstructure exhibited very thick boundaries, as shown in the figure, due to the extremely prominent reversion of martensite to austenite. The retained austenite resulted from the microsegregation of solute elements, especially nickel, at cellular boundaries during solidification [15]. Retained austenite is easily distinguishable in the high magnification microstructure, where it appears as a bright phase that aggregates at cell boundaries. It arises in the as-built alloy from an incomplete transformation into martensite during rapid cooling or solidification. The presence of these tiny segregated solute-rich regions not only softens the as-built alloy but also enhances additional austenite reversion during the high temperature precipitation heat treatment cycle [16].
Figure 3 illustrates the XRD pattern changes depending on the aging temperature. The (110)α, (200)α, and (211)α peaks, which are martensite phases, are observed in the as-built and all heat treatment conditions. However, no austenite phase appeared in the as-built specimen. When heat-treated at 430 °C, the intensity of the martensite peaks decreased, and in the austenite phase, the (111)γ peak appeared. The (111)γ peak was also present at 490 °C, and an additional (002)γ peak was formed. At 550 °C, the (022)γ peak formed, and the intensity of the existing (111)γ and (002)γ peaks increased. The peak of the austenite phase was observed during the aging treatment. As the aging temperature increased, the peak intensity of martensite decreased, while the peak intensity of austenite increased. This indicates that the amount of austenite increased as the aging temperature increased. Austenite increases during aging due to the enrichment of the matrix containing Ni, which is called reverted austenite. Precipitates such as Ni3X (X = Al, Ti, Mo, Fe) are present at 20° to 40°, but they are difficult to distinguish because of their small size and low volume fraction [17,18,19]. Table 2 presents the quantitative analysis of the volume fraction of austenite and martensite as they vary with aging temperature, utilizing the Rietveld method. Previous literature has shown a low fraction of austenite in untreated specimens [20,21]. In this study, the austenite fraction was 0%, and only martensite was present. This is estimated to be due to the very low content of austenite, which did not appear in the XRD pattern. The austenite fraction was the lowest at 1.2% when aged at 430 °C and the highest at 21.8% when aged at 550 °C. The amount of austenite increased significantly to 12.9% during the heat treatment at 490 °C, indicating that overaging has occurred [22]. As the aging temperature increased, the fraction of austenite increased, but the fraction of martensite decreased.
Figure 4 shows the results of Vickers hardness measurements varying with heat treatment temperature. The differences between the heat treatment conditions were minimal, but the hardness of the specimens subjected to heat treatment increased significantly compared to the as-built specimens. The as-built specimens, which were not subjected to heat treatment, have the lowest hardness at 380.3 ± 10.8 HV due to the absence of precipitate formation [23]. In the heat treatment process, the highest hardness value of 593.9 ± 12.5 HV was obtained at 430 °C. Hardness values of 575.6 ± 7.6 HV and 558.7 ± 8.9 HV were measured for specimens heat-treated at 490 °C and 550 °C, respectively. The hardness increased with heat treatment and decreased with higher aging temperatures. The hardness decreased with the increasing thickness of the bright phase (austenite), as shown in Figure 2d. The Rietveld results show that the highest hardness values were obtained when the amount of austenite was 1.2%, with a sharp decrease in hardness above this amount. Therefore, the presence of austenite affects the hardness, which decreases as it dissolves intermetallic precipitates [24].
Figure 5 compares the tensile strength and elongation at different annealing temperatures of 18Ni-300 maraging steel produced by PBF. The as-built specimen without heat treatment had the lowest tensile and yield strengths of 960 MPa and 740 MPa, respectively, but the highest elongation at break of 10.88%. The tensile strength at 430 °C was 1350 MPa, which was similar to the value obtained when heat-treated at 550 °C (1340 MPa). However, the yield strength after heat treatment at 430 °C was 870 MPa, which was 120 MPa higher than the specimen heat-treated at 550 °C. Heat treatment at 490 °C resulted in a tensile strength of 1400 MPa, which was a 45.83% increase over the as-built condition and represented the highest tensile strength among the heat-treated conditions. However, the yield strength was lower than that of the specimen heat-treated at 430 °C. After the heat treatment, the tensile and yield strength increased, but the elongation decreased. The difference in tensile strength with aging temperature was not significant. This is likely due to the small variation in heat treatment temperature. However, the yield strength decreased with increasing aging temperature. The intensity of the austenite peak increased as the aging temperature increased, as shown in Figure 3. Enhancing the austenite content can lead to improved elongation but may result in a decrease in strength [25,26]. In addition, the yield strength of the metal was inversely proportional to the size of the cellular structure, and these structures contribute to improving the yield strength of AM parts [27,28]. Therefore, the highest elongation and lowest tensile strength were obtained at 550 °C, where the amount of austenite was the highest at 21.8%. At 430 °C, the finest cellular structure was observed, with an optimal combination of martensite and austenite structures, resulting in the highest yield strength. Heat treatment was performed at similar temperatures as in previous studies, but the tensile and yield strengths were lower [29]. However, high mechanical properties were achieved at lower temperatures compared to the original study.
In Figure 6, the fracture surface after tensile testing was observed using SEM. All specimens in Figure 6 exhibit mixed fracture behavior, with both ductile and brittle fractures observed. In ductile fracture, dimples are observed, which are created by the coalescence of micropores. The smaller and more numerous the dimples, the less ductile and more brittle they become when broken. In brittle fracture, cracks propagate rapidly in a specific direction, leading to cleavages and river patterns [30,31]. The as-built specimen (Figure 6a) has the highest elongation due to the highest number of dimples and larger dimples with a river pattern compared to other heat treatments. Figure 6b shows fewer dimples compared to the as-built sample, and transgranular fracture occurred with crack propagation across the grain. Figure 6c, which was heat-treated at 490 °C, exhibited more brittle fracture with increased transgranular fractures compared to 430 °C, and cracks were observed. In addition, the river patterns became more numerous and distinct. Figure 6d shows that the number of dimples is the highest among the heat-treated specimens, and large dimples are observed. Therefore, the ductility is better than that of the specimens heat-treated at 430 °C and 490 °C, resulting in a more ductile fracture. In addition, a small number of transgranular fractures occurred. Similar to the other specimens, cracks and river patterns were observed, although the river patterns were faintly visible. Transgranular fracture was not observed in the as-built specimen, but it occurred in the heat-treated specimen. Additionally, the dimple size increased as the aging temperature rose.
Figure 7 shows SEM images of the width of the wear track after the wear test. Debris was observed around the wear tracks of all specimens. Under a certain load, abrasives move across the surface of a material, producing wear debris. Material was removed from the wear track and pushed to the sides of the track due to the plowing effect. Successive layers of materials were squeezed out with each pass of the counterpart movement to accommodate the wear. It can be observed that material from the wear track is extruded, forming plate-like debris due to repeated sliding. The plates then fracture and break away as debris [32,33]. In Figure 7a, numerous large and small debris are observed at the edge of the wear track, with the widest wear width measuring 1190 µm from the as-built specimen. However, specimens subjected to annealing heat treatment at 430 °C (Figure 7b) show that only a few small debris are observed around the wear track, and the wear width is the narrowest at 854 µm. It can be seen from Figure 7c that the specimens annealed at 490 °C and 550 °C have an average wear width of 1010 µm, and Figure 7d is 1050 µm. The degree of specimen adhesion around the wear track is similar, but the specimens heat-treated at 490 °C and 550 °C exhibit wider wear tracks. The heat-treated specimens tended to have a narrower wear width compared to the untreated specimens. In addition, the amount of debris and the width of wear around the wear track decreased with increasing hardness.
In Figure 8, observations were made at high magnification to analyze the wear mechanism, and the average friction coefficients are shown in Figure 9. There are wear mechanisms including abrasive wear, adhesive wear, fatigue wear, and fretting wear. This is influenced by factors such as the applied load, sliding velocity, surface hardness, roughness, lubrication, and other related variables [34,35]. In Figure 8a, more adhesive wear is observed compared to abrasive wear. The average friction coefficient is the highest at 0.453 µ because the ball does not slide well due to the adhered part. On the other hand, in Figure 8b, which was heat-treated at 430 °C, a significant amount of grinding wear with grooves was observed. The width of the groove was the narrowest among the specimens, and delamination occurred. Delamination was more prevalent in specimens with heat treatment than in those without heat treatment. The EDS results showed a high content of oxygen and tungsten in the area affected by adhesive wear. The presence of W is believed to be caused by the WC ball counterpart detaching during the wear test. In addition, the formation of these oxides is associated with reduced wear and friction. Therefore, the oxide present on the wear surface in Figure 8b has the lowest average friction coefficient of 0.302 µ as the friction force is reduced due to the oxide on the wear surface. Figure 8c,d exhibits similar mechanisms involving a combination of abrasive and adhesive wear. Abrasive wear typically occurs when softer surfaces come into contact with rough, hard protrusions, while adhesive wear takes place when two essentially flat solid surfaces are in sliding contact, with or without lubrication [36,37]. The average coefficient of friction in Figure 8c is 0.349 µ, while in Figure 8d it is 0.375 µ, which is the highest among the heat-treated specimens. Figure 8c exhibits higher hardness compared to Figure 8d, leading to increased abrasive wear and observed galling. Figure 8d also shows significant galling due to adhesive wear, with the widest groove width and the least delamination. As the heat treatment temperature decreased, abrasive wear increased, leading to more delamination and a reduction in groove width. In addition, the average coefficient of friction (Figure 9) exhibited a sharp decrease with heat treatment and increased with higher heat treatment temperatures due to the presence of adhesive wear.

4. Conclusions

In this study, 18Ni-300 maraging steel was processed using powder bed fusion (PBF) with optimized processing parameters. The effect of various heat treatment conditions on the microstructure, mechanical properties, and wear behavior of 18Ni-300 maraging steel was analyzed. The microstructure of the as-built sample exhibits columnar and cellular structures within the boundaries of the melt pool. When annealed at temperatures above 490 °C, the boundaries of the melt pool disappeared. The average microhardness of the samples in the as-built state and after aging at 430 °C, 490 °C, and 550 °C are 380.3, 593.9, 575.6, and 558.7 HV, respectively. The austenite present at the grain boundaries precipitated during the aging treatment, leading to an increase in cell thickness and an irregular microstructure. Although the microhardness could be improved by aging, the hardness decreased due to excessive precipitation of austenite when heat-treated above 490 °C. Aging also affects the transformation of martensite to austenite. The austenite phase fraction analyzed by the Rietveld method was 1.2%, 12.9%, and 21.8% at 430 °C, 490 °C, and 550 °C. The austenite fraction increased with the rise in aging temperature. Specimens aged at 430 °C exhibited UTS of 1350 MPa, which is nearly 40.63% higher than that of the as-built specimen and 0.75% higher than the specimen aged at 550 °C, attributed to the finer microstructure and moderate amount of austenite. The as-built specimens exhibited the highest elongation because of the distribution of large dimples. The heat-treated specimens had both ductile and brittle fractures, with a higher occurrence of brittle fracture observed at 490 °C. The average coefficient of friction (COF) of the sample, heat-treated at 430 °C (COF = 0.302 µ), is reduced by 33.33% compared to the as-built state. The cause of the reduced friction was the high oxygen content in the area affected by adhesive wear. This results in excellent wear resistance when aged at 430 °C. Adhesive wear is the dominant wear mechanism in the as-built samples, while the heat-treated specimen exhibits a mixed mode of abrasive and adhesive wear. The optimal heat treatment conditions we found (430 °C, 3 h) are expected to enhance mechanical and wear properties when implemented in the industry.

Author Contributions

Conceptualization, I.J.; Data curation, S.H., C.J. and Y.L.; Formal analysis, N.K. and S.H.; Funding acquisition, I.J.; Investigation, N.K., C.J. and Y.L.; Methodology, S.H. and C.J.; Project administration, I.J.; Supervision, I.J.; Visualization, I.J.; Writing—original draft, N.K. and Y.L.; Writing—review and editing, I.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (2021R1I1A3057115).

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Floreen, S. The Physical Metallurgy of Maraging Steels. Metall. Rev. 1968, 13, 115–128. [Google Scholar] [CrossRef]
  2. Casalino, G.; Campanelli, S.L.; Contuzzi, N.; Ludovico, A.D. Experimental Investigation and Statistical Optimisation of the Selective Laser Melting Process of a Maraging Steel. Opt. Laser Technol. 2015, 65, 151–158. [Google Scholar] [CrossRef]
  3. Hermann Becker, T.; Dimitrov, D. The Achievable Mechanical Properties of SLM Produced Maraging Steel 300 Components. Rapid Prototyp. J. 2016, 22, 487–494. [Google Scholar] [CrossRef]
  4. Zhang, S.; Zhou, N.; Ma, C.; Liu, J.; Liu, S.; Misra, R.D.K. Simultaneous Enhancement of Strength and Hydrogen Embrittlement Resistance of Laser-Powder Bed Fusion Maraging Steel via Long-Term Low-Temperature Aging. Corros. Sci. 2023, 223, 111440. [Google Scholar] [CrossRef]
  5. Jägle, E.A.; Choi, P.-P.; Humbeeck, J.V.; Raabe, D. Precipitation and austenite reversion behavior of a maraging steel produced by selective laser melting. J. Mater. Res. 2014, 29, 2072–2079. [Google Scholar] [CrossRef]
  6. Chen, T.; Shea, K. Computational design-to-fabrication using spatial grammars: Automatically generating printable car wheel design variants. Des. Soc. A Worldw. Community 2015, 35–44. Available online: https://www.designsociety.org/publication/37710/COMPUTATIONAL+DESIGN-TO-FABRICATION+USING+SPATIAL+GRAMMARS%3A+AUTOMATICALLY+GENERATING+PRINTABLE+CAR+WHEEL+DESIGN+VARIANTS (accessed on 20 February 2024).
  7. Dong, Z.; Ouyang, P.; Zhang, S.; Liu, L.; Li, H.; Wu, Y. Effect of Building Direction on Anisotropy of Mechanical Properties of GH4169 Alloy Fabricated by Laser Powder Bed Fusion. Mater. Sci. Eng. A 2023, 862, 144430. [Google Scholar] [CrossRef]
  8. Shamsdini, S.; Pirgazi, H.; Ghoncheh, M.H.; Sanjari, M.; Amirkhiz, B.S.; Kestens, L.; Mohammadi, M. A Relationship between the Build and Texture Orientation in Tensile Loading of the Additively Manufactured Maraging Steels. Addit. Manuf. 2021, 41, 101954. [Google Scholar] [CrossRef]
  9. Song, J.; Tang, Q.; Feng, Q.; Ma, S.; Setchi, R.; Liu, Y.; Han, Q.; Fan, X.; Zhang, M. Effect of Heat Treatment on Microstructure and Mechanical Behaviours of 18Ni-300 Maraging Steel Manufactured by Selective Laser Melting. Opt. Laser Technol. 2019, 120, 105725. [Google Scholar] [CrossRef]
  10. Casati, R.; Lemke, J.N.; Tuissi, A.; Vedani, M. Aging Behaviour and Mechanical Performance of 18-Ni 300 Steel Processed by Selective Laser Melting. Metals 2016, 6, 218. [Google Scholar] [CrossRef]
  11. Tan, C.; Zhou, K.; Tong, X.; Huang, Y.; Li, J.; Ma, W.; Li, F.; Kuang, T. Microstructure and Mechanical Properties of 18Ni-300 Maraging Steel Fabricated by Selective Laser Melting; Atlantis Press: Amsterdam, The Netherlands, 2017; pp. 404–410. [Google Scholar]
  12. Kim, D.; Kim, T.; Ha, K.; Oak, J.-J.; Jeon, J.B.; Park, Y.; Lee, W. Effect of Heat Treatment Condition on Microstructural and Mechanical Anisotropies of Selective Laser Melted Maraging 18Ni-300 Steel. Metals 2020, 10, 410. [Google Scholar] [CrossRef]
  13. Wang, S.; Zhao, T.; Liu, Z.; Du, C.; Li, X. Exploration of the Processing Scheme of a Novel Ni(Fe, Al)-Maraging Steel. J. Mater. Res. Technol. 2021, 10, 225–239. [Google Scholar] [CrossRef]
  14. Yan, J.; Meng, X.; Ou, B.; Xie, Y.; Cai, B.; Zhang, Y.; Fang, S. Dependence of Microstructure and Properties in Additive Manufactured 18Ni300 on Heat Treatment and Surface Enhancement. J. Mater. Res. Technol. 2024, 29, 969–981. [Google Scholar] [CrossRef]
  15. Yin, S.; Chen, C.; Yan, X.; Feng, X.; Jenkins, R.; O’Reilly, P.; Liu, M.; Li, H.; Lupoi, R. The Influence of Aging Temperature and Aging Time on the Mechanical and Tribological Properties of Selective Laser Melted Maraging 18Ni-300 Steel. Addit. Manuf. 2018, 22, 592–600. [Google Scholar] [CrossRef]
  16. Mooney, B.; Kourousis, K.I. A Review of Factors Affecting the Mechanical Properties of Maraging Steel 300 Fabricated via Laser Powder Bed Fusion. Metals 2020, 10, 1273. [Google Scholar] [CrossRef]
  17. Nouri, N.; Li, Q.; Damon, J.; Mühl, F.; Graf, G.; Dietrich, S.; Schulze, V. Characterization of a Novel Maraging Steel for Laser-Based Powder Bed Fusion: Optimization of Process Parameters and Post Heat Treatments. J. Mater. Res. Technol. 2022, 18, 931–942. [Google Scholar] [CrossRef]
  18. Bae, K.; Shin, D.; Kim, J.-H.; Lee, W.; Jo, I.; Lee, J. Influence of Post Heat Treatment Condition on Corrosion Behavior of 18Ni300 Maraging Steel Manufactured by Laser Powder Bed Fusion. Micromachines 2022, 13, 1977. [Google Scholar] [CrossRef] [PubMed]
  19. Guo, Z.; Sha, W. Quantification of Precipitate Fraction in Maraging Steels by X-ray Diffraction Analysis. Mater. Sci. Technol. Mater. Sci. Technol. 2004, 20, 126–130. [Google Scholar] [CrossRef]
  20. Calro, M. Microstructural and Mechanical Properties of Maraging Steel Parts Produced by Selective Laser Melting. Master’s Thesis, School of Industrial and Information Engineering, Politecnico di Milano, Milan, Italy, 2015. [Google Scholar]
  21. Kempen, K.; Yasa, E.; Thijs, L.; Kruth, J.-P.; Van Humbeeck, J. Microstructure and Mechanical Properties of Selective Laser Melted 18Ni-300 Steel. Phys. Procedia 2011, 12, 255–263. [Google Scholar] [CrossRef]
  22. Zhang, C.; Wang, C.; Zhang, S.L.; Ding, Y.L.; Ge, Q.L.; Su, J. Effect of Aging Temperature on the Precipitation Behavior and Mechanical Properties of Fe–Cr–Ni Maraging Stainless Steel. Mater. Sci. Eng. A 2021, 806, 140763. [Google Scholar] [CrossRef]
  23. Herzog, D.; Seyda, V.; Wycisk, E.; Emmelmann, C. Additive Manufacturing of Metals. Acta Mater. 2016, 117, 371–392. [Google Scholar] [CrossRef]
  24. Pardal, J.M.; Tavares, S.S.M.; Cindra Fonseca, M.P.; da Silva, M.R.; Neto, J.M.; Abreu, H.F.G. Influence of Temperature and Aging Time on Hardness and Magnetic Properties of the Maraging Steel Grade 300. J. Mater. Sci. 2007, 42, 2276–2281. [Google Scholar] [CrossRef]
  25. Bai, Y.; Wang, D.; Yang, Y.; Wang, H. Effect of Heat Treatment on the Microstructure and Mechanical Properties of Maraging Steel by Selective Laser Melting. Mater. Sci. Eng. A 2019, 760, 105–117. [Google Scholar] [CrossRef]
  26. Rohit, B.; Muktinutalapati, N.R. Austenite Reversion in 18% Ni Maraging Steel and Its Weldments. Mater. Sci. Technol. 2018, 34, 253–260. [Google Scholar] [CrossRef]
  27. Kong, D.; Dong, C.; Wei, S.; Ni, X.; Zhang, L.; Li, R.; Wang, L.; Man, C.; Li, X. About Metastable Cellular Structure in Additively Manufactured Austenitic Stainless Steels. Addit. Manuf. 2021, 38, 101804. [Google Scholar] [CrossRef]
  28. Kong, D.; Dong, C.; Ni, X.; Liang, Z.; Man, C.; Li, X. Hetero-Deformation-Induced Stress in Additively Manufactured 316 L Stainless Steel. Mater. Res. Lett. 2020, 8, 390–397. [Google Scholar] [CrossRef]
  29. Mei, X.; Yan, Y.; Fu, H.; Gao, X.; Huang, S.; Qiao, L. Effect of Aging Temperature on Microstructure Evolution and Strengthening Behavior of L-PBF 18Ni(300) Maraging Steel. Addit. Manuf. 2022, 58, 103071. [Google Scholar] [CrossRef]
  30. Wang, W.; Chen, Z.; Lu, W.; Meng, F.; Zhao, T. Heat Treatment for Selective Laser Melting of Inconel 718 Alloy with Simultaneously Enhanced Tensile Strength and Fatigue Properties. J. Alloys Compd. 2022, 913, 165171. [Google Scholar] [CrossRef]
  31. Lei, S.; Xu, Z.; Peng, L.; Lai, X. Effect of Grain Size on the Ductile-Brittle Fracture Behavior of Commercially Pure Titanium Sheet Metals. Mater. Sci. Eng. A 2021, 822, 141630. [Google Scholar] [CrossRef]
  32. Yingjie, L.; Xingui, B.; Keqiang, C. A Study on the Formation of Wear Debris during Abrasion. Tribol. Int. 1985, 18, 107–111. [Google Scholar] [CrossRef]
  33. Mabrouk, A.; Farhat, Z. Novel Ni-P-Tribaloy Composite Protective Coating. Materials 2023, 16, 3949. [Google Scholar] [CrossRef]
  34. Swain, B.; Bhuyan, S.; Behera, R.; Mohapatra, S.S.; Behera, A. Wear: A Serious Problem in Industry. In Tribology in Materials and Manufacturing—Wear, Friction and Lubrication; IntechOpen: London, UK, 2020; ISBN 978-1-83880-575-3. [Google Scholar]
  35. Hong, W.; Cai, W.; Wang, S.; Tomovic, M.M. Mechanical Wear Debris Feature, Detection, and Diagnosis: A Review. Chin. J. Aeronaut. 2018, 31, 867–882. [Google Scholar] [CrossRef]
  36. Wei, M.X.; Chen, K.M.; Wang, S.Q.; Cui, X.H. Analysis for Wear Behaviors of Oxidative Wear. Tribol. Lett. 2011, 42, 1–7. [Google Scholar] [CrossRef]
  37. Zhang, C. 14—Understanding the Wear and Tribological Properties of Ceramic Matrix Composites. In Advances in Ceramic Matrix Composites; Low, I.M., Ed.; Woodhead Publishing: Cambridge, UK, 2014; pp. 312–339. ISBN 978-0-85709-120-8. [Google Scholar]
Figure 1. Schematics of (a) wear test specimen, (b) tensile test specimen (KS B 0801 13B, gauge length 50 mm), and (c) heat treatment conditions.
Figure 1. Schematics of (a) wear test specimen, (b) tensile test specimen (KS B 0801 13B, gauge length 50 mm), and (c) heat treatment conditions.
Metals 14 00375 g001
Figure 2. Microstructure of 18Ni-300 maraging steel with different annealing heat treatment temperature of (a) as-built, (b) 430 °C, (c) 490 °C, and (d) 550 °C.
Figure 2. Microstructure of 18Ni-300 maraging steel with different annealing heat treatment temperature of (a) as-built, (b) 430 °C, (c) 490 °C, and (d) 550 °C.
Metals 14 00375 g002
Figure 3. XRD pattern of PBFed 18Ni-300 maraging steel with varying aging heat treatments.
Figure 3. XRD pattern of PBFed 18Ni-300 maraging steel with varying aging heat treatments.
Metals 14 00375 g003
Figure 4. Vickers hardness at different heat treatment temperatures.
Figure 4. Vickers hardness at different heat treatment temperatures.
Metals 14 00375 g004
Figure 5. Tensile strength, yield strength, and elongation with variations at different aging temperatures.
Figure 5. Tensile strength, yield strength, and elongation with variations at different aging temperatures.
Metals 14 00375 g005
Figure 6. Tensile fracture surface SEM images with different aging temperature of (a) as-built, (b) 430 °C, (c) 490 °C, and (d) 550 °C.
Figure 6. Tensile fracture surface SEM images with different aging temperature of (a) as-built, (b) 430 °C, (c) 490 °C, and (d) 550 °C.
Metals 14 00375 g006
Figure 7. Wear surface analysis on the specimen with aging treatment temperature of (a) as-built, (b) 430 °C, (c) 490 °C, (d) 550 °C.
Figure 7. Wear surface analysis on the specimen with aging treatment temperature of (a) as-built, (b) 430 °C, (c) 490 °C, (d) 550 °C.
Metals 14 00375 g007
Figure 8. Wear mechanism on the specimen with aging treatment temperature of (a) as-built, (b) 430 °C, (c) 490 °C, (d) 550 °C (high magnification).
Figure 8. Wear mechanism on the specimen with aging treatment temperature of (a) as-built, (b) 430 °C, (c) 490 °C, (d) 550 °C (high magnification).
Metals 14 00375 g008
Figure 9. Average friction coefficient of aging treatment temperature.
Figure 9. Average friction coefficient of aging treatment temperature.
Metals 14 00375 g009
Table 1. Chemical composition of 18Ni-300 maraging steel powder.
Table 1. Chemical composition of 18Ni-300 maraging steel powder.
ElementNiCoMoTiAlSiCPSFe
Wt%17.708.664.580.680.0890.010.0043<0.005<0.003Bal.
Table 2. Rietveld analysis of different aging temperatures.
Table 2. Rietveld analysis of different aging temperatures.
Martensite (%)Austenite (%)
as-built1000
430 °C98.81.2
490 °C87.112.9
550 °C78.221.8
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

Kwak, N.; Lim, Y.; Heo, S.; Jeon, C.; Jo, I. Effect of Aging Temperature on Microstructure, Mechanical, and Wear Properties of 18Ni-300 Maraging Steel Produced by Powder Bed Fusion. Metals 2024, 14, 375. https://doi.org/10.3390/met14040375

AMA Style

Kwak N, Lim Y, Heo S, Jeon C, Jo I. Effect of Aging Temperature on Microstructure, Mechanical, and Wear Properties of 18Ni-300 Maraging Steel Produced by Powder Bed Fusion. Metals. 2024; 14(4):375. https://doi.org/10.3390/met14040375

Chicago/Turabian Style

Kwak, Nawon, Yujin Lim, Seokha Heo, Chami Jeon, and Ilguk Jo. 2024. "Effect of Aging Temperature on Microstructure, Mechanical, and Wear Properties of 18Ni-300 Maraging Steel Produced by Powder Bed Fusion" Metals 14, no. 4: 375. https://doi.org/10.3390/met14040375

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