Next Article in Journal
The Effect of Powder Reuse on Electron Beam Melting for Biomedical Implants
Next Article in Special Issue
Enhanced Durability of Wood Cutting Tools through Thermal Cycling
Previous Article in Journal
Molecular Dynamics Analysis of Multi-Factor Influences on Structural Defects in Deposited Mg-Matrix Zn Atom Films
Previous Article in Special Issue
Impact of Size and Distribution of k-Carbides on the Hydrogen Embrittlement and Trapping Behaviors of a Fe-Mn-Al-C Low-Density Steel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress in Laser Powder Bed Fusions Processes of Advanced High-Strength Steels

by
Aleksandra Królicka
* and
Julia Malawska
Department of Metal Forming, Welding and Metrology, Faculty of Mechanical Engineering, Wroclaw University of Science and Technology, 50-370 Wroclaw, Poland
*
Author to whom correspondence should be addressed.
Materials 2024, 17(19), 4699; https://doi.org/10.3390/ma17194699
Submission received: 9 September 2024 / Revised: 18 September 2024 / Accepted: 20 September 2024 / Published: 25 September 2024
(This article belongs to the Special Issue Enhancing In-Use Properties of Advanced Steels)

Abstract

:
This review is focused on the perspectives of the application of Advanced High Strength Steels (AHSSs) in the field of additive technologies directed at the laser powder bed fusion/selective laser melting processes. In principle, AHSSs require significant attention due to their promising mechanical properties for usage in the automotive industry towards reducing the weight of vehicles. Although additive manufacturing represents a promising perspective towards expanding the industrialization of AHSSs in a wider area of their applications, they have not been sufficiently investigated concerning their usage in LPBF/SLM processes. AM techniques enable the fabrication of complex machine parts, including those with a cellular structure, which can contribute to further reducing the weight of vehicles or structures. Maraging steels have recently attracted the attention of researchers, and today are a common grade of steel produced by LPBF techniques. The other group of AHSSs are high-Mn steels with an austenitic matrix characterized by the TRIP and TWIP effects. Less published research has been conducted on medium-Mn steels, which require additional intercritical annealing and preheating during printing. Moreover, the advanced bainitic steels and low-density, high-strength steels represent a new window for further research into the use of the LPBF processes for their fabrication.

Graphical Abstract

1. Introduction

Considering the European and global directives relating to reducing CO2 emissions in the steel sector industry, developing new materials and adjusting advanced technology processes constitute crucial challenges to overcome. The purpose of current trends is focused on achieving high-strength and in-use properties of steels, which lead to the diminution of the structure’s weight and fuel consumption. Consequently, Advanced High-Strength Steels (AHSSs) are a material group of increasing interest in the scientific communities and the industry sector [1]. The scope of AHSSs mainly includes:
  • dual-phase steels [2,3];
  • multi-phase steels [4,5];
  • medium-Mn steels [6,7];
  • high-Mn steels [8,9];
  • twinning-induced plasticity (TWIP) steels [10];
  • transformation-induced plasticity (TRIP) steels [11];
  • nanocrystalline bainitic steels [12,13,14];
  • maraging steels [15].
Current research mainly focuses on controlling in-use properties and facilitating industrialization [16,17,18,19,20,21,22]. AHSSs constitute promising materials for use in the automotive industry considering the associated strong demands [23]. The vehicle parts made of AHSSs are prone to absorb more kinetic energy during the crash and are characterized by higher resistance to intrusion in comparison to low-strength steels [24]. The controlled contribution of the TRIP and TWIP effect constitutes a research direction focused on energy absorption capabilities, and high mechanical properties enable a reduction in vehicle weight. The significant in-use properties cover the main fields [24]: (i) weldability; (ii) coatability, primarily galvanic coatings; (iii) processability; (iv) cost; (v) recyclability. It was also indicated that the current trends for AHSSs constitute the light-weight alloy design strategy and improving the formability without the loss of mechanical properties level [25]. Although AHSSs are characterized by outstanding mechanical properties, their susceptibility to hydrogen embrittlement constitutes one of the particular challenges. For the majority of these steels, the local microstructure spots characterized by high gradients of local stress occur [1].
Over the last decade, there has been a significant increase in interest in additive manufacturing (AM) methods for producing demanding machine parts. Additive technologies provide prominent capabilities for production the of machine parts using various advanced engineering materials, and at the same time have a few crucial limitations. Moreover, the AM processes of steel constitute one of the relevant research trends [26].
Referring to current trends both in terms of novel material solutions and advanced technological processes, the scope of the current literature overview is focused on the laser powder bed fusion (LPBF) technology concerning AHSS. The LBPF is considered one of the most versatile additive manufacturing processes. It is used to produce metallic parts with complex geometries in a relatively short time. The increasing use of this method is associated with advantages such as the possibility of obtaining defect-free parts with good mechanical properties, probably due to the microstructure obtained during the process. Also, AM processes may constitute the solution for the limited formability and processability that occurred during processing in conventional manufacturing routes. Moreover, the use of, among others, cellular structures will enable further reduction of the machine parts, which is one of the advantages of AM techniques considering lightweight design. Despite these advantages, the LBPF process still has limitations due to the limited knowledge of the process and the influence of the parameters [27]. Moreover, the additive manufacturing trade beyond adjusting process parameters also includes, among other factors, the quality of the powder supplied to the process [28], which affects the repeatability of the process and the quality of produced parts. It should also be emphasized that the mechanical properties of the part of AHSSs are associated with the thermomechanical processes [29,30,31,32] that are not present during LPBF fabrication. This is also a significant limitation.
The LBPF process, also known as selective laser melting (SLM), is based on the use of a laser beam to melt particles of metal powder. A high-energy laser beam melts a specific contour layer by layer until a three-dimensional element is obtained. During the process, complex thermodynamic mechanisms take place, the energy absorbed by the powder particles generates a high temperature within them, and the nature of the process affects the high cooling rate. The interaction between the laser and the powder can be associated with problems such as balling, pore formation, and thermal cracking [27,33].
Based on a critical insight into the published investigations, future directions and prospects related to LBPF and AHSS processes were formulated. Moreover, the most crucial limitations and challenges were identified. The results of this literature analysis may constitute a basis for the design of future LPBF processes for wider grades of AHSSs and to optimize the process parameters aimed at controlling the microstructure and mechanical/in-use properties.

2. Insight into the Steels Manufactured by Laser Powder Bed Fusion

Laser powder bed fusion as one of the most common additive manufacturing processes for metal production has been successfully used to produce different types of steel. With the use of this technology, it has been reported to produce components with materials such as conventional steels, maraging steels, and AHSSs [34].
However, it has to be addressed that components produced with LPBF will be characterized by different microstructures than in conventional processes [26]. The focus of the current review is directed towards LPBF processes and AHSSs, and it may be noticed that this group of steels constitutes a relatively novel scope of materials considering addictive manufacturing methods.

2.1. Short Overview of the Steels Commonly Produced by LPBF Not Covering the AHSS Group

2.1.1. Austenitic Stainless Steels (ASS)

Austenitic stainless steel produced by LPBF has been found to have excellent mechanical properties with an even better strength/ductility ratio, in comparison to conventionally produced components. The strength increase is sometimes attributed to microstructure obtained in the process, more specifically to the cellular or cellular–dendritic solidification structure. This specific microstructure is directly connected to the temperature gradients and rapid solidification rates inherent in the process [35,36]. AISI 316L is one of the most commonly used austenitic stainless steels in additive manufacturing, valued for its favorable mechanical properties and excellent corrosion resistance. The corrosion resistance of 3D-printed 316L is generally regarded as superior to that of forged material. It has been suggested that the cellular sub-structure formed during the laser powder bed fusion process enhances corrosion resistance. This is attributed to the segregation of molybdenum (Mo) and chromium (Cr) at the cell walls, creating a more corrosion-resistant barrier that helps limit the depth of corrosion attack [37]. The current state of additive manufacturing perspectives for ASS is formulated in an in-depth review developed by Haghdadi et al. [38].

2.1.2. Duplex Stainless Steels (DSS)

Duplex stainless steels consist of approximately 50% ferrite and 50% austenite and combine beneficial corrosion resistance with mechanical properties and twice the strength of traditional austenitic stainless steels [39,40]. However, DSSs constitute the demanding material for the LBPF process as they still require post-processing such as solution annealing to control the austenite and ferrite content and morphology. The main problem with those steels produced by the LPBF process is that they demonstrate mainly ferritic microstructure, which is attributed to very high cooling rates. A high cooling rate inherent to LPBF processing favors the formation of a dominantly ferritic and N loss during the process [41]. To overcome this problem, dedicated heat treatment is used, and research indicates that the duplex microstructure may be successfully recovered. One of the commonly used grades of DSSs is 2205 with the composition of Fe-22Cr-5Ni-3Mo-1Mn (wt.%). For 2205 steel produced by LBPF, the solution annealing at 1100 °C for 1 h in an air atmosphere followed by water quenching is typically applied to control the ferrite and austenite content [40]. The current state of additive manufacturing perspectives for DSSs is demonstrated in an in-depth review developed by Zhang et al. [42].

2.1.3. Oxide Dispersion Strengthened (ODS) Steels

Oxide Dispersion Strengthened steels represent a groundbreaking class of alloys, designed for industries demanding exceptional high-temperature resistance and mechanical strength. These materials are typically built on a ferritic/martensitic steel matrix, enriched with high levels of chromium (9–20%) to ensure corrosion resistance. Reinforcement comes from nanometer-sized dispersoids, which consist of titanium-containing yttrium-based oxides. In the nuclear industry, ODS ferritic alloys are particularly valuable as a potential alternative for radiation-resistant cladding and structural materials. Fine oxides, with their small size and high density, create numerous composite interfaces that help to eliminate point defects, preventing defect clustering and failure. This distinctive microstructure enhances their creep resistance and durability, making ODS steels the material of choice for applications in power plants, aerospace, and nuclear reactors [43,44]. The well-known way to fabricate ODS steel is mechanical alloying of Y2O3 which requires high-energy mechanical milling. Another way to produce ODS steel is the use of gas atomization reaction synthesis powders (GARS). The GARS-produced steel powders consist of a base ferritic steel matrix. As a result of the rapid solidification during atomization, stable Y-containing intermetallic (typically Y2Fe17) are formed along interdendritic regions or with Y that is retained in solution or at grain boundaries [45]. The current state of additive manufacturing perspectives for ODS steel is presented in an in-depth review developed by Wilms et al. [46].

2.1.4. Precipitation-Hardened (PH) Stainless Steels

Precipitation-hardened steels are a class of high-strength stainless steels that achieve superior mechanical properties through the precipitation-hardening process. This process involves heating the steel to precipitate fine intermetallic compounds, which strengthen the material. PH steels offer an excellent combination of strength, corrosion resistance, and toughness, making them ideal for demanding applications in the aerospace, automotive, and chemical industries. They are commonly used in components like shafts, gears, and fasteners, where both high strength and resistance to harsh environments are essential. One of the most commonly used PH steel is 17-4 PH, this grade of steel is fully martensitic when it is fabricated through conventional methods such as forging, casting, or welding. PH steels are also commonly fabricated using AM [47,48,49,50]. However, when produced with the LBPF method, which involves extremely high-temperature gradients and rapid cooling rates, the microstructure is in a non-equilibrium state with metastable austenitic and BCC ferritic phases. PH steels produced by LPBF can be characterized by a wide range of phase compositions. Unlike conventionally manufactured PH 17-4 SS, LPBF-printed PH 17-4 SS can retain metastable austenite at room temperature [51]. It is found that 17-4 PH exhibits a complex, inhomogeneous microstructure with various amounts of δ ferrite, austenite, and martensite in the as-built condition. Research shows that even small variations of the chemical composition of used powders have a significant effect on the phase composition in the as-built samples [52]. The current state of additive manufacturing perspectives for PH stainless steel is included in reviews developed by Ko et al. [53] and Zadi-Maad et al. [54].

2.2. Advanced High-Strength Steels Produced by LPBF

2.2.1. Medium- and High-Mn TWIP and TRIP Steels

Overall, the TWIP effect in steel occurs when the mechanical twins are formed during plastic deformation. When the steel is subjected to stress, instead of dislocation movement, twinning occurs, accommodating the deformation. TWIP steels typically have a medium or high Mn content. Medium-Mn steels refer to the 3–12 wt.% of Mn [55], while high-Mn steels are characterized by manganese content exceeding 13 wt.% [56]. The presence of martensitic transformation of retained austenite may be also the result of transformation-induced plasticity. Moreover, TWIP and TRIP steels benefit from the so-called “dynamic Hall–Petch effect”, a deformation-induced grain refinement leading to the evolution of new obstacles to dislocation motion [57]. TWIP steels are known for their excellent combination of high strength and ductility. This makes them perspective material for applications where the machine parts are subjected to absorbing large amounts of energy, such as in automotive crash components. Thus, TRIP steels also offer a beneficial balance between strength and ductility. The transformation to martensite absorbs energy, improving ductility while increasing strength.
The mechanical properties depend on the chemical composition of the steel and the volume fraction of retained austenite in its microstructure. The activation of the TRIP and TWIP deformation mechanisms is determined by factors such as volume fraction of structure constituents, grain size, and stacking fault energy (SFE) of the austenite [58], as well as the strain rate [59]. The current challenges of medium- and high-Mn steels are associated with the understanding of the physical metallurgy phenomena, twinning mechanisms, deformation mechanisms, texture evolution, fracture mechanisms, hydrogen embrittlement behavior, and fatigue performance [8,60,61].
The high content of manganese may lead to difficulties with recrystallization processes, which typically leads to a depletion of manganese and may occur during high-temperature processes such as SLM. However, Niendorf et al. [62], first reported that the effective manufacturing of TWIP steels is achieved using the SLM technique. Steels produced with this process are characterized by mechanical properties similar to the conventionally processed material. Despite an anisotropic microstructure featuring elongated grains, preferred grain orientation, and process-induced imperfections such as inclusions and pores, the SLM-processed TWIP steel demonstrates high strength and high ductility even in a not-post-treated condition. The effective manufacturing of high-Mn TRIP and TWIP steels was also reported in other works [63,64,65,66,67]. The scheme of solidification processes during AM processes of high-Mn steels is presented in Figure 1. Haase et al. [63] investigated the high-Mn steel characterized by reduced chemical element segregation compared to conventionally manufactured alloys. The limited segregation constitutes a prospect to avoiding the additional post-treatment. Due to very fast cooling rates, the structure consisted of an austenite matrix with ε- and α’-martensite. Both, TRIP and TWIP effects were activated after LPBF processes. Authors suggest that the challenge related to LPBF processes of high-Mn steels is related to reduced formability. Ewald et al. [65] investigated the influence of Al content on the deformation mechanisms of LPBF-processed high-Mn steel. After the LPBF processes the TRIP and TWIP effects were observed, which contribution was related to the Al content. Moreover, Kohnen et al. [67] observed the tendency that increasing the Al content, a transition from austenite (FCC) to ferrite–austenite (BCC-FCC) occurs during solidification for 4 wt. % of Al. As a result, strong grain refinement and texture randomization were achieved. The presence of strong grain refinement and suppression of the contribution of the TRIP effect in the dual phase high-Mn steels (above 4 wt.%) plays a crucial role in enhancing the energy absorption capacity and ductile deformation behavior. Kies et al. [64] noticed that high-Mn steels subjected to LPBF processes are characterized by reduced segregation and suppressed grain boundary embrittlement. The authors proposed the fabrication of a lattice structure intended for energy-absorbing applications. Both TRIP and TWIP effects were revealed. Overall, high-Mn austenitic TRIP/TWIP steels are characterized by sufficient processability and demonstrate comparable mechanical properties compared to the hot-rolling processes without application of additional post-treatment. The overview of performed investigations is presented in Table 1.
Pawlak et al. [68] and Heemann [69] reported investigations involving low-carbon, medium-Mn steel. In medium-Mn steels, a martensitic or martensitic–bainitic matrix with a controlled content of retained austenite typically occurs. In as-built conditions, the content of retained austenite is lower compared to conventional manufacturing routes. The post-treatment (intercritical annealing) significantly enhances the austenite susceptibility to the TRIP effect and increases the fraction of retained austenite compared to the as-built condition (Figure 2). It should be also noted that both papers applied preheating treatment at 200 °C to avoid cold cracking, and high relative densities were achieved (99.8–99.9%). The conducted research indicates that post-treatment (intercritical annealing) is necessary to enhance mechanical behavior up to the level of conventional manufacturing processes of medium-Mn steels.

2.2.2. Maraging Steels

Maraging steels constitute a group of AHSSs characterized by ultra-high mechanical properties caused mainly by the precipitation strengthening of intermetallic compounds (among others Ni3Ti, Ni3Mo, Ni3 (Ti,Mo), Fe2Mo Laves phase, or B2-ordered NiAl) that are semi-coherent or coherent or with the low-carbon martensitic matrix [15,70,71]. Overall, maraging steel contains a low content of carbon (up to 0.03 wt.%) and a significant content of nickel (17–19 wt.%), together with lesser contents of cobalt (8–12%), molybdenum (3–5 wt.%), titanium (0.2–1.8 wt.%), and aluminum (0.1–0.15 wt.%). The alloying elements contribute to the formation of intermetallic phases during aging. The addition of Cr to the chemical composition results in the design of corrosion-resistant maraging steels [72]. The microstructure of maraging steels consists of a ductile martensitic matrix reinforced by the precipitation of nanometric intermetallic phases. The name “maraging” comes from the used heat treatment. These steels are subjected to heat treatment—hardening and then aging at a temperature of 450–510 °C—which leads to a significant increase in hardness and strength properties as a result of intermetallic strengthening. The current challenges associated with the development and industrialization of maraging steel include stress corrosion cracking of welded joints, fatigue life, and deformation behaviors [73,74,75].
Maraging steels are commonly manufactured using additive manufacturing, especially LPBF processes. In recent years, a few in-depth review papers have been published on the current developments in AM and manufacturing of maraging steels [76,77,78,79]. For this reason, the scope of this literature review presents the most crucial assumptions and findings regarding maraging steel and LPBF processes focused on papers published during the last few years. It is stated that the AM methods and additional heat treatment (aging processes) enable the manufacturing of the maraging steels characterized by comparable strength properties and hardness properties compared to the conventionally fabricated parts [76,80]. The differences in the ductility and strength properties of maraging steels are related to the presence of retained and reversed austenite caused by the chemical inhomogeneity related to the micro-segregation upon solidification [81] (visible also in Figure 3). It is also revealed that after AM processes the anisotropy of mechanical properties occurs related to the build orientation [77]. For the final structure and quality of LPBF-fabricated maraging steels, the process parameters play a crucial role (laser power, hatch distance, laser scanning velocity, build direction, layer thickness). The influence of process parameters on the structure, defects, and properties of maraging steels is presented in a review developed by B.S Rao and T.B. Rao [79]. One of the advantages of producing maraging steels by AM processes such as LBPF is the intrinsic rapid solidification rate during the process, which may facilitate the austenite to martensitic phase transformation and reduce the post-heat treatment time compared to the conventional manufacturing process [81].
Based on recent findings, a summary is presented regarding the fabrication of maraging steel using LPBF/SLM processes (also in Table 2). It is reported that one of the most commonly used maraging steels in LPBF processes is the M300 grade. This grade is characterized by the capability to produce parts without cracks and minimum porosity, and after postprocessing may demonstrate an expected strength and fracture toughness combination. Wei et al. [83] revealed that approximately 9.2% of retained austenite in the as-build sample is preserved. Takata et al. [82] noticed that fine retained austenite with a <001> orientation along the build direction is mainly formed at the melt pool boundaries. After complete heat treatment (quenching + aging; and solutionizing + austenitizing) it is possible to obtain fully martensitic microstructures [83,84] Moreover, during the direct aging process, austenite reversion occurs and may lead to an increase in the overall austenite fraction [83,85].
Król et al. [86] revealed the presence of two solid-state reactions during the heating of SLM-manufactured maraging samples:
  • precipitation of intermetallic phases;
  • reversion of martensite into austenite.
Based on the Kissinger equation, the authors indicated the activation energy for each reaction (precipitation of intermetallic phases 301 kJ mol−1, the martensite to austenite reversion 478 kJ mol−1). Moreover, Eres-Castellanos et al. [87] revealed that during reheating related to the subsequent deposition, grains subjected to re-austenitization keep the same crystallographic orientation as the surrounding retained austenite. Santana et al. [88] demonstrated that no crucial influence of the used printing parameters on the microstructure evolution occurred during aging. Furthermore, it was noticed that the hardness of aged maraging steel is comparable to that obtained by conventional manufacturing routes. Kannan et al. [89] observed the direct aging processes for Ti-free maraging steel manufactured using LPBF for various aging times. The authors indicated that aging at a lower temperature (or shorter time) led to reduced strain hardening caused by a lower content of reverted austenite. On the other hand, aging at a higher temperature led to extensive recrystallization of martensite, precipitation coarsening, and significant austenite reversion, which caused the softening of the fabricated samples. Piekło et al. [90] developed the relation between aging temperature and fatigue life. From the perspective of corrosion resistance, Zhao et al. [91] demonstrated that SLM-fabricated 18Ni300 maraging steel was negatively correlated to the size of pores formed during printing. A.G. Demir and B. Previtali [92] attempted to use the preheating of the 18Ni300 maraging steel (at 170 °C), aiming at reducing the porosity. However, the preheating was not significantly effective in terms of enhancing the density.
Table 2. Overview of selected maraging steels subjected to LPBF processes.
Table 2. Overview of selected maraging steels subjected to LPBF processes.
PowderParametersCommentsRef.
X3NiCoMoTi18-9-5 steel (in wt %)Gas atomized, particle size: 15 to 45 μmEnergy density: 60–70 J/mm3,
Laser powder: 300 W,
Scanning speed:
1 cm/s,
Layer thickness:
50 μm,
Argon gas atmosphere
In as-built state: elongated grains with nano-scale cellular dendritic structure. This microstructure transformed into a martensitic lath microstructure after heat treatment.
The retained austenite completely transformed into martensite after solution annealing heat treatment.
The aging process produced reverted austenite.
[85]
M789
(0.02C-12.2Cr-10Ni-1Mo0.06Al-1Ti-Fe)
Gas atomized,
Particle size: 15–45 μm
Energy density:
75–85 J/mm3
Argon gas atmosphere
In as-built structures: melt pool boundaries with nano-scale columnar dendritic structures were present. The elongated grains transformed into martensitic needle-like structures after annealing and aging treatments.[93]
18Ni-350
(18.44Ni-11.91Co-4.88Mo-1.45Ti-0.12Al-0.01C)
Gas-atomization under nitrogen environment,
average particle size: 30 μm
Laser power: 175 W,
Hatch spacing:
100 μm,
Scan speed:
300 mm/s,
Layer thickness:
30 μm,
Scan rotation:
67°,
Energy density: 194.4 J/mm3,
Relative density: 99.6–99.8 %
After direct aging, austenite reversion occurred, increasing the overall austenite fraction (21.8%).
Austenitizing–aging, and solutionizing–austenitizing–aging treatments eliminated the austenite and a fully martensitic microstructure was achieved.
A high density of dislocations was found in the martensitic laths in all the conditions examined.
[83]
18Ni-300
0.025Mn-0.01Cr-4.5Mo-18Ni-0.07Al-9.2Co-0.91Ti
Atomized in an inert argon medium using electrode induction melting gas atomization, average particle size: 40 umLaser power: 300 W,
Focus diameter: 70 μm,
Scan speed: 1300 mm/s,
Hatch space:
70 μm,
Layer thickness: 30 μm,
Rotation angle: 33°
Absence of nano-precipitate in the as-built state, while precipitations of Ni-Ti-based and Mo-based precipitates were confirmed in both the aged and solution-aged states.[94]
18Ni-300Gas atomized,
particle size: 15 to 53 μm
Laser power: 285 W,
Scanning speed: 960 mm/s,
Hatching space: 110 μm,
Layer thickness:
40 μm,
Scanning strategy: Orthogonal, Relative density: >99.5%
Significant anisotropies in the thermal expansion and CTE were observed between the OX (parallel) direction and OZ (perpendicular) direction in the LPBFed samples.
The solution-treated and solution+ageing-treated samples showed differences in the coefficient of thermal expansion CTE compared to direct aging-treated samples.
[95]
18Ni-300 Gas atomized, average particle size: 88.8 μmLaser power:
2000 W,
Scanning speed: 1000–2000 mm/s, Hatch distance: 0.06–0.1 mm,
Layer thickness: 0.1–0.2 mm,
Hatch angle:
90°,
Energy density:
50–180 J/mm3,
Build rate:
6–40 mm3/s
High power (2000 W) contributed to typical HP-LPBF dendritic microstructure inside columnar prior-austenite grains (PAGs) growing along the building direction.
The original cellular and columnar dendrites were replaced by parallel lath martensite and nanoprecipitates could be observed in the sample with solution-aging treatment.
[96]
Based on the literature overview, the following summary was formulated:
  • Maraging steels constitute a steel group that may be effectively fabricated using LPBF/SLM processes after adjusting the process parameters to achieve high-density samples and reduced fraction of defects;
  • The mechanical properties of LPBF-fabricated maraging steels are generally comparable to the conventional manufacturing routes;
  • The LPBF-fabricated maraging steels to achieve expected mechanical properties are subjected to heat treatment (quenching + aging; solutionizing+austenitizing) or direct aging;
  • During aging, the intermetallic precipitation growth is similar to the conventional manufacturing routes;
  • During aging, the martensite is prone to transformation into reverted austenite;
  • In the as-built conditions, the retained austenite is preserved. The retained austenite is formed especially at the melting pool boundaries due to occurred micro-segregation of local chemical composition.

2.2.3. Bainitic Steels

Over three decades, ultra-fine bainitic steels have attracted scientists and the industry sector due to their beneficial combination of mechanical properties in comparison to conventional martensitic steels [97,98,99]. Obtaining a bainitic structure with ultra-fine or even nanoscale dimension is related to the solid-state reaction at low temperatures and through dedicated alloy design high silicon steels [14]. The structure of ultra-fine bainitic steels contained carbon-super saturated bainitic ferrite and stable retained austenite [100]. The higher content Si (typically above 1.5 wt.%) leads to incomplete transformation and retained austenite stabilization at room temperature [101]. A similar role to Si is associated with Al considering the suppression of cementite precipitation from austenite [102]. The bainitic steels with increased Al and Si content due to limited carbide precipitations are called carbide-free [103]. The current challenges for the development of advanced bainitic steels cover, among other, the strategy for reducing carbon content, fatigue and in-service performance, the thermal stability of the metastable structure, deformation behavior, and the thermomechanical processes [104,105,106,107,108].
Despite extensive research covering the development of advanced bainitic steels, the current challenge to overcome is connected to extending industrial applications [17]. AM processes in the manufacturing of ultra-fine bainitic steels open a new window related to their potential applications. The majority of the literature [109,110,111,112,113] indicate that the most common AM process for producing bainitic structures is Wire and Arc Additive Manufacturing (WAAM).
Prior to publication of this paper, only a few published works have attempted to produce bainitic steels using LPBF processes (Table 3). Bartels et al. [114] investigated low-carbon bainitic steel subjected to LPBF processes. The minor content of retained austenite was detected in as-built conditions, while post-treatment led to the reduction of their content. The bainite morphology was different in the fusion zone and heat-affected zone (upper and lower bainite with cementite precipitations). Franceschi et al. [115] investigated carbide-free bainitic steel subjected to the LPBF process. The samples after LPBF were characterized by a refined carbide-free bainitic structure due to the formation of smaller prior austenite grains in comparison to conventional manufacturing processes. Moreover, the samples subjected to LPBF demonstrated a faster bainitic transformation kinetic caused by the finer prior austenite grain size. The inhomogeneity of structure constituents was noticed within the single layer following the layer-by-layer manufacturing process. After the deposition on a novel layer, the austenitized layer will transform into martensite, and further layers will be subjected to tempering processes. The post-treatment involved a complete heat treatment route for ultra-fine bainitic steels. However, the structure after post-treatment was characterized by various bainite morphologies and bimodal distribution of bainitic ferrite plate thickness in the fusion zone, core, and heat-affected zone (Figure 4).
The lack of literature related to the production of LPBF ultra fine bainitic steels is directly related to its limitations:
  • The ultra-fine bainitic steels are characterized by medium or high carbon content. The bainitic structure is achieved during prolonged isothermal heat treatment. Thus, during fast cooling after the process, the cold cracks related to martensite transformation are prone to be formed, similar to welding processes [116]. The solution for avoiding the cold crack is the usage of heated support with the temperature above martensite start (Ms) temperature.
  • The LPBF is connected with formation layer-by-layer. The metastable bainitic structure together with the limited thermal stability of retained austenite tend to the decomposition process [106] and the deposition of subsequent layers. This process is similar to the tempering process.
  • It seems that LPBF-fabricated bainitic steel should be subjected to a complete post-treatment process similar to conventional heat treatment (quenching and isothermal heat treatment). However, the processing of bainitic steels with lower carbon content and faster bainitic transformation may be produced using heated support at the temperature of bainitic transformation during the whole LPBF process. However, due to the segregation of chemical elements, the distribution of retained austenite may be not homogeneous in the fusion zone and heat-affected zone [115].
  • The future direction of investigations of AM processes of bainitic steels is also associated with the fabrication capabilities of secondary hardened bainitic steels [117,118,119], continuous cooling bainitic steels [120,121,122], and intermetallic strengthened bainitic steels [123,124,125].
The experimental studies of developing the optimized LPBF technology for ultra-fine bainitic steels constitute a research gap and require further research aimed at controlling the structure morphology and economical aspects of manufacturing. It cannot be ruled out that there is a necessity to develop a new design alloy strategy oriented towards AM processes to overcome the most crucial difficulties.

2.2.4. Novel Concepts of Advanced Steels—Low-Density High-Strength Steels

The necessity for reducing the weight of steel for structural purposes has been growing, especially considering the transportation and automotive sectors [126]. Thus, a novel class of high-strength steels called low-density or lightweight steels was developed [127]. The current challenges for low-density steels include optimization of heat treatment, reducing the cost of alloys, and formability [128]. The promising grade of low-density steels is TRIPLEX steels (Fe-Al-Mn-C alloy system) characterized by a microstructure consisting of an austenite matrix and finely dispersed nanoscale κ-carbides [129,130,131]. Overall, the industrial applications of high-Mn TRIPLEX steels are limited due to the inherent difficulties related to conventional manufacturing processes, among others: (i) the necessity of performing energy-intensive post-treatment caused by strong segregation [132]; (ii) limited machinability [133]; (iii) reduced high-temperature formability [134]; and (iv) generation of reactive slag [135]. It seems that these challenges may be solved by using powder additive manufacturing processes aimed at directly fabricating components considering minor post-processing. Seede et al. [136] were the first to introduce the LPBF to produce steel with reduced density (Fe-30Mn-9Al-1Si-0.5Mo-9C wt.%). Promising tensile properties were achieved (ultimate tensile strength up to 1.3 GPa, while elongation was approximately 36% considering the build direction). However, after the LPBF processes, the presence of solidification micro-cracks in the whole cross-sectional area related to the hot cracking was detected. The microstructure constituted of austenite with cellular solidification morphology. Moreover, the strong micro-segregation of Si and Mn was revealed in the cellular–dendritic structures, as well as macro-segregation at the melt pool boundaries. On the other hand, Sanchez-Poncela et al. [137] were the first to manufacture TRIPLEX steel using LPBF technology. Based on the performed investigations, the hot cracking HC is prone to be formed when the alloy composition is not tailored. The hot cracking occurs as vertical cracks along the as-building direction and further may propagate across the whole cross-section of produced parts (Figure 5). The authors suggest that the presence of hot cracks in TRIPLEX steels may be predicted using hot cracking coefficient (HCS) criteria determined based on the solidification curve determined by the CALPHAD method. Particular attention was paid to the P and S, which segregate into the liquid during solidification by forming thin regions enriched with these chemical elements, which constitute a potential risk for hot cracking. Despite that austenite is the matrix of TRIPLEX steels, the ferrite formation at the dendrite boundaries was noticed for alloys characterized by higher Si content. Based on performed tests, the authors stated that AM processes constitute a reliable alternative to conventional metallurgical processes aimed at achieving thin, small, or complex shape components produced from low-density TRIPLEX steels.
However, the scientific field of low-density steels and AM processes still constitutes a large research gap and requires further investigations focused on eliminating cracks, micro- and macro-segregation, controlling the morphology of the microstructure and carbide precipitation, and tailoring the structure–property relationship. A crucial role is also played by process parameters and the development of effective post-treatment processes aimed at enhancing the mechanical properties and enabling the producing larger parts.

3. Conclusions

Based on the literature overview, a review of AHSSs produced by additive manufacturing methods was performed, focusing on laser powder bed fusion/selective laser melting processes. Although maraging steels are a common group of steels produced by AM techniques, other AHSSs still constitute a large research field where there are more questions than answers. AM techniques open a new window of perspectives towards the industrialization of steel grades with excellent mechanical and in-use properties, which are promising for reducing the weight of structures and designing steels with low environmental impact. The most relevant findings from the literature review include:
  • Maraging steels: Adjusting process parameters enabled the fabrication of the parts with high density and a limited fraction of defects. After heat treatment or direct aging, strengthening by intermetallic phases is achieved similarly to conventional manufacturing processes. After the LBPF processes, the structure retains austenite located mainly at the melting pool boundaries, which is related to the micro-segregation of the chemical composition. During aging, the transformation of the martensite into reverted austenite occurs.
  • High-Mn steels with TWIP/TRIP effect: Mechanical properties after LPBF are comparable to the conventional processing. The fully austenitic matrix is achieved. Additional post-printing heat treatment is not necessary to introduce TWIP and TRIP effects. The design strategy to reduce the segregation of chemical elements is favorable for AM processes. Both ε- and α´-martensite are formed during deformation. The contribution of TRIP and TWIP effects is related to the Al content.
  • Medium-Mn steels with TRIP effect: A martensitic or martensitic–bainitic matrix with a controlled content of retained austenite is formed during the LPBF. Martensite is also subjected to the tempering process during printing. In as-built conditions, the content of retained austenite is lower compared to conventional manufacturing routes. The post-process heat treatment (intercritical annealing) significantly enhances the austenite susceptibility to the TRIP effect and increases the fraction of retained austenite compared to the as-built condition. Preheating at 200 °C is used to avoid cold cracking. The conducted research indicates that post-process heat treatment (intercritical annealing) is necessary to enhance mechanical behavior up to the level of conventional manufacturing processes of medium-Mn steels.
  • Bainitic steels: They still constitute a research gap. After the LPBF process, a complete heat treatment allows for the formation of the expected carbide-free bainitic microstructure, which is, however, still characterized by heterogeneity. In terms of bainitic steels, a new chemical composition design strategy is required to achieve a bainitic structure during printing characterized by a combination of strength and ductility. Process parameters also require further investigation to prevent cold and hot cracking.
  • Low-Density, High-Strength Steels: Although the low-density still attracts huge attention of researchers, the AM processes constitute a large research gap and require further investigations focused on eliminating cracks, micro- and macro-segregation, controlling the morphology of the microstructure and carbide precipitation (especially Kappa carbide), and tailoring the structure–property relationship. A crucial role is also played by process parameters and the development of effective post-treatment processes aimed at enhancing the mechanical properties and enabling the producing larger parts. Only TRIPLEX steel and Fe-30Mn-9Al-1Si-0.5Mo-9C were subjected to LPBF processes, which is a promising premise for further research.
Based on the analysis of the published papers on the LPBF/SLM technique and AHSS fabrication, this issue constitutes a significant research challenge. Extensive research into the manufacturing of maraging steel confirms the effectiveness of LPBF techniques in achieving mechanical properties comparable to conventional manufacturing routes. However, several phenomena in this area require further explanation and understanding (e.g., chemical composition segregation, the relationship between the retained austenite reversion and intermetallic formation, and the relationship between the texture and process parameters). When it comes to high- and medium-Mn, bainitic, or high-strength, low-density steels, there are still more questions than answers. It is difficult to state unequivocally whether LPBF techniques will be truly effective in fabricating these steel grades, which requires further experimental research. However, research on high-Mn and medium-Mn steels is prospective in these terms. On the other hand, bainitic steels and high-strength, low-density steels constitute a research gap, as confirmed by the first published works (published in 2024: TRIPLEX steel [137] and carbide-free bainitic steel [115]). Therefore, AHSSs represent an open window for future research and constitute currently a “hot” research topic.

Author Contributions

Conceptualization, A.K. and J.M., validation, A.K., formal analysis, J.M.; resources, A.K.; data curation, J.M., and A.K.; writing—original draft preparation, J.M., and A.K.; writing—review and editing, A.K.; visualization, J.M.; supervision, A.K.; project administration, A.K.; funding acquisition, A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Centre (Opus Project, No. 2023/51/B/ST8/01393, title: “Intermetallic Strengthening in nanocrystalline bainitic steels considering the additive manufacturing method—INSTaBAIN”.

Acknowledgments

Supported by the Foundation for Polish Science (FNP)—Program START.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Raabe, D.; Sun, B.; Kwiatkowski Da Silva, A.; Gault, B.; Yen, H.-W.; Sedighiani, K.; Thoudden Sukumar, P.; Souza Filho, I.R.; Katnagallu, S.; Jägle, E.; et al. Current Challenges and Opportunities in Microstructure-Related Properties of Advanced High-Strength Steels. Metall. Mater. Trans. A 2020, 51, 5517–5586. [Google Scholar] [CrossRef]
  2. Tasan, C.C.; Diehl, M.; Yan, D.; Bechtold, M.; Roters, F.; Schemmann, L.; Zheng, C.; Peranio, N.; Ponge, D.; Koyama, M.; et al. An Overview of Dual-Phase Steels: Advances in Microstructure-Oriented Processing and Micromechanically Guided Design. Annu. Rev. Mater. Res. 2015, 45, 391–431. [Google Scholar] [CrossRef]
  3. Kalhor, A.; Soleimani, M.; Mirzadeh, H.; Uthaisangsuk, V. A review of recent progress in mechanical and corrosion properties of dual phase steels. Arch. Civ. Mech. Eng. 2020, 20, 85. [Google Scholar] [CrossRef]
  4. Heibel, S.; Dettinger, T.; Nester, W.; Clausmeyer, T.; Tekkaya, A.E. Damage Mechanisms and Mechanical Properties of High-Strength Multiphase Steels. Materials 2018, 11, 761. [Google Scholar] [CrossRef]
  5. Grajcar, A.; Skowronek, A.; Radwański, K. Mechanical behavior and stability of dispersed retained austenite in thermomechanically rolled and isothermally-treated TRIP-aided multiphase steel. Mater. Sci. Eng. A 2022, 830, 142300. [Google Scholar] [CrossRef]
  6. Hu, B.; Luo, H.; Yang, F.; Dong, H. Recent progress in medium-Mn steels made with new designing strategies, a review. J. Mater. Sci. Technol. 2017, 33, 1457–1464. [Google Scholar] [CrossRef]
  7. Grajcar, A.; Kilarski, A.; Kozlowska, A. Microstructure–Property Relationships in Thermomechanically Processed Medium-Mn Steels with High Al Content. Metals 2018, 8, 929. [Google Scholar] [CrossRef]
  8. Jacob, R.; Raman Sankaranarayanan, S.; Kumaresh Babu, S.P. Recent advancements in manganese steels—A review. Mater. Today Proc. 2020, 27, 2852–2858. [Google Scholar] [CrossRef]
  9. Grajcar, A.; Borek, W. Thermo-mechanical processing of high-manganese austenitic TWIP-type steels. Arch. Civ. Mech. Eng. 2008, 8, 29–38. [Google Scholar] [CrossRef]
  10. De Cooman, B.C.; Estrin, Y.; Kim, S.K. Twinning-induced plasticity (TWIP) steels. Acta Mater. 2018, 142, 283–362. [Google Scholar] [CrossRef]
  11. Soleimani, M.; Kalhor, A.; Mirzadeh, H. Transformation-induced plasticity (TRIP) in advanced steels: A review. Mater. Sci. Eng. A 2020, 795, 140023. [Google Scholar] [CrossRef]
  12. Caballero, F.G.; Bhadeshia, H.K.D.H.; Mawella, K.J.A.; Jones, D.G.; Brown, P. Design of novel high strength bainitic steels: Part 1. Mater. Sci. Technol. 2001, 17, 512–516. [Google Scholar] [CrossRef]
  13. Caballero, F.G.; Bhadeshia, H.K.D.H.; Mawella, K.J.A.; Jones, D.G.; Brown, P. Design of novel high strength bainitic steels: Part 2. Mater. Sci. Technol. 2001, 17, 517–522. [Google Scholar] [CrossRef]
  14. Caballero, F.G.; Garcia-Mateo, C.; Miller, M.K. Design of novel bainitic steels: Moving from ultrafine to nanoscale structures. Jom 2014, 66, 747–755. [Google Scholar] [CrossRef]
  15. da Fonseca, D.P.M.; Feitosa, A.L.M.; de Carvalho, L.G.; Plaut, R.L.; Padilha, A.F. A Short Review on Ultra-High-Strength Maraging Steels and Future Perspectives. Mater. Res. 2021, 24, e20200470. [Google Scholar] [CrossRef]
  16. Rana, R.; Cordova-Tapia, E.; Jimenez, J.A.; Morales-Rivas, L.; Garcia-Mateo, C. Design of carbide free bainitic steels for hot rolling practices. Philos. Mag. Lett. 2024, 104, 2322552. [Google Scholar] [CrossRef]
  17. Morales-Rivas, L. Viewpoints on Technological Aspects of Advanced High-Strength Bainitic Steels. Metals 2022, 12, 195. [Google Scholar] [CrossRef]
  18. Garcia-Mateo, C.; Sourmail, T.; Caballero, F.G.; Smanio, V.; Kuntz, M.; Ziegler, C.; Leiro, A.; Vuorinen, E.; Elvira, R.; Teeri, T. Nanostructured steel industrialisation: Plausible reality. Mater. Sci. Technol. 2014, 30, 1071–1078. [Google Scholar] [CrossRef]
  19. Sun, L.; Simm, T.H.; Martin, T.L.; McAdam, S.; Galvin, D.R.; Perkins, K.M.; Bagot, P.A.J.; Moody, M.P.; Ooi, S.W.; Hill, P.; et al. A novel ultra-high strength maraging steel with balanced ductility and creep resistance achieved by nanoscale β-NiAl and Laves phase precipitates. Acta Mater. 2018, 149, 285–301. [Google Scholar] [CrossRef]
  20. Li, J.; Zhan, D.; Jiang, Z.; Zhang, H.; Yang, Y.; Zhang, Y. Progress on improving strength-toughness of ultra-high strength martensitic steels for aerospace applications: A review. J. Mater. Res. Technol. 2023, 23, 172–190. [Google Scholar] [CrossRef]
  21. Pereira, R.; Peixinho, N.; Costa, S.L. A Review of Sheet Metal Forming Evaluation of Advanced High-Strength Steels (AHSS). Metals 2024, 14, 394. [Google Scholar] [CrossRef]
  22. Kvackaj, T.; Bidulská, J.; Bidulský, R. Overview of HSS Steel Grades Development and Study of Reheating Condition Effects on Austenite Grain Size Changes. Materials 2021, 14, 1988. [Google Scholar] [CrossRef] [PubMed]
  23. Grajcar, A.; Kuziak, R.; Zalecki, W. Third generation of AHSS with increased fraction of retained austenite for the automotive industry. Arch. Civ. Mech. Eng. 2012, 12, 334–341. [Google Scholar] [CrossRef]
  24. Bouaziz, O.; Zurob, H.; Huang, M. Driving Force and Logic of Development of Advanced High Strength Steels for Automotive Applications. Steel Res. Int. 2013, 84, 937–947. [Google Scholar] [CrossRef]
  25. Lesch, C.; Kwiaton, N.; Klose, F.B. Advanced High Strength Steels (AHSS) for Automotive Applications − Tailored Properties by Smart Microstructural Adjustments. Steel Res. Int. 2017, 88, 1700210. [Google Scholar] [CrossRef]
  26. Bajaj, P.; Hariharan, A.; Kini, A.; Kürnsteiner, P.; Raabe, D.; Jägle, E.A. Steels in additive manufacturing: A review of their microstructure and properties. Mater. Sci. Eng. A 2020, 772, 138633. [Google Scholar] [CrossRef]
  27. Narasimharaju, S.R.; Zeng, W.; See, T.L.; Zhu, Z.; Scott, P.; Jiang, X.; Lou, S. A comprehensive review on laser powder bed fusion of steels: Processing, microstructure, defects and control methods, mechanical properties, current challenges and future trends. J. Manuf. Process. 2022, 75, 375–414. [Google Scholar] [CrossRef]
  28. Gruber, K.; Smolina, I.; Stopyra, W. Assessing metal powder quality for additive manufacturing using diffuse light spectroscopy. Powder Technol. 2024, 434, 119366. [Google Scholar] [CrossRef]
  29. Romano-Acosta, L.F.; García-Rincon, O.; Pedraza, J.P.; Palmiere, E.J. Influence of thermomechanical processing parameters on critical temperatures to develop an Advanced High-Strength Steel microstructure. J. Mater. Sci. 2021, 56, 18710–18721. [Google Scholar] [CrossRef]
  30. Krbata, M.; Krizan, D.; Eckert, M.; Kaar, S.; Dubec, A.; Ciger, R. Austenite Decomposition of a Lean Medium Mn Steel Suitable for Quenching and Partitioning Process: Comparison of CCT and DCCT Diagram and Their Microstructural Changes. Materials 2022, 15, 1753. [Google Scholar] [CrossRef]
  31. Krbaťa, M.; Barényi, I.; Eckert, M.; Križan, D.; Kaar, S.; Breznická, A. Hot deformation analysis of lean medium-manganese 0.2C3Mn1.5Si steel suitable for quenching et partitioning process. Met. Mater. 2021, 59, 379–390. [Google Scholar] [CrossRef]
  32. Zhao, J.; Jiang, Z. Thermomechanical processing of advanced high strength steels. Prog. Mater. Sci. 2018, 94, 174–242. [Google Scholar] [CrossRef]
  33. Liu, Y.; Yang, Y.; Mai, S.; Wang, D.; Song, C. Investigation into spatter behavior during selective laser melting of AISI 316L stainless steel powder. Mater. Des. 2015, 87, 797–806. [Google Scholar] [CrossRef]
  34. Köhnen, P.; Létang, M.; Voshage, M.; Schleifenbaum, J.H.; Haase, C. Understanding the process-microstructure correlations for tailoring the mechanical properties of L-PBF produced austenitic advanced high strength steel. Addit. Manuf. 2019, 30, 100914. [Google Scholar] [CrossRef]
  35. Funch, C.V.; Somlo, K.; Christiansen, T.L.; Somers, M.A.J. Thermochemical post-processing of additively manufactured austenitic stainless steel. Surf. Coatings Technol. 2022, 441, 128495. [Google Scholar] [CrossRef]
  36. 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]
  37. Dong, P.; Vecchiato, F.; Yang, Z.; Hooper, P.A.; Wenman, M.R. The effect of build direction and heat treatment on atmospheric stress corrosion cracking of laser powder bed fusion 316L austenitic stainless steel. Addit. Manuf. 2021, 40, 101902. [Google Scholar] [CrossRef]
  38. Haghdadi, N.; Laleh, M.; Moyle, M.; Primig, S. Additive manufacturing of steels: A review of achievements and challenges. J. Mater. Sci. 2021, 56, 64–107. [Google Scholar] [CrossRef]
  39. Köhler, M.L.; Kunz, J.; Herzog, S.; Kaletsch, A.; Broeckmann, C. Microstructure analysis of novel LPBF-processed duplex stainless steels correlated to their mechanical and corrosion properties. Mater. Sci. Eng. A 2021, 801, 140432. [Google Scholar] [CrossRef]
  40. Nigon, G.N.; Burkan Isgor, O.; Pasebani, S. The effect of annealing on the selective laser melting of 2205 duplex stainless steel: Microstructure, grain orientation, and manufacturing challenges. Opt. Laser Technol. 2021, 134, 106643. [Google Scholar] [CrossRef]
  41. Haghdadi, N.; Laleh, M.; Chen, H.; Chen, Z.; Ledermueller, C.; Liao, X.; Ringer, S.; Primig, S. On the pitting corrosion of 2205 duplex stainless steel produced by laser powder bed fusion additive manufacturing in the as-built and post-processed conditions. Mater. Des. 2021, 212, 110260. [Google Scholar] [CrossRef]
  42. Zhang, D.; Liu, A.; Yin, B.; Wen, P. Additive manufacturing of duplex stainless steels—A critical review. J. Manuf. Process. 2022, 73, 496–517. [Google Scholar] [CrossRef]
  43. Paul, B.K.; Lee, K.; He, Y.; Ghayoor, M.; Chang, C.; Pasebani, S. Oxide dispersion strengthened 304 L stainless steel produced by ink jetting and laser powder bed fusion. CIRP Ann. 2020, 69, 193–196. [Google Scholar] [CrossRef]
  44. Wilms, M.B.; Streubel, R.; Frömel, F.; Weisheit, A.; Tenkamp, J.; Walther, F.; Barcikowski, S.; Schleifenbaum, J.H.; Gökce, B. Laser additive manufacturing of oxide dispersion strengthened steels using laser-generated nanoparticle-metal composite powders. Procedia CIRP 2018, 74, 196–200. [Google Scholar] [CrossRef]
  45. Horn, T.; Rock, C.; Kaoumi, D.; Anderson, I.; White, E.; Prost, T.; Rieken, J.; Saptarshi, S.; Schoell, R.; DeJong, M.; et al. Laser powder bed fusion additive manufacturing of oxide dispersion strengthened steel using gas atomized reaction synthesis powder. Mater. Des. 2022, 216, 110574. [Google Scholar] [CrossRef]
  46. Wilms, M.B.; Rittinghaus, S.-K.; Goßling, M.; Gökce, B. Additive manufacturing of oxide-dispersion strengthened alloys: Materials, synthesis and manufacturing. Prog. Mater. Sci. 2023, 133, 101049. [Google Scholar] [CrossRef]
  47. Sabooni, S.; Chabok, A.; Feng, S.C.; Blaauw, H.; Pijper, T.C.; Yang, H.J.; Pei, Y.T. Laser powder bed fusion of 17–4 PH stainless steel: A comparative study on the effect of heat treatment on the microstructure evolution and mechanical properties. Addit. Manuf. 2021, 46, 102176. [Google Scholar] [CrossRef]
  48. Sabzi, H.E.; Lim, S.-H.; Della Crociata, D.; Castellote-Alvarez, R.; Simonelli, M.; San-Martín, D.; Hao, X.; Choi, P.-P.; Rivera-Díaz-del-Castillo, P.E.J. Genetic design of precipitation-hardening stainless steels for additive manufacturing. Acta Mater. 2024, 274, 120018. [Google Scholar] [CrossRef]
  49. Roberts, D.; Zhang, Y.; Charit, I.; Zhang, J. A comparative study of microstructure and high-temperature mechanical properties of 15-5 PH stainless steel processed via additive manufacturing and traditional manufacturing. Prog. Addit. Manuf. 2018, 3, 183–190. [Google Scholar] [CrossRef]
  50. Fields, B.; Amiri, M.; Lim, J.; Pürstl, J.T.; Begley, M.R.; Apelian, D.; Valdevit, L. Microstructural Control of a Multi-Phase PH Steel Printed with Laser Powder Bed Fusion. Adv. Mater. Technol. 2024, 9, 2301037. [Google Scholar] [CrossRef]
  51. Singh, K.K.; Moridi, A. Metastability mediated grain size control in PH 17-4 stainless steel fabricated using Laser-Powder Bed Fusion (LPBF). Materialia 2023, 30, 101857. [Google Scholar] [CrossRef]
  52. Vunnam, S.; Saboo, A.; Sudbrack, C.; Starr, T.L. Effect of powder chemical composition on the as-built microstructure of 17-4 PH stainless steel processed by selective laser melting. Addit. Manuf. 2019, 30, 100876. [Google Scholar] [CrossRef]
  53. Ko, G.; Kim, W.; Kwon, K.; Lee, T.-K. The Corrosion of Stainless Steel Made by Additive Manufacturing: A Review. Metals 2021, 11, 516. [Google Scholar] [CrossRef]
  54. Zadi-Maad, A.; Rohib, R.; Irawan, A. Additive manufacturing for steels: A review. IOP Conf. Ser. Mater. Sci. Eng. 2018, 285, 012028. [Google Scholar] [CrossRef]
  55. Sun, B.; Kwiatkowski da Silva, A.; Wu, Y.; Ma, Y.; Chen, H.; Scott, C.; Ponge, D.; Raabe, D. Physical metallurgy of medium-Mn advanced high-strength steels. Int. Mater. Rev. 2023, 68, 786–824. [Google Scholar] [CrossRef]
  56. Miller, R.L. Ultrafine-grained microstructures and mechanical properties of alloy steels. Metall. Trans. 1972, 3, 905–912. [Google Scholar] [CrossRef]
  57. Gutierrez-Urrutia, I.; Raabe, D. Grain size effect on strain hardening in twinning-induced plasticity steels. Scr. Mater. 2012, 66, 992–996. [Google Scholar] [CrossRef]
  58. Zambrano, O.A. A general perspective of Fe–Mn–Al–C steels. J. Mater. Sci. 2018, 53, 14003–14062. [Google Scholar] [CrossRef]
  59. Wang, M.; Huang, M.X. Abnormal TRIP effect on the work hardening behavior of a quenching and partitioning steel at high strain rate. Acta Mater. 2020, 188, 551–559. [Google Scholar] [CrossRef]
  60. Kim, S.; Jeong, D.; Sung, H. Reviews on factors affecting fatigue behavior of high-Mn steels. Met. Mater. Int. 2018, 24, 1–14. [Google Scholar] [CrossRef]
  61. Koyama, M.; Akiyama, E.; Lee, Y.-K.; Raabe, D.; Tsuzaki, K. Overview of hydrogen embrittlement in high-Mn steels. Int. J. Hydrogen Energy 2017, 42, 12706–12723. [Google Scholar] [CrossRef]
  62. Niendorf, T.; Brenne, F. Steel showing twinning-induced plasticity processed by selective laser melting—An additively manufactured high performance material. Mater. Charact. 2013, 85, 57–63. [Google Scholar] [CrossRef]
  63. Haase, C.; Bültmann, J.; Hof, J.; Ziegler, S.; Bremen, S.; Hinke, C.; Schwedt, A.; Prahl, U.; Bleck, W. Exploiting Process-Related Advantages of Selective Laser Melting for the Production of High-Manganese Steel. Materials 2017, 10, 56. [Google Scholar] [CrossRef] [PubMed]
  64. Kies, F.; Köhnen, P.; Wilms, M.B.; Brasche, F.; Pradeep, K.G.; Schwedt, A.; Richter, S.; Weisheit, A.; Schleifenbaum, J.H.; Haase, C. Design of high-manganese steels for additive manufacturing applications with energy-absorption functionality. Mater. Des. 2018, 160, 1250–1264. [Google Scholar] [CrossRef]
  65. Ewald, S.; Schaukellis, M.; Koehnen, P.; Schleifenbaum, J.H. Laser Powder Bed Fusion of Advanced High-Strength Steels—Modification of Deformation Mechanisms by Increasing Stacking Fault Energy. BHM Berg Hüttenmännische Monatshefte 2019, 164, 127–132. [Google Scholar] [CrossRef]
  66. Niendorf, T.; Brenne, F.; Hoyer, P.; Schwarze, D.; Schaper, M.; Grothe, R.; Wiesener, M.; Grundmeier, G.; Maier, H.J. Processing of New Materials by Additive Manufacturing: Iron-Based Alloys Containing Silver for Biomedical Applications. Metall. Mater. Trans. A 2015, 46, 2829–2833. [Google Scholar] [CrossRef]
  67. Köhnen, P.; Ewald, S.; Schleifenbaum, J.H.; Belyakov, A.; Haase, C. Controlling microstructure and mechanical properties of additively manufactured high-strength steels by tailored solidification. Addit. Manuf. 2020, 35, 101389. [Google Scholar] [CrossRef]
  68. Pawlak, A.; Dziedzic, R.; Kasprowicz, M.; Stopyra, W.; Kuźnicka, B.; Chlebus, E.; Schob, B.; Zopp, C.; Kroll, L.; Kordass, R.; et al. Properties of medium-manganese steel processed by laser powder bed fusion: The effect of microstructure in as-built and intercritically annealed state on energy absorption during tensile and impact tests. Mater. Sci. Eng. A 2023, 870, 144859. [Google Scholar] [CrossRef]
  69. Heemann, L.; Mostaghimi, F.; Schob, B.; Schubert, F.; Kroll, L.; Uhlenwinkel, V.; Steinbacher, M.; Toenjes, A.; von Hehl, A. Adjustment of Mechanical Properties of Medium Manganese Steel Produced by Laser Powder Bed Fusion with a Subsequent Heat Treatment. Materials 2021, 14, 3081. [Google Scholar] [CrossRef]
  70. Wang, W.; Yan, W.; Duan, Q.; Shan, Y.; Zhang, Z.; Yang, K. Study on fatigue property of a new 2.8GPa grade maraging steel. Mater. Sci. Eng. A 2010, 527, 3057–3063. [Google Scholar] [CrossRef]
  71. Galindo-Nava, E.I.; Rainforth, W.M.; Rivera-Díaz-del-Castillo, P.E.J. Predicting microstructure and strength of maraging steels: Elemental optimisation. Acta Mater. 2016, 117, 270–285. [Google Scholar] [CrossRef]
  72. Hadadzadeh, A.; Shahriari, A.; Amirkhiz, B.S.; Li, J.; Mohammadi, M. Additive manufacturing of an Fe–Cr–Ni–Al maraging stainless steel: Microstructure evolution, heat treatment, and strengthening mechanisms. Mater. Sci. Eng. A 2020, 787, 139470. [Google Scholar] [CrossRef]
  73. Bagchi, A.; Gope, D.K.; Chattopadhyaya, S.; Wuriti, G. A critical review on susceptibility of stress corrosion cracking in maraging steel weldments. Mater. Today Proc. 2020, 27, 2303–2307. [Google Scholar] [CrossRef]
  74. Rohit, B.; Muktinutalapati, N.R. Fatigue Behavior of 18% Ni Maraging Steels: A Review. J. Mater. Eng. Perform. 2021, 30, 2341–2354. [Google Scholar] [CrossRef]
  75. Nageswara Rao, M.; Narayana Murty, S.V.S. Hot Deformation of 18% Ni Maraging Steels: A Review. Mater. Perform. Charact. 2019, 8, 742–765. [Google Scholar] [CrossRef]
  76. Guo, L.; Zhang, L.; Andersson, J.; Ojo, O. Additive manufacturing of 18% nickel maraging steels: Defect, structure and mechanical properties: A review. J. Mater. Sci. Technol. 2022, 120, 227–252. [Google Scholar] [CrossRef]
  77. Mooney, B.; Kourousis, K. 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]
  78. Kizhakkinan, U.; Seetharaman, S.; Raghavan, N.; Rosen, D.W. Laser Powder Bed Fusion Additive Manufacturing of Maraging Steel: A Review. J. Manuf. Sci. Eng. 2023, 145, 110801. [Google Scholar] [CrossRef]
  79. Rao, B.S.; Rao, T.B. Effect of Process Parameters on Powder Bed Fusion Maraging Steel 300: A Review. Lasers Manuf. Mater. Process. 2022, 9, 338–375. [Google Scholar] [CrossRef]
  80. Simson, T.; Koch, J.; Rosenthal, J.; Kepka, M.; Zetek, M.; Zetková, I.; Wolf, G.; Tomčík, P.; Kulhánek, J. Mechanical Properties of 18Ni-300 maraging steel manufactured by LPBF. Procedia Struct. Integr. 2019, 17, 843–849. [Google Scholar] [CrossRef]
  81. Jägle, E.; Sheng, Z.; Kürnsteiner, P.; Ocylok, S.; Weisheit, A.; Raabe, D. Comparison of Maraging Steel Micro- and Nanostructure Produced Conventionally and by Laser Additive Manufacturing. Materials 2016, 10, 8. [Google Scholar] [CrossRef] [PubMed]
  82. Takata, N.; Nishida, R.; Suzuki, A.; Kobashi, M.; Kato, M. Crystallographic Features of Microstructure in Maraging Steel Fabricated by Selective Laser Melting. Metals 2018, 8, 440. [Google Scholar] [CrossRef]
  83. Wei, S.; Kumar, P.; Lau, K.B.; Wuu, D.; Liew, L.-L.; Wei, F.; Teo, S.L.; Cheong, A.; Ng, C.K.; Zhang, B.; et al. Effect of heat treatment on the microstructure and mechanical properties of 2.4 GPa grade maraging steel fabricated by laser powder bed fusion. Addit. Manuf. 2022, 59, 103190. [Google Scholar] [CrossRef]
  84. Karlapudy, S.P.; Nancharaiah, T.; Subba Rao, V.V. Influence of post-heat treatment on microstructure, texture, and mechanical properties of 18Ni-300 maraging steel fabricated by using LPBF technique. Prog. Addit. Manuf. 2023, 1–16. [Google Scholar] [CrossRef]
  85. Chadha, K.; Tian, Y.; Bocher, P.; Spray, J.G.; Aranas, C. Microstructure Evolution, Mechanical Properties and Deformation Behavior of an Additively Manufactured Maraging Steel. Materials 2020, 13, 2380. [Google Scholar] [CrossRef]
  86. Król, M.; Snopiński, P.; Czech, A. The phase transitions in selective laser-melted 18-NI (300-grade) maraging steel. J. Therm. Anal. Calorim. 2020, 142, 1011–1018. [Google Scholar] [CrossRef]
  87. Eres-Castellanos, A.; Santana, A.; De-Castro, D.; Jimenez, J.A.; Rementeria, R.; Capdevila, C.; Caballero, F.G. Effect of processing parameters on texture and variant selection of as-built 300 maraging steel processed by laser powder bed fusion. Sci. Rep. 2022, 12, 16168. [Google Scholar] [CrossRef]
  88. Santana, A.; Eres-Castellanos, A.; Jimenez, J.A.; Rementeria, R.; Capdevila, C.; Caballero, F.G. Effect of layer thickness and laser emission mode on the microstructure of an additive manufactured maraging steel. J. Mater. Res. Technol. 2023, 25, 6898–6912. [Google Scholar] [CrossRef]
  89. Kannan, R.; Leonard, D.N.; Nandwana, P. Optimization of direct aging temperature of Ti free grade 300 maraging steel manufactured using laser powder bed fusion (LPBF). Mater. Sci. Eng. A 2021, 817, 141266. [Google Scholar] [CrossRef]
  90. Piekło, J.; Garbacz-Klempka, A.; Myszka, D.; Figurski, K. Numerical and Experimental Analysis of Strength Loss of 1.2709 Maraging Steel Produced by Selective Laser Melting (SLM) under Thermo-Mechanical Fatigue Conditions. Materials 2023, 16, 7682. [Google Scholar] [CrossRef]
  91. Zhao, Z.; Dong, C.; Kong, D.; Wang, L.; Ni, X.; Zhang, L.; Wu, W.; Zhu, L.; Li, X. Influence of pore defects on the mechanical property and corrosion behavior of SLM 18Ni300 maraging steel. Mater. Charact. 2021, 182, 111514. [Google Scholar] [CrossRef]
  92. Demir, A.G.; Previtali, B. Investigation of remelting and preheating in SLM of 18Ni300 maraging steel as corrective and preventive measures for porosity reduction. Int. J. Adv. Manuf. Technol. 2017, 93, 2697–2709. [Google Scholar] [CrossRef]
  93. Palad, R.; Tian, Y.; Chadha, K.; Rodrigues, S.; Aranas, C. Microstructural features of novel corrosion-resistant maraging steel manufactured by laser powder bed fusion. Mater. Lett. 2020, 275, 128026. [Google Scholar] [CrossRef]
  94. Allam, T.; Pradeep, K.G.; Köhnen, P.; Marshal, A.; Schleifenbaum, J.H.; Haase, C. Tailoring the nanostructure of laser powder bed fusion additively manufactured maraging steel. Addit. Manuf. 2020, 36, 101561. [Google Scholar] [CrossRef]
  95. Bai, Y.; Zhao, C.; Zhang, J.; Wang, H. Abnormal thermal expansion behaviour and phase transition of laser powder bed fusion maraging steel with different thermal histories during continuous heating. Addit. Manuf. 2022, 53, 102712. [Google Scholar] [CrossRef]
  96. Huang, G.; Wei, K.; Deng, J.; Zeng, X. High power laser powder bed fusion of 18Ni300 maraging steel: Processing optimization, microstructure and mechanical properties. Mater. Sci. Eng. A 2022, 856, 143983. [Google Scholar] [CrossRef]
  97. Bhadeshia, H.K.D.H. Nanostructured bainite. Proc. R. Soc. A Math. Phys. Eng. Sci. 2010, 466, 3–18. [Google Scholar] [CrossRef]
  98. Caballero, F.G.; Bhadeshia, H.K.D.H. Very strong bainite. Curr. Opin. Solid State Mater. Sci. 2004, 8, 251–257. [Google Scholar] [CrossRef]
  99. Caballero, F.G.; Bhadeshia, H.K.D.H.; Mawella, K.J.A.; Jones, D.G.; Brown, P. Very strong low temperature bainite. Mater. Sci. Technol. 2002, 18, 279–284. [Google Scholar] [CrossRef]
  100. Królicka, A.; Caballero, F.G.; Jimenez, J.A.; Zalecki, W.; Kuziak, R. Development of advanced cementite-free bainitic steel using Ni, Al and Cu additions—Design concept and phase transformations study. J. Mater. Res. Technol. 2023, 27, 1265–1280. [Google Scholar] [CrossRef]
  101. Bhadeshia, H.K.D.H.; Edmonds, D.V. The bainite transformation in a silicon steel. Metall. Trans. A 1979, 10, 895–907. [Google Scholar] [CrossRef]
  102. Huang, H.; Sherif, M.Y.; Rivera-Díaz-del-Castillo, P.E.J. Combinatorial optimization of carbide-free bainitic nanostructures. Acta Mater. 2013, 61, 1639–1647. [Google Scholar] [CrossRef]
  103. Caballero, F.G. Carbide-free bainite in steels. In Phase Transformations in Steels; Elsevier: Amsterdam, The Netherlands, 2012; pp. 436–467. [Google Scholar]
  104. Garcia-Mateo, C.; Paul, G.; Somani, M.; Porter, D.; Bracke, L.; Latz, A.; Garcia De Andres, C.; Caballero, F. Transferring Nanoscale Bainite Concept to Lower C Contents: A Perspective. Metals 2017, 7, 159. [Google Scholar] [CrossRef]
  105. Zhang, F.; Yang, Z. Development of and Perspective on High-Performance Nanostructured Bainitic Bearing Steel. Engineering 2019, 5, 319–328. [Google Scholar] [CrossRef]
  106. Królicka, A.; Żak, A.M.; Caballero, F.G. Enhancing technological prospect of nanostructured bainitic steels by the control of thermal stability of austenite. Mater. Des. 2021, 211, 110143. [Google Scholar] [CrossRef]
  107. Benrabah, I.-E.; Deboer, A.; Geandier, G.; Van Landeghem, H.; Hutchinson, C.; Brechet, Y.; Zurob, H. Effect of deformation on the bainitic transformation. Acta Mater. 2024, 277, 120195. [Google Scholar] [CrossRef]
  108. Caballero, F.G.; Allain, S.; Cornide, J.; Puerta Velásquez, J.D.; Garcia-Mateo, C.; Miller, M.K. Design of cold rolled and continuous annealed carbide-free bainitic steels for automotive application. Mater. Des. 2013, 49, 667–680. [Google Scholar] [CrossRef]
  109. Youheng, F.; Guilan, W.; Haiou, Z.; Liye, L. Optimization of surface appearance for wire and arc additive manufacturing of Bainite steel. Int. J. Adv. Manuf. Technol. 2017, 91, 301–313. [Google Scholar] [CrossRef]
  110. Mu, S.; Li, Y.; Song, D.; Xu, B.; Chen, X. Low Cycle Fatigue Behavior and Failure Mechanism of Wire Arc Additive Manufacturing 16MND5 Bainitic Steel. J. Mater. Eng. Perform. 2021, 30, 4911–4924. [Google Scholar] [CrossRef]
  111. Xing, X.; Qin, G.; Zhou, Y.; Yu, H.; Liu, L.; Zhang, L.; Yang, Q. Microstructure Optimization and Cracking Control of Additive Manufactured Bainite Steel by Gas Metal Arc Welding Technology. J. Mater. Eng. Perform. 2019, 28, 5138–5145. [Google Scholar] [CrossRef]
  112. Fu, Y.; Zhang, H.; Wang, G.; Wang, H. Investigation of mechanical properties for hybrid deposition and micro-rolling of bainite steel. J. Mater. Process. Technol. 2017, 250, 220–227. [Google Scholar] [CrossRef]
  113. Panicker, S.; Nagarajan, H.P.N.; Tuominen, J.; Patnamsetty, M.; Coatanéa, E.; Haapala, K.R. Investigation of thermal influence on weld microstructure and mechanical properties in wire and arc additive manufacturing of steels. Mater. Sci. Eng. A 2022, 853, 143690. [Google Scholar] [CrossRef]
  114. Bartels, D.; Novotny, T.; Mohr, A.; van Soest, F.; Hentschel, O.; Merklein, C.; Schmidt, M. PBF-LB/M of Low-Alloyed Steels: Bainite-like Microstructures despite High Cooling Rates. Materials 2022, 15, 6171. [Google Scholar] [CrossRef]
  115. Franceschi, M.; Yazdanpanah, A.; Leone, D.; Pezzato, L.; Dabalà, M. Laser Powder Bed Fusion Fabrication of a Novel Carbide-Free Bainitic Steel: The Possibilities and a Comparative Study with the Conventional Alloy. Metals 2024, 14, 113. [Google Scholar] [CrossRef]
  116. Królicka, A.; Ambroziak, A.; Żak, A. Welding Capabilities of Nanostructured Carbide-Free Bainite: Review of Welding Methods, Materials, Problems, and Perspectives. Appl. Sci. 2019, 9, 3798. [Google Scholar] [CrossRef]
  117. Kuntz, M.; Wicks, G.; Sourmail, T.; Garcia-Mateo, C. Fatigue Properties at Elevated Temperatures of Secondary Tempered Nanostructured Bainite: A Study of Deformation Mechanisms. Steel Res. Int. 2024, 95, 2400105. [Google Scholar] [CrossRef]
  118. Królicka, A.; Caballero, F.G.; Zalecki, W.; Kuziak, R.; Rozmus, R. Controlling the Thermal Stability of a Bainitic Structure by Alloy Design and Isothermal Heat Treatment. Materials 2023, 16, 2963. [Google Scholar] [CrossRef]
  119. Huang, B.M.; Yang, J.R.; Yen, H.W.; Hsu, C.H.; Huang, C.Y.; Mohrbacher, H. Secondary hardened bainite. Mater. Sci. Technol. 2014, 30, 1014–1023. [Google Scholar] [CrossRef]
  120. Shahriari, B.; Vafaei, R.; Sharifi, E.M.; Farmanesh, K. Continuous cooling transformation behavior and the kinetics of bainite formation in a bainitic–martensitic steel. Int. J. Mater. Res. 2017, 108, 715–724. [Google Scholar] [CrossRef]
  121. Królicka, A.; Lesiuk, G.; Kuziak, R.; Radwański, K.; Janik, A. The role of microstructure morphology on fracture mechanisms of continuously cooled bainitic steel designed for rails application. Metall. Mater. Trans. A 2023, 54, 487–504. [Google Scholar] [CrossRef]
  122. Zheng, Y.-F.; Wu, R.-M.; Li, X.-C.; Wu, X.-C. Continuous cooling transformation behaviour and bainite formation kinetics of new bainitic steel. Mater. Sci. Technol. 2017, 33, 454–463. [Google Scholar] [CrossRef]
  123. Hulme-Smith, C.N.; Ooi, S.W.; Bhadeshia, H.K.D.H. Intermetallic-strengthened nanocrystalline bainitic steel. Mater. Sci. Technol. 2018, 34, 1976–1979. [Google Scholar] [CrossRef]
  124. Królicka, A.; Żak, A.M.; Caballero, F.G. Revealing intermetallic strengthening in bainitic steel designed for nickel aluminide formation. Scr. Mater. 2024, 249, 116153. [Google Scholar] [CrossRef]
  125. Królicka, A.; Caballero, F.G.; Kuziak, R.; Radwański, K.; Sozańska-Jędrasik, L.; Stawarczyk, P. Two design strategies for enhancing the thermal stability of bainitic structures. J. Mater. Res. Technol. 2023, 24, 7004–7020. [Google Scholar] [CrossRef]
  126. Zhang, W.; Xu, J. Advanced lightweight materials for Automobiles: A review. Mater. Des. 2022, 221, 110994. [Google Scholar] [CrossRef]
  127. Chen, S.; Rana, R.; Haldar, A.; Ray, R.K. Current state of Fe-Mn-Al-C low density steels. Prog. Mater. Sci. 2017, 89, 345–391. [Google Scholar] [CrossRef]
  128. Hu, S.; Zheng, Z.; Yang, W.; Yang, H. Fe–Mn–C–Al Low-Density Steel for Structural Materials: A Review of Alloying, Heat Treatment, Microstructure, and Mechanical Properties. Steel Res. Int. 2022, 93, 2200191. [Google Scholar] [CrossRef]
  129. Kim, H.; Suh, D.-W.; Kim, N.J. Fe–Al–Mn–C lightweight structural alloys: A review on the microstructures and mechanical properties. Sci. Technol. Adv. Mater. 2013, 14, 014205. [Google Scholar] [CrossRef] [PubMed]
  130. Sutou, Y.; Kamiya, N.; Umino, R.; Ohnuma, I.; Ishida, K. High-strength Fe-20Mn-Al-C-based Alloys with Low Density. ISIJ Int. 2010, 50, 893–899. [Google Scholar] [CrossRef]
  131. Gutierrez-Urrutia, I.; Raabe, D. Influence of Al content and precipitation state on the mechanical behavior of austenitic high-Mn low-density steels. Scr. Mater. 2013, 68, 343–347. [Google Scholar] [CrossRef]
  132. Daamen, M.; Haase, C.; Dierdorf, J.; Molodov, D.A.; Hirt, G. Twin-roll strip casting: A competitive alternative for the production of high-manganese steels with advanced mechanical properties. Mater. Sci. Eng. A 2015, 627, 72–81. [Google Scholar] [CrossRef]
  133. Akhlaghi, S.; Yue, S. Effect of Thermomechanical Processing on the Hot Ductility of a Nb-Ti Microalloyed Steel. ISIJ Int. 2001, 41, 1350–1356. [Google Scholar] [CrossRef]
  134. Steenken, B.; Rezende, J.L.L.; Senk, D. Hot ductility behaviour of high manganese steels with varying aluminium contents. Mater. Sci. Technol. 2017, 33, 567–573. [Google Scholar] [CrossRef]
  135. Wietbrock, B.; Bambach, M.; Seuren, S.; Hirt, G. Homogenization Strategy and Material Characterization of High-Manganese TRIP and TWIP Steels. Mater. Sci. Forum 2010, 638–642, 3134–3139. [Google Scholar] [CrossRef]
  136. Seede, R.; Whitt, A.; Ye, J.; Gibbons, S.; Flater, P.; Gaskey, B.; Elwany, A.; Arroyave, R.; Karaman, I. A lightweight Fe–Mn–Al–C austenitic steel with ultra-high strength and ductility fabricated via laser powder bed fusion. Mater. Sci. Eng. A 2023, 874, 145007. [Google Scholar] [CrossRef]
  137. Sanchez-Poncela, M.; Jin, Q.; Hebert, R.; Aindow, M.; Amigó-Borrás, V.; Rementeria, R. Triplex steel powder design to avoid hot cracking in laser-powder bed fusion using computational thermodynamics. Addit. Manuf. 2024, 86, 104196. [Google Scholar] [CrossRef]
Figure 1. Microstructure evolution in AM manufactured high-Mn steel solidification. (a) The formation of columnar and equiaxed dendrites occurred during the deposition of a laser-melted track and heat-affected zone (HAZ). ∇T and CET indicate temperature gradient and columnar morphology into equiaxed morphology transition. (b) Origin of the texture caused by the solidification related to the 25° inclination of sustaining grains connected with the remelting process. (c) Texture originated from the route [64].
Figure 1. Microstructure evolution in AM manufactured high-Mn steel solidification. (a) The formation of columnar and equiaxed dendrites occurred during the deposition of a laser-melted track and heat-affected zone (HAZ). ∇T and CET indicate temperature gradient and columnar morphology into equiaxed morphology transition. (b) Origin of the texture caused by the solidification related to the 25° inclination of sustaining grains connected with the remelting process. (c) Texture originated from the route [64].
Materials 17 04699 g001
Figure 2. Phase distribution maps of medium-Mn steel manufactured using LPBF. Comparison of retained austenite content of as-built conditions and intercritical annealed processes at 690 °C and 750 °C. Area of the fracture surface indicates susceptibility to TRIP effect of retained austenite [69].
Figure 2. Phase distribution maps of medium-Mn steel manufactured using LPBF. Comparison of retained austenite content of as-built conditions and intercritical annealed processes at 690 °C and 750 °C. Area of the fracture surface indicates susceptibility to TRIP effect of retained austenite [69].
Materials 17 04699 g002
Figure 3. (ac) Scheme of microstructure formation of maraging steel during SLM/LPBF processes. (d) Thermal history of the sample related to local laser heating [82]. γ—austenite; G.B.—grain boundary; α—ferrite; Ms—martensite start temperature; As—solidification temperature.
Figure 3. (ac) Scheme of microstructure formation of maraging steel during SLM/LPBF processes. (d) Thermal history of the sample related to local laser heating [82]. γ—austenite; G.B.—grain boundary; α—ferrite; Ms—martensite start temperature; As—solidification temperature.
Materials 17 04699 g003
Figure 4. (ac) Bainitic microstructure at prior melt pool boundary. γf—austenite films, γb—austenite blocks, αb—bainitic ferrite [115].
Figure 4. (ac) Bainitic microstructure at prior melt pool boundary. γf—austenite films, γb—austenite blocks, αb—bainitic ferrite [115].
Materials 17 04699 g004
Figure 5. (af) Overview of hot cracking presence in six various TRIPLEX grades (Triplex, Triplex-HP, Triplex-HP-2Si, Triplex-1Si, Triplex-2Si, Triplex-4Si) produced by LPBF process with parameters: 200 W, 700 mm/s, and VED of 130 J/mm3 [137].
Figure 5. (af) Overview of hot cracking presence in six various TRIPLEX grades (Triplex, Triplex-HP, Triplex-HP-2Si, Triplex-1Si, Triplex-2Si, Triplex-4Si) produced by LPBF process with parameters: 200 W, 700 mm/s, and VED of 130 J/mm3 [137].
Materials 17 04699 g005
Table 1. Overview of medium- and high-Mn steels subjected to LPBF processes.
Table 1. Overview of medium- and high-Mn steels subjected to LPBF processes.
PowderParametersCommentsRef.
Fe-0.15C-0.46Si-7.21Mn-1.81Mn (wt.%)
Medium-Mn steel
Argon atomization, particle size: 20–65 μmLayer thickness:
30–60 μm,
Laser power:
250–350 W,
Scanning speed:
700–1000 mm/s,
Linear energy:
0.25–0.50 J/mm,
Preheatment:
200 °C,
Relative density:
99.76–99.87%
Intercritical annealing at 670 °C (0.5–6 h). The as-built microstructure: lath martensite and retained austenite (~8%). Microstructure after intercritical annealing: tempered martensite and retained austenite (~40%). Increased strength properties, elongation, and impact toughness. TRIP effect.[68]
Fe-0.23C-3.93Mn-2.01Al-0.51Si (wt.%)
Medium-Mn steel
Atomized using a Close-Coupled-Atomizer, particle size:
20–63 µm
Layer thickness:
60 μm,
Laser power:
300 W,
Scanning speed:
500 mm/s,
Preheatment:
200 °C,
Energy input:
111.1 J/mm3,
Relative density:
99.9%
Intercritical annealing at 630–770 °C (3, 10 and 60 min). The as-built microstructure: bainitic–martensitic matrix and retained austenite. Microstructure after intercritical annealing: martensite/bainite, M/A constituents, and retained austenite. After post-treatment mechanical behavior was comparable to conventional production routes. TRIP effect.[69]
Fe-0.60C-22.36Mn-0.25V-0.2Cr-0.25Si (wt.%)
High-Mn steel
Spray aeration in an argon inert gas, particle size: 40 μmLaser power:
Up to 400 W
An anisotropic structure covering elongated grains, texture, and process-related defects (inclusions and pores). Mechanical properties after LPBF are comparable to the conventional processing. Dynamic Hall–Petch effect was revealed. TWIP effect.[62]
Fe-0.27C-20.15Mn-0.05Si-0.01Al (wt.%)
High-Mn steel
Electrode induction melting gas atomization with argon, particle size: below 45 μmScan line spacing: 100 μm,
Scanning speed:
571 mm/s,
Relative density:
99.9%
Due to the reduced chemical elements segregation compared to strip-cast material the additional post-treatment may be avoided. Revealed crystallographic texture related to the solidification affected the mechanical behavior. The austenite matrix contained ε- and α′-martensite. TRIP and TWIP effect.[63]
Fe-0.33C-21.9Mn-0.01Al-0.03Si (wt.%)
High-Mn steel
Gas-atomized, particle size: 15–50 μmLayer thickness:
30 μm,
Laser power:
120 W,
Scanning speed:
700 mm/s
SLM lattice structure manufacturing. Reduced segregation after LPBF processes. TRIP and TWIP effects. Formation of ε- and α′-martensite during deformation.[64]
Fe-(21-22)Mn-(0-5)Al-0.3C (wt.%)
High-Mn steel
-Layer thickness:
30 μm,
Laser power:
90–120 W,
Scanning speed:
200–700 mm/s,
Relative density:
99.85–99.95%
TRIP and TWIP effects were revealed after the LPBF process. The Al content influenced the contribution of TRIP and TWIP effects.[65]
X30Mn21 steel and controlled content of aluminum powder (0.03–5.39 wt.%)
High-Mn steel
Argon gas atomized, particle size:
10–45 μm
Layer thickness:
30 μm,
Laser power:
120 W,
Scanning speed:
750 mm/s,
Relative density:
99.9%
TRIP and TWIP effects were revealed after the LPBF process. Fully austenite matrix below 4 wt.% of Al. Above 4 wt.% of Al the dual phase structure occurs (ferrite and austenite).[67]
Table 3. Overview of bainitic steels subjected to LPBF processes.
Table 3. Overview of bainitic steels subjected to LPBF processes.
PowderParametersCommentsRef.
Fe-0.22C-0.7Si-1.2Mn-1.0Cr-0.9Mo-(wt.%)
Low-carbon bainite
Particle size: 15–45 μmLayer thickness:
60 μm,
Laser power:
225–275 W,
Scanning speed:
550–850 mm/s,
Relative density:
above 99.7%
Post-treatment: tempering (150–600 °C), quenching, and tempering. Various bainite morphologies in the fusion zone and heat-affected zone. Post-treatment reduced the retained austenite.[114]
Fe-0.4C-3.2Si-2.6Mn-0.1Al
Medium-carbon carbide-free bainite
Argon atomization particle size:
15–53 µm
Layer thickness:
60 μm,
Laser power:
170–295 W,
Scanning speed:
500–750 mm/s,
Energy density:
22–89 J/mm3,
Pore density:
from 2.15%
Post-treatment: austenitization at 900 °C and isothermal annealing at 325 °C for 3 h. A refined carbide-free bainitic structure due to the smaller prior austenite grains formed during the LPBF process. Non-homogeneous bainite morphology. The cracks were observed.[115]
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

Królicka, A.; Malawska, J. Recent Progress in Laser Powder Bed Fusions Processes of Advanced High-Strength Steels. Materials 2024, 17, 4699. https://doi.org/10.3390/ma17194699

AMA Style

Królicka A, Malawska J. Recent Progress in Laser Powder Bed Fusions Processes of Advanced High-Strength Steels. Materials. 2024; 17(19):4699. https://doi.org/10.3390/ma17194699

Chicago/Turabian Style

Królicka, Aleksandra, and Julia Malawska. 2024. "Recent Progress in Laser Powder Bed Fusions Processes of Advanced High-Strength Steels" Materials 17, no. 19: 4699. https://doi.org/10.3390/ma17194699

APA Style

Królicka, A., & Malawska, J. (2024). Recent Progress in Laser Powder Bed Fusions Processes of Advanced High-Strength Steels. Materials, 17(19), 4699. https://doi.org/10.3390/ma17194699

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