Open-access Influence of Solution Annealing Temperature on Crystallographic Texture and Grain Disorientation of Nb-Enhanced Maraging Steel: An EBSD Analysis

Abstract

This study investigates the microstructural and crystallographic changes in a novel Nb-enhanced, Ti-reduced maraging steel. The hot-rolled steel was solution annealed at two different temperatures followed by aging. The Electron Backscatter Diffraction (EBSD) analysis of hot-rolled sample revealed a predominant {112}//RP texture, reoriented to {110}//RP during aging to minimize internal energy and stress. Solution annealing at 820oC followed by aging favored a {111} orientation, resulting in minimal crystallographic defects and grain distortion. In contrast, higher solution annealing temperatures promoted the formation of {001} cleavage planes, increasing brittleness and crystal defects, thereby impacting the material's suitability for high-performance applications.

Keywords:
Martensite lath; crystal orientation; grain distortion; crystal defects


1. Introduction

Maraging steels, a blend of martensitic and aging properties, are renowned for their exceptional strength and toughness while retaining ductility1-3. These ultra-high-strength steels, with very low carbon content, owe their remarkable properties to the precipitation of intermetallic compounds rather than carbon. Primarily composed of 15-25 wt% nickel, these steels also include secondary elements such as cobalt (Co), molybdenum (Mo), and titanium (Ti), which form intermetallic precipitates that enhance strength1. Developed in the late 1950s, maraging steels have evolved to include Co-free variants due to economic considerations. Common non-stainless grades typically contain 17-19 wt% Ni, 8-12 wt% Co, 3-5 wt% Mo, and 0.2-1.6 wt% Ti3. The addition of chromium produces stainless grades, improving corrosion resistance and indirectly increasing hardenability. The importance of maraging steels in engineering lies in their exceptional strength (≥1000 MPa) and good ductility (15-20%), making them critical for applications in aerospace and other demanding fields2. The strengthening mechanism of maraging steels involves forming a martensitic structure in the FeNi matrix. Upon aging, fine, uniformly dispersed Ni-rich precipitates, such as Ni3(Ti, Mo) form within the martensitic matrix1. These precipitates inhibit dislocation movement, significantly enhancing the material's strength without compromising toughness. This process, known as precipitation hardening, allows maraging steels to maintain high strength and thermal stability even under repeated heating and cooling cycles, making them ideal for high-performance engineering applications.

Niobium in maraging steels has impact on mechanical properties and microstructural stability. Research by Parvinian et al.1 demonstrated that microalloying 18% Ni maraging steels with small amounts of Nb and B delayed precipitate coarsening and maintained peak aged strength over extended periods. Lee et al.4 found that adding Nb improved both strength and fatigue life, with a notable increase in yield strength and elongation. Nb additions also promoted transformation-induced plastic deformation and martensite hardening, as shown by Li et al.2, who observed refined grains and the formation of nanometer-scaled precipitates that enhanced both strength and plasticity. Li et al.5 confirmed that multi-component alloying, including Nb, led to Ni3M precipitate formation, significantly boosting tensile strength and corrosion resistance. Sha et al.6 highlighted different age-hardening behaviors with Nb microalloying, where Nb-rich Laves phases formed, enhancing the mechanical properties. Hossein-Nejad et al.7 reported that Nb additions mitigated intergranular brittleness by hindering planar slip and increasing grain boundary fracture stress. Lastly, Ahmed's work8 ⁠on recycling 18Ni(350) maraging steel with Nb additions showed that Nb restored the mechanical properties and improved toughness due to a more homogeneous and refined microstructure. These studies collectively underscore Nb's critical role in enhancing the mechanical performance and microstructural stability of maraging steels.

The use of Electron Backscatter Diffraction (EBSD) in analyzing maraging steel's BCC matrix aims to provide a comprehensive understanding of its microstructural characteristics, crucial for optimizing mechanical properties9-12. EBSD captures spatially-linked crystallographic orientation and phase information, which can be processed using various analytical tools12. These include Inverse Pole Figure (IPF) orientation images, which map the crystal directions relative to sample directions, and pole figures that plot the 3D orientation data in 2D for texture analysis9. Misorientation measurements identify disorientation angles between adjacent grains, crucial for understanding grain boundary properties. Texture analysis using Orientation Distribution Functions (ODFs) represents preferred crystallographic orientations, providing insights into the material's deformation and processing history. Effective grain size is determined by defining grain boundaries with a critical disorientation angle, highlighting grain size's influence on mechanical properties via the Hall-Petch relationship10. EBSD also measures dislocation density, a key factor in understanding the distribution and behavior of dislocations within the microstructure11. Pattern quality maps reveal features such as grains, grain boundaries, and surface damage, offering a detailed view of the material's microstructural integrity13. This suite of EBSD tools collectively enables a thorough interpretation of the maraging steel's microstructure, essential for enhancing its performance in engineering applications.

Our study experimentally produces a Titanium-reduced, Niobium-enhanced Maraging 300 steel, examining its microstructural evolution under hot rolling, aging, and hot rolling-solubilization at two different temperatures followed by aging. EBSD technique was used to characterize the microstructure, texture and grain boundaries characteristics.

2. Experimental Procedure

The experimental maraging steel with a nominal composition of 18.6 Ni - 9.5 Co - 5.1 Mo - 0.4 Ti - 0.6 Nb - 0.002 C (wt.%) was prepared using vacuum induction melting to cast ingots with a thickness of 70 mm. The ingots underwent a homogenization heat treatment at 1200 °C for 6 hours. Following this treatment, the ingots were rolled within a temperature range of 1200–1030 °C, reducing their thickness to final dimensions of 60×15 mm. This rolling process was followed by air cooling, referred to as HR in this paper.

Subsequent standard aging (precipitation hardening) at a temperature of 480 °C for 3 hours was conducted to produce a fine dispersion of Ni3(X,Y) intermetallic phases along dislocations left by the martensitic transformation, where X and Y are solute elements added for such precipitation. This treatment is denoted as HR-A480. To avoid microsegregation during the previous hot rolling, the hot-rolled samples were first solution annealed at 820 °C and 860 °C for 1 hour, followed by air cooling. These solution-annealed samples were then aged at 480 °C for 3 hours to compare the influence of different solid solution treatments on the deformed structure. These treatments are denoted as HR-S820-A480 and HR-S860-A480, respectively.

Microstructural analysis was conducted on the rolled samples along their rolling plane using a FEI Quanta FEG 450 scanning electron microscope equipped with an EBSD detector. For metallographic preparation, the samples underwent grinding with SiC papers up to a grit of 1200, then polishing using diamond pastes of 6, 3, and 1 μm. EBSD measurements were subsequently performed at a magnification of X1240, with a step size of 0.25 µm, an acceleration voltage of 20 kV, and a horizontal field width (HFW) of 300 μm. The EBSD data were thoroughly examined and interpreted using ATEX14 and ARPGE15 software's, respectively. The fine step size enhances resolution, enabling the detection of subtle microstructural features like martensite plates, retained austenite, or precipitates that are crucial for understanding the steel's strengthening mechanisms. This setup is particularly useful for assessing the homogeneity of the microstructure and identifying any localized strain or deformation patterns within the steel matrix.

X-ray diffraction (XRD) measurements were carried out using a detector with Co Kα1 radiation (wavelength = 1.788 Å), within a 2θ range of 40°-110°, using a step size of 0.005° and a counting time of 1.00 s per step. The position and broadening of each peak were analyzed, and the full width at half maximum (FWHM, β1/2) was recorded for each peak. Lattice microstrain and dislocation density for each sample were separately calculated using the methods proposed by Williamson16 and Murugesan17, based on individual peaks from a face-centered cubic (FCC) structure.

The lattice microstrain (ε) was determined using the relationship: εL2=β1/24.tanθ. The crystallite size (D) was estimated using the Scherrer equation: D=0.9λβ1/2.cosθ. The dislocation density (ρ) was calculated by considering both the contribution from the domain size (ρD) and lattice strain (ρS), with the total dislocation density (ρtot) expressed as: ρtot=ρD.ρS2. The contribution of the domain size to the dislocation density is given by: ρD=3/D2. The contribution of lattice strain to the dislocation density is: ρS=0.9εL2b2.

Here, b is the Burgers vector, defined as: b=a3, for the body-centered cubic (BCC) structure, where a is the lattice parameter for each peak/plane.

3. Results

Figure 1 shows the Inverse Pole Figure (IPF) and Orientation Image Microscopy (OIM) maps for the as-received hot-rolled (HR) sample. The microstructure reveals a fine martensite structure. Detailed IPF analysis indicates a predominance of the {112} crystal orientation along the rolling plane (RP), with the {112}//RP fiber texture comprising approximately 30%. In BCC metals, slip occurs in the close-packed <111> directions, but the slip plane can be any of the {110}, {112}, or {123} planes, each containing the close-packed <111> slip direction. The choice of slip plane is influenced by the temperature of deformation: below Tm/4, {112} slip predominates; between Tm/4 and Tm/2, {110} slip is favored; and above Tm/2, {123} slip is preferred. At room temperature, iron exhibits slip on all three planes in the common <111> direction, a process known as pencil glide18. Static Orientation Distribution Functions (ODFs) analyses revealed that (112)[11-1] is the primary texture in these hot-rolled and dynamically recrystallized samples.

Figure 1
IPF, OIM, and ODF maps, along with the volumetric fractions for each main texture fiber of the HR sample.

Figure 2a presents the effective grain size map. Effective grain size refers to grains enclosed by boundaries with a misorientation greater than 15°, known as high-angle boundaries (HABs), which have high stored energy and effectively serve as barriers to dislocation movement19. In this sample, EBSD analysis revealed an effective grain size of approximately 16.75±2.50 µm. The disorientation within the effective grains is determined by analyzing several spatial pixels within these grains20. Although their variation is less than 15°, there is a certain degree of disorientation due to crystal distortion and crystallographic and metallurgical defects. Regions near the boundaries and interfaces with the highest lattice distortion exhibited the most distortion. However, the average lattice distortion across the effective grain is approximately 5.20°±0.25° in the HR sample. Finally, the dislocation density map, based on the Entrywise Norm of the Nye Tensor analysis21, is presented in Figure 2c. The finer FeNi martensite laths show a well-distributed local stress/strain with a dislocation density of approximately 2.7E14 m-2.

Figure 2
(a) Grain sizes, (b) Disorientation/reference orientation, and (c) Crystal defects maps of the HR sample.

The HR sample was aged without solution annealing treatment to preserve the martensite structure. The EBSD analysis map of specimen submitted to aging treatment at 480 °C for 3 hours (sample HR-A480) is presented in Figure 3. Interestingly, we observed localized lath reorientation along the {101} orientation, which minimized lattice distortion. The {101} planes in the BCC martensite structure, being the most compact planes, possess the available slip system to facilitate crystallographic defects. Consequently, 36% of all spatial crystal orientations in this map aligned with the {101} planes. ODF analysis revealed that the (011)[11-1] crystal texture is the dominant texture. Liu et al.22 explained that after deformation, a significant amount of the (011) distorted structure had partially or entirely recovered, resulting in stress relief.

Figure 3
IPF, OIM, and ODF maps, along with the volumetric fractions for each main texture fiber of the HR-A480 sample.

HThe grain size map of the HR-A480 sample, shown in Figure 4a, indicates a slight refinement, with the effective grain size reduced to 12.0±1.2 µm. Additionally, the disorientation within the grains decreased to 4.25±0.25, as depicted in Figure 4b. It is understood that the (101) misorientation angle-axis can effectively absorb crystallographic defects and distribute them throughout the refined structure. Consequently, the dislocation densities in the sample, as analyzed in Figure 4c, also reduced to 2.6E14 m-2. According to Niu et al.23, nano-sized Ni3Ti precipitates in FeNi lath martensite after aging cause a high dislocation density around 5E14 m-2.

Figure 4
(a) Grain sizes, (b) Disorientation/reference orientation, and (c) Crystal defects maps of the HR-A480 sample.

Figure 5 details the OIM, IPF, ODFs, and texture fiber volumetric fractions of solution treated sample at 820 °C for one hour and aged at 480 °C for three hours, labeled HR-S820-A480. Notably, the {111} crystal orientation, aligned with the rolling direction, was predominant. Although austenitization can help eliminate texture, further transformation into the martensite phase selectively promotes specific atomic orientations that minimize internal energy, favoring the most compact <111> directions. ODF analyses revealed a dominant (111)[10-1] crystal orientation, aligning with both the compact plane and direction in the resultant martensite structure. Additionally, Haghdadi et al.24 reported on the {111} crystal orientation's predominance, its relationship with the rolling direction, and the role of austenitization in altering texture. Their findings demonstrate a 'texture memory' effect where austenite retains similar textures to the as-received condition despite phase transformations, crucially explaining the persistent predominance of the {111} orientation.

Figure 5
IPF, OIM, and ODF maps, along with the volumetric fractions for each main texture fiber of the HR-S820-A480 sample.

The grain size map shown in Figure 6a indicates that the average effective grain size varies between 10-12 µm, demonstrating that the austenitization process did not cause significant grain growth. Additionally, the disorientation within the effective grain is slightly reduced to 3.5±0.25°, the resulting grain refinement and lower internal disorientation, as shown in Figure 6b. Dislocation density is estimated at approximately 1.9E14 m-2, depicted in Figure 6c. This density suggests that martensite transformation from the prior austenite likely induces localized micro-stress/strain within individual fine austenite grains, in line with findings from Yadunandan et al.25.

Figure 6
(a) Grain sizes, (b) Disorientation/reference orientation, and (c) Crystal defects maps of the HR-S820-A480 sample.

Austenitization at higher temperatures such as 860 °C on the one hand could potentially increase the prior austenite grain size and decrease the mechanical properties of the investigated samples. On the other hand, solubilization at higher temperatures offers benefits such as more homogeneous austenite26. The IPF, OIM, texture fiber volumetric fraction, and ODF maps of the HR-S860-A480 sample are shown in Figure 7. Static crystallographic data analyses reveal a slight predominance of {212} and {001} crystal orientations in these samples. This indicates that solubilization at this higher temperature could eliminate the previous texture observed in the HR-S820-A480 sample. However, the {001} cleavage plane could deteriorate mechanical properties as there is no activate slip system in these planes, making the grains prone to accumulating stress/strain concentration without dislocation movement or stress relief27.

Figure 7
IPF, OIM, and ODF maps, along with the volumetric fractions for each main texture fiber of the HR-S860-A480 sample.

Figure 8 displays the effective grain size, disorientation throughout the individual grain, and dislocation density of the HR-S860-A480 sample. The effective grain size has increased to 18.4±3.2 µm, indicating grain growth due to the higher solution temperature. Disorientation throughout the individual grain also increased to 4.8±0.3°, slightly lower than in the HR condition. This amplifies the impact of the formation of {001} planes, which lack of activated slip system. Consequently, crystallographic defects cannot traverse the grains, leading to increased lattice distortion. Accordingly, the dislocation density has risen to 2.7E14 m-2.

Figure 8
(a) Grain sizes, (b) Disorientation/reference orientation, and (c) Crystal defects maps of the HR-S860-A480 sample.

3.1. XRD analysis for microstructure investigation

Figure 9 presents the XRD patterns of the investigated samples, where the Ti content was reduced and Nb was added. As expected, the martensitic matrix exhibited characteristic peaks corresponding to the BCC phase. However, specific new peaks were observed at approximately ~48.64° and ~59.78°. To identify these possible phases, Nb-based precipitates and alloying elements were analyzed. The most probable phases include Iron Niobium (Fe7Nb3, reference code: 00-012-0599) and Niobium Nitride (Nb4N3.4, reference code: 01-089-5132). Notably, with increasing solution treatment temperature, these precipitates dissolved into the matrix, indicating greater stability of the martensitic structure at elevated temperatures.

Figure 9
XRD patterns of the investigated samples.

Leveraging methodologies established by Bragg, Williamson, Scherrer, and Murugesan, the lattice parameter, lattice microstrain, and dislocation density were calculated using the position and broadening of each peak in the XRD patterns. The precise determination of the lattice parameter was achieved using peak positions, while peak broadening provided insights into lattice microstrain and crystallite size, enabling dislocation density estimation. These parameters were determined using standard approaches, such as the Scherrer equation to correlate peak broadening with crystallite size and the Williamson-Hall equation to separate the contributions of microstrain and crystallite size to peak broadening. The dislocation density was subsequently calculated by combining the contributions from crystallite size and lattice microstrain. The calculated values for lattice parameter, microstrain, and dislocation density, along with their respective uncertainties, are summarized in Table 1.

Table 1
Analysis of lattice parameter, lattice microstrain, and dislocation density calculated from XRD patterns.

The analysis of the calculated values reveals that the highest lattice parameter (2.870 Å) and the lowest lattice microstrain (0.262) were observed in the HR-S820-A480 sample, indicating a relatively relaxed crystalline structure with reduced internal strain. Conversely, the HR-S860-A480 sample exhibited the highest microstrain (0.296) and dislocation density (2.415 × 1014 m-2), suggesting a more distorted lattice structure. The HR sample displayed the lowest dislocation density (2.036 × 1014 m-2), reflecting minimal defects, while HR-S860-A480 showed the largest number of lattice defects due to the combined effects of solution treatment and alloying. These results highlight the impact of processing conditions and alloying elements on the microstructure and intrinsic properties of the investigated samples.

4. Discussion

The results from Figure 10 illustrate a distinctive variation in the volumetric fraction of main fiber textures among different samples, which significantly influences their mechanical properties and formability28. The HR sample exhibited the highest volumetric fraction of the {112}//RP crystallographic texture, a result of the activated {112} slip plane at high rolling temperatures, consistent with prior findings on BCC metals where {112} slip predominates at lower deformation temperatures29. The transformation of texture during aging, observed in the HR-A480 sample, where the {110}//RP texture became more prevalent, indicates a typical re-orientation of martensitic structures under thermal influence to minimize internal energy and stress, aligning with the mechanism described as 'pencil glide' in BCC metals30. This texture realignment is critical for applications requiring specific directional properties and highlights the importance of controlled rolling and aging processes in optimizing the microstructural characteristics of maraging steels.

Figure 10
Comparative analysis of volumetric fractions of main fiber textures.

In contrast, the HR-S820-A480 sample showed a remarkable increase in the {111} texture, which is beneficial for ductility and toughness due to its dense atomic packing and multiple slip systems availability, as noted in the enhanced dislocation movement without significant pile-up or strain accumulation. The predominant {111} orientation aligns with previous studies, suggesting that this texture can enhance mechanical performance and more favorable stress distribution across the grain structure31.

However, the HR-S860-A480 sample, which exhibited a predominant {001} texture, underscores the challenges associated with higher solubilization temperatures. Although such treatments aim to achieve grain homogeneity and increased solute atoms mobility, they inadvertently enhance the {001} cleavage planes, making the material more prone to brittle failure under stress. This observation is critical for applications where fracture toughness is a crucial requirement, and it suggests that lower solubilization temperatures might be more beneficial in maintaining an optimal balance between strength and ductility. The predominance of the {001} texture and its impact on the mechanical properties further corroborates studies highlighting the role of texture in influencing the fracture and deformation behavior of BCC martensite structure32,33.

Figure 11a illustrates the variation in grain disorientation, revealing that the HR-S860-A480 sample exhibits a high proportion of boundaries with angles ranging from 50-60°. This high proportion of boundaries indicates significant lattice distortion at the martensite lath interfaces, acting as centers of strain concentration. Similar findings are depicted in Figure 11b, where the highest disorientation relative to the reference effective grain orientation corresponds to the HR-S860-A480 sample. This suggests that prior austenite grain growth, due to solubilization at higher temperatures, reduces martensite nucleation sites, leading to collisions at sharp lath interfaces and, consequently, the highest dislocation density among the samples, as shown in Figure 11c. Comparatively, no significant differences are revealed between the HR and HR-A480 samples. However, the HR-S820-A480 sample displays significantly lower dislocation densities than all other samples, which can be well justified by the {111} crystallographic texture that enhances dislocation movement, as previously discussed. Figure 11d compares the variations in average grain size, grain disorientation, and dislocation density across all investigated samples, revealing that the HR-S820-A480 sample exhibits the slightest crystallographic defects and minor grain distortion.

Figure 11
Comprehensive analysis of microstructural characteristics. (a) Grain disorientation maps, (b) Disorientation relative to the reference effective grain orientation, (c) Dislocation density maps, and (d) Comparative analysis of average grain size, grain disorientation, and dislocation density.

Using EBSD analyses of martensite products, ARPGE software15,34 was employed to reconstruct the parent austenite grains using the Kurdjumov-Sachs (K-S) orientation relationship, with a tolerance of 3° for parent grain nucleation and 6° for parent grain growth. Figure 12 provides a detailed visualization of these results, presenting Euler maps of the daughter grains (room temperature martensite lath) in the first row, parent reconstruction Euler maps for each condition in the second row, and combined maps of parent grain boundaries and daughter grains in the third row.

Figure 12
Microstructural analysis of Nb-enhanced maraging steel. Top: Euler maps of martensite lath.

The comparative analysis illustrated in Figure 12 reveals several critical insights. Applying solubilization treatment after the thermomechanical process significantly alters the microstructural features while maintaining the martensite structure. Notably, there is a remarkable change in crystallographic orientation, as seen in the first row of Figure 12. The influence of Nb on preserving the austenite grain size is evident from the parent reconstruction maps, where the HR sample and solubilized samples (S820 and S860) do not show an increase in prior austenite grain size, underscoring Nb's role in stabilizing grain size. Furthermore, the combined maps of parent grain boundaries and the resulting martensite structure demonstrate the utility of this mathematical program in designing appropriate plastic deformation in the austenite region, achieving optimal martensite orientations and ultra-refinement.

Middle: Reconstructed parent austenite grains. Bottom: Combined parent grain boundaries and martensite structures. Shows impact of solubilization treatments on grain structure and orientation.

5. Conclusion

  • The HR-S820-A480 sample demonstrates a predominant {111} orientation, enhancing ductility and mechanical performance due to optimal stress distribution across the grain structure.

  • Solubilization at 820°C followed by aging significantly improves the HR-S820-A480 sample’s ductility and reduces dislocation density to 1.9E14 m-2, making it ideal for high-performance engineering applications.

  • The HR-S860-A480 sample highlights the drawbacks of higher solubilization temperatures, such as increased {001} texture presence, which exacerbates brittleness and raises concerns for applications requiring high fracture toughness.

  • Comparative analysis shows the HR-S820-A480 sample as the most resilient, with the lowest incidence of crystallographic defects and grain distortion and no detected retained austenite, underlining its superior structural integrity.

6. Acknowledgments

The authors greatly appreciate the support from CNPQ (Process: 408380/2022-5) and are grateful Villares Metals for the production of the experimental maraging steel used in this study. The microstructural characterization was performed at the Central Analítica UFC/CT - INFRAFINEP/Pro-Equipamentos - CAPES/CNPq-SisNano-MCTI 2019 (Grant 442577/2019-2) - INCT FUNCAP.

  • Data Availability
    Data will be made available on request.

7. References

  • 1 Parvinian S, Sievers DE, Garmestani H, Kalidindi SR. Evaluation of the influence of B and Nb microalloying on the microstructure and strength of 18% Ni maraging steels (C350) using hardness, spherical indentation and tensile tests. Acta Mater. 2021;215:117071. http://doi.org/10.1016/j.actamat.2021.117071
    » http://doi.org/10.1016/j.actamat.2021.117071
  • 2 Li H, Liu Y, Liu B, Zhao W. Effects of Cu and Nb on martensite hardening and transformation-induced plastic deformation behavior of maraging TRIP-aided steel. J Mater Res Technol. 2023;24:4271-84. http://doi.org/10.1016/j.jmrt.2023.04.035
    » http://doi.org/10.1016/j.jmrt.2023.04.035
  • 3 Zhou X, Zeng T, Shi X, Zhao M, Wang W, Li Y, et al. An ultra-high strength bainitic aging steel. Scr Mater. 2024;245:116061. http://doi.org/10.1016/j.scriptamat.2024.116061
    » http://doi.org/10.1016/j.scriptamat.2024.116061
  • 4 Lee DG, Jang KC, Kuk JM, Kim IS. The influence of niobium and aging treatment in the 18% Ni maraging steel. J Mater Process Technol. 2005;162-163:342-9. http://doi.org/10.1016/j.jmatprotec.2005.02.102
    » http://doi.org/10.1016/j.jmatprotec.2005.02.102
  • 5 Li Z, Zhang R, Zha Q, Wang Y, Qiang J, Dong C. Composition design of superhigh strength maraging stainless steels using a cluster model. Prog Nat Sci. 2014;24(1):35-41. http://doi.org/10.1016/j.pnsc.2014.01.010
    » http://doi.org/10.1016/j.pnsc.2014.01.010
  • 6 Sha W, Ye A, Malinov S, Wilson EA. Microstructure and mechanical properties of low nickel maraging steel. Mater Sci Eng A. 2012;536:129-35. http://doi.org/10.1016/j.msea.2011.12.086
    » http://doi.org/10.1016/j.msea.2011.12.086
  • 7 Hossein Nedjad S, Teimouri J, Tahmasebifar A, Shirazi H, Nili Ahmadabadi M. A new concept in further alloying of Fe-Ni-Mn maraging steels. Scr Mater. 2009;60(7):528-31. http://doi.org/10.1016/j.scriptamat.2008.11.046
    » http://doi.org/10.1016/j.scriptamat.2008.11.046
  • 8 Ahmed M, Salam I, Nasim I, Hussain SW, Hashmi FH, Khan AQ. Reclamation and additional alloying of 18Ni(350) maraging steel. J Mater Eng Perform. 1994;3(3):386-92. http://doi.org/10.1007/BF02645336
    » http://doi.org/10.1007/BF02645336
  • 9 Wisniewski W, Saager S, Böbenroth A, Rüssel C. Experimental evidence concerning the significant information depth of electron backscatter diffraction (EBSD). Ultramicroscopy. 2017;173:1-9. http://doi.org/10.1016/j.ultramic.2016.11.004
    » http://doi.org/10.1016/j.ultramic.2016.11.004
  • 10 Alam MK, Mehdi M, Urbanic RJ, Edrisy A. Electron Backscatter Diffraction (EBSD) analysis of laser-cladded AISI 420 martensitic stainless steel. Mater Charact. 2020;161:110138. http://doi.org/10.1016/j.matchar.2020.110138
    » http://doi.org/10.1016/j.matchar.2020.110138
  • 11 Mineta T. Surface treatment and electron backscatter diffraction (EBSD) analysis of the body-centered cubic phase in Mg-Li based alloys processed by severe plastic deformation. Micron. 2020;137:102914. http://doi.org/10.1016/j.micron.2020.102914
    » http://doi.org/10.1016/j.micron.2020.102914
  • 12 Beirau T, Kilian R, Stipp M, Ewing RC. Probing structural disorder in zircon by electron backscatter diffraction (EBSD): radiation damage and Kikuchi pattern. J Nucl Mater. 2023;581:154440. http://doi.org/10.1016/j.jnucmat.2023.154440
    » http://doi.org/10.1016/j.jnucmat.2023.154440
  • 13 Koko A, Tong V, Wilkinson AJ, Marrow TJ. An iterative method for reference pattern selection in high-resolution electron backscatter diffraction (HR-EBSD). Ultramicroscopy. 2023;248:113705. http://doi.org/10.1016/j.ultramic.2023.113705
    » http://doi.org/10.1016/j.ultramic.2023.113705
  • 14 Beausir B, Fundenberger J-J. Analysis Tools for Electron and X-Ray Diffraction, ATEX – Software. Metz: Université de Lorraine; 2017 [cited 2024 Sep 21]. Available from: Www.Atex-Software.Eu
    » Www.Atex-Software.Eu
  • 15 Cayron C. ARPGE: a computer program to automatically reconstruct the parent grains from electron backscatter diffraction data. J Appl Crystallogr. 2007;40(Pt 6):1183-8. http://doi.org/10.1107/S0021889807048777
    » http://doi.org/10.1107/S0021889807048777
  • 16 Williamson GK, Hall WH. X-ray line broadening from filed aluminium and wolfram. Acta Metall. 1953;1(1):22-31. http://doi.org/10.1016/0001-6160(53)90006-6
    » http://doi.org/10.1016/0001-6160(53)90006-6
  • 17 Murugesan S, Kuppusami P, Mohandas E, Vijayalakshmi M. X-ray diffraction Rietveld analysis of cold worked austenitic stainless steel. Mater Lett. 2012;67(1):173-6. http://doi.org/10.1016/j.matlet.2011.09.065
    » http://doi.org/10.1016/j.matlet.2011.09.065
  • 18 Humphreys F, Hatherly M. Recrystallization and related annealing phenomena. Oxford: Pergamon; 2017.
  • 19 Miyamoto G, Iwata N, Takayama N, Furuhara T. Mapping the parent austenite orientation reconstructed from the orientation of martensite by EBSD and its application to ausformed martensite. Acta Mater. 2010;58(19):6393-403. http://doi.org/10.1016/j.actamat.2010.08.001
    » http://doi.org/10.1016/j.actamat.2010.08.001
  • 20 Suikkanen PP, Cayron C, DeArdo AJ, Karjalainen LP. Crystallographic analysis of martensite in 0.2C-2.0Mn-1.5Si-0.6Cr steel using EBSD. J Mater Sci Technol. 2011;27(10):920-30. http://doi.org/10.1016/S1005-0302(11)60165-5
    » http://doi.org/10.1016/S1005-0302(11)60165-5
  • 21 Tu X, Shi X, Yan W, Li C, Shi Q, Shan Y, et al. Tensile deformation behavior of ferrite-bainite dual-phase pipeline steel. Mater Sci Eng A. 2022;831:142230. http://doi.org/10.1016/j.msea.2021.142230
    » http://doi.org/10.1016/j.msea.2021.142230
  • 22 Liu Y, Xie ZL, Van Humbeeck J, Delaey L. Effect of texture orientation on the martensite deformation of NiTi shape memory alloy sheet. Acta Mater. 1999;47(2):645-60. http://doi.org/10.1016/S1359-6454(98)00376-0
    » http://doi.org/10.1016/S1359-6454(98)00376-0
  • 23 Niu M, Zhou G, Wang W, Shahzad MB, Shan Y, Yang K. Precipitate evolution and strengthening behavior during aging process in a 2.5 GPa grade maraging steel. Acta Mater. 2019;179:296-307. http://doi.org/10.1016/j.actamat.2019.08.042
    » http://doi.org/10.1016/j.actamat.2019.08.042
  • 24 Haghdadi N, Cizek P, Hodgson PD, Tari V, Rohrer GS, Beladi H. Effect of ferrite-to-austenite phase transformation path on the interface crystallographic character distributions in a duplex stainless steel. Acta Mater. 2018;145:196-209. http://doi.org/10.1016/j.actamat.2017.11.057
    » http://doi.org/10.1016/j.actamat.2017.11.057
  • 25 Yadunandan D. Characterization of stresses and strains involved in the martensitic phase transformations [thesis]. Reino Unido: The Open University; 2017.
  • 26 Masoumi M, de Barros IF, Herculano LFG, Coelho HLF, de Abreu HFG. Effect of microstructure and crystallographic texture on the Charpy impact test for maraging 300 steel. Mater Charact. 2016;120:203-9. http://doi.org/10.1016/j.matchar.2016.09.003
    » http://doi.org/10.1016/j.matchar.2016.09.003
  • 27 Barik RK, Ghosh A, Basiruddin Sk M, Biswal S, Dutta A, Chakrabarti D. Bridging microstructure and crystallography with the micromechanics of cleavage fracture in a lamellar pearlitic steel. Acta Mater. 2021;214:116988. http://doi.org/10.1016/j.actamat.2021.116988
    » http://doi.org/10.1016/j.actamat.2021.116988
  • 28 Diehl M, Niehuesbernd J, Bruder E. Quantifying the contribution of crystallographic texture and grain morphology on the elastic and plastic anisotropy of bcc steel. Metals (Basel). 2019;9(12):1252. http://doi.org/10.3390/met9121252
    » http://doi.org/10.3390/met9121252
  • 29 Chatterjee A, Sk MB, Ghosh A, Mitra R, Chakrabarti D. Combining crystal plasticity and electron microscopy to elucidate texture dependent micro-mechanisms of tensile deformation in lath martensitic steel. Int J Plast. 2022;153:103251. http://doi.org/10.1016/j.ijplas.2022.103251
    » http://doi.org/10.1016/j.ijplas.2022.103251
  • 30 Na H, Nambu S, Ojima M, Inoue J, Koseki T. Crystallographic and microstructural studies of lath martensitic steel during tensile deformation. Metall Mater Trans, A Phys Metall Mater Sci. 2014;45(11):5029-43. http://doi.org/10.1007/s11661-014-2461-4
    » http://doi.org/10.1007/s11661-014-2461-4
  • 31 Kitsuya S, Ohtsubo H, Fujita N, Ichimiya K, Hase K. Effect of crystallographic texture on anisotropy of mechanical properties in high strength martensitic steel. ISIJ Int. 2020;60(2):346-51. http://doi.org/10.2355/isijinternational.ISIJINT-2019-164
    » http://doi.org/10.2355/isijinternational.ISIJINT-2019-164
  • 32 Lobanov ML, Pyshmintsev IY, Urtsev VN, Danilov SV, Urtsev NV, Redikultsev AA. Texture inheritance in the ferrito-martensite structure of low-alloy steel after thermomechanical controlled processing. Phys Met Metallogr. 2019;120(12):1180-6. http://doi.org/10.1134/S0031918X1912010X
    » http://doi.org/10.1134/S0031918X1912010X
  • 33 Masoumi M, de Barros IF, Herculano LFG, Coelho HLF, de Abreu HFG. Effect of microstructure and crystallographic texture on the Charpy impact test for maraging 300 steel. Mater Charact. 2016;120:203-9. http://doi.org/10.1016/j.matchar.2016.09.003
    » http://doi.org/10.1016/j.matchar.2016.09.003
  • 34 Li D, Zhao X, Zhang H, Li J, Han H. The effect of network cementite dissolution on the nucleation and growth of prior austenite grains in high carbon low alloy steels. J Mater Res Technol. 2024;30:565-79. http://doi.org/10.1016/j.jmrt.2024.03.046
    » http://doi.org/10.1016/j.jmrt.2024.03.046

Data availability

Data will be made available on request.

Publication Dates

  • Publication in this collection
    19 May 2025
  • Date of issue
    2025

History

  • Received
    21 Sept 2024
  • Reviewed
    08 Jan 2025
  • Accepted
    02 Feb 2025
location_on
ABM, ABC, ABPol UFSCar - Dep. de Engenharia de Materiais, Rod. Washington Luiz, km 235, 13565-905 - São Carlos - SP- Brasil. Tel (55 16) 3351-9487 - São Carlos - SP - Brazil
E-mail: pessan@ufscar.br
rss_feed Acompanhe os números deste periódico no seu leitor de RSS
Reportar erro