3,151
Views
0
CrossRef citations to date
0
Altmetric
Review Article

Process-microstructure-corrosion of additively manufactured steels: a review

, , &

Figures & data

Figure 1. Research outputs on the topic of “metal additive manufacturing” (a) documents by year, and (b) documents by subject area (Adapted with permission from[Citation3]).

Figure 1. Research outputs on the topic of “metal additive manufacturing” (a) documents by year, and (b) documents by subject area (Adapted with permission from[Citation3]).

Figure 2. Schematic design of (a) water atomization, and (b) gas atomization processes of metal powder (Reproduced with permission from[Citation22,Citation31]).

Figure 2. Schematic design of (a) water atomization, and (b) gas atomization processes of metal powder (Reproduced with permission from[Citation22,Citation31]).

Figure 3. Gas atomization techniques based on melting mechanisms and gas nozzle design (Reproduced with permission from[Citation29]).

Figure 3. Gas atomization techniques based on melting mechanisms and gas nozzle design (Reproduced with permission from[Citation29]).

Figure 4. (a) Capillary waves causing the liquid stream atomization on the top surface of the transducer Horn, and (b) a layout of the UA setup (Reproduced with permission from[Citation36]).

Figure 4. (a) Capillary waves causing the liquid stream atomization on the top surface of the transducer Horn, and (b) a layout of the UA setup (Reproduced with permission from[Citation36]).

Figure 5. (a) Schematic layout of the rotating disk atomization of liquid stream, and (b) droplet formation during the CA process (Reproduced with permission from[Citation47]).

Figure 5. (a) Schematic layout of the rotating disk atomization of liquid stream, and (b) droplet formation during the CA process (Reproduced with permission from[Citation47]).

Figure 6. (a) the Fe-Ni-Cr Isopleth at 19% Cr showing various solidification paths,[Citation68] (b) solidification modes in austenitic stainless steels,[Citation68] (c and d) EBSD phase maps and corresponding inverse pole figure maps of virgin and reused 316 L stainless steel powders (Reproduced with permission from[Citation74]).

Figure 6. (a) the Fe-Ni-Cr Isopleth at 19% Cr showing various solidification paths,[Citation68] (b) solidification modes in austenitic stainless steels,[Citation68] (c and d) EBSD phase maps and corresponding inverse pole figure maps of virgin and reused 316 L stainless steel powders (Reproduced with permission from[Citation74]).

Table 1. Various types of powder morphology in ASTM B243.[Citation89]

Figure 7. SEM micrographs and optical Appearance (in the insets) of 316 L stainless steel powder produced by (a) GA-, and (b) WA-techniques (Reproduced with permission from[Citation93]).

Figure 7. SEM micrographs and optical Appearance (in the insets) of 316 L stainless steel powder produced by (a) GA-, and (b) WA-techniques (Reproduced with permission from[Citation93]).

Figure 8. Gas atomized 316 L stainless steel (a) virgin powder, (b) recycled powder after EBM cycle,[Citation97] and (c,d) small satellites and surface defects in 304 L stainless steel powder (Reproduced with permission from[Citation99]).

Figure 8. Gas atomized 316 L stainless steel (a) virgin powder, (b) recycled powder after EBM cycle,[Citation97] and (c,d) small satellites and surface defects in 304 L stainless steel powder (Reproduced with permission from[Citation99]).

Table 2. Conventional techniques for powder size measurement relevant to AM.[Citation88,Citation111]

Figure 9. (a) Particle size distribution of gas atomized AISI H13 tool steel powder, and (b) the SEM-SE micrograph (Reproduced with permission from[Citation109]).

Figure 9. (a) Particle size distribution of gas atomized AISI H13 tool steel powder, and (b) the SEM-SE micrograph (Reproduced with permission from[Citation109]).

Figure 10. Bottom-up diagram presenting the effect of powder properties on the properties of the bulk powder, in-process performance, and the as-built product (Reproduced with permission from[Citation114]).

Figure 10. Bottom-up diagram presenting the effect of powder properties on the properties of the bulk powder, in-process performance, and the as-built product (Reproduced with permission from[Citation114]).

Figure 11. Illustration of two techniques that measure the apparent density: (a) a Scott volumeter, and (b) an Arnold meter (Reproduced with permission from[Citation13]).

Figure 11. Illustration of two techniques that measure the apparent density: (a) a Scott volumeter, and (b) an Arnold meter (Reproduced with permission from[Citation13]).

Figure 12. An interplay among the terms “flowability,” “flow properties,” and respective parameters (Reproduced with permission from[Citation114]).

Figure 12. An interplay among the terms “flowability,” “flow properties,” and respective parameters (Reproduced with permission from[Citation114]).

Table 3. Degrees of powder flowability in terms of the Hausner ratio.[Citation88]

Table 4. Gas concentration and porosity level in rapidly solidified 304 SS metal powder.[Citation127]

Figure 13. Schematic Illustration of the bag break-up or folding mechanisms during the powder production (Adapted with permission from[Citation126]).

Figure 13. Schematic Illustration of the bag break-up or folding mechanisms during the powder production (Adapted with permission from[Citation126]).

Figure 14. Ishikawa diagram of processing parameters controlling the properties of LPBF product (Reproduced with permission from[Citation131]).

Figure 14. Ishikawa diagram of processing parameters controlling the properties of LPBF product (Reproduced with permission from[Citation131]).

Figure 15. (a) Complete-fusion criterion relied on melt pool geometry, where Md and Md0 are the melt pool depth and the minimum ELT to have complete fusion (Adapted with permission from[Citation132]), (b) a relationship between the AM product quality and the optimum laser energy density,[Citation130] (c) transverse section of a single-track cladding layer of 1Cr18Ni9Ti stainless steel at ELT of 60, 80, and 100 μm,[Citation137] (d) bridge-like sample before and after detachment from the substrate.[Citation140]

Figure 15. (a) Complete-fusion criterion relied on melt pool geometry, where Md and Md0 are the melt pool depth and the minimum ELT to have complete fusion (Adapted with permission from[Citation132]), (b) a relationship between the AM product quality and the optimum laser energy density,[Citation130] (c) transverse section of a single-track cladding layer of 1Cr18Ni9Ti stainless steel at ELT of 60, 80, and 100 μm,[Citation137] (d) bridge-like sample before and after detachment from the substrate.[Citation140]

Table 5. Specifications of beam sources used in AM.[Citation141]

Figure 16. (a) Various types of power-density distribution,[Citation144] (b) energy density vs. laser power at three types of melt zone profile, 0: no deposition, 1: low substrate wetting between a melt bead and the substrate, 2: good substrate wetting by a melt bead showing no substrate penetration, 3: shallow substrate penetration with conduction mode laser melting, 4: Intermediate substrate penetration, 5: deep substrate penetration with key-hole laser melting, and (c) solidification microstructure of the melt pool at different energy densities and powers, 0: no fusion, 1: equiaxed, 2: equiaxed-columnar, 3: columnar (Reproduced with permission from[Citation145]).

Figure 16. (a) Various types of power-density distribution,[Citation144] (b) energy density vs. laser power at three types of melt zone profile, 0: no deposition, 1: low substrate wetting between a melt bead and the substrate, 2: good substrate wetting by a melt bead showing no substrate penetration, 3: shallow substrate penetration with conduction mode laser melting, 4: Intermediate substrate penetration, 5: deep substrate penetration with key-hole laser melting, and (c) solidification microstructure of the melt pool at different energy densities and powers, 0: no fusion, 1: equiaxed, 2: equiaxed-columnar, 3: columnar (Reproduced with permission from[Citation145]).

Figure 17. (a) How the scanning speed affects the deposition shape of a single track in 316 L stainless steel,[Citation150] (b,c) Macrohardness and relative density variations in 18Ni-300 steel samples at various scanning speeds and layer thicknesses (t) (Reproduced with permission from[Citation161]).

Figure 17. (a) How the scanning speed affects the deposition shape of a single track in 316 L stainless steel,[Citation150] (b,c) Macrohardness and relative density variations in 18Ni-300 steel samples at various scanning speeds and layer thicknesses (t) (Reproduced with permission from[Citation161]).

Figure 18. Effect of hatch spacing on (a) cooling rate of the melt pools, and (b) LOF voids percentage in 316 L stainless steel (Reproduced with permission from[Citation164]).

Figure 18. Effect of hatch spacing on (a) cooling rate of the melt pools, and (b) LOF voids percentage in 316 L stainless steel (Reproduced with permission from[Citation164]).

Figure 19. Layout of different scanning strategies (Reproduced with permission from[Citation165]).

Figure 19. Layout of different scanning strategies (Reproduced with permission from[Citation165]).

Figure 20. (a–c) Residual stresses and deformation contours, (d) max S11 and S22 stresses, (e) max deflection magnitude at various scanning strategies (Reproduced with permission from[Citation165]).

Figure 20. (a–c) Residual stresses and deformation contours, (d) max S11 and S22 stresses, (e) max deflection magnitude at various scanning strategies (Reproduced with permission from[Citation165]).

Table 6. Summary on the effect of scan strategy on the properties of AM steels.[Citation167]

Figure 21. (a,b) Schematic Representations of process by-products and spherical particles formed through redeposition of them,[Citation170] and (c–e) the effect of normal gas flow velocity on porosity level, melt pool width and penetration (1–9 correspond to location numbers on the build plate given in legend) (Reproduced with permission from[Citation172]).

Figure 21. (a,b) Schematic Representations of process by-products and spherical particles formed through redeposition of them,[Citation170] and (c–e) the effect of normal gas flow velocity on porosity level, melt pool width and penetration (1–9 correspond to location numbers on the build plate given in legend) (Reproduced with permission from[Citation172]).

Figure 22. (a,b) Melt pool shape in the LPBF SS 316 L printed via Gaussian and elliptical beam profiles, respectively, (c–g) ALE3D-assisted FE simulation of a melt track fused using Gaussian, transverse elliptical, longitudinal, annular, and Bessel beam profiles, respectively (mean beam diameter: 100 μm, laser power: 550 W, scanning speed: 1800 mm/s, domain size: 250 × 250 × 750 μm3) (Reproduced with permission from[Citation180]).

Figure 22. (a,b) Melt pool shape in the LPBF SS 316 L printed via Gaussian and elliptical beam profiles, respectively, (c–g) ALE3D-assisted FE simulation of a melt track fused using Gaussian, transverse elliptical, longitudinal, annular, and Bessel beam profiles, respectively (mean beam diameter: 100 μm, laser power: 550 W, scanning speed: 1800 mm/s, domain size: 250 × 250 × 750 μm3) (Reproduced with permission from[Citation180]).

Figure 23. Grain morphology over a cross-section of melt track roots at (a-c) various laser powers (C-M: circular-medium size beam profile) and (d-f) various beam profiles where the equiaxed area is 2%, 28%, and 77% for the C-M, LE-M, and T E-M, respectively (C: circular, LE: longitudinal elliptical, TE: transverse elliptical) (Reproduced with permission from[Citation145]), (g,h) optical micrographs of melt pools in samples scanned parallel and perpendicular to shielding gas direction, respectively, and (i,j) their corresponding IPF maps (Reproduced with permission from[Citation183]).

Figure 23. Grain morphology over a cross-section of melt track roots at (a-c) various laser powers (C-M: circular-medium size beam profile) and (d-f) various beam profiles where the equiaxed area is 2%, 28%, and 77% for the C-M, LE-M, and T E-M, respectively (C: circular, LE: longitudinal elliptical, TE: transverse elliptical) (Reproduced with permission from[Citation145]), (g,h) optical micrographs of melt pools in samples scanned parallel and perpendicular to shielding gas direction, respectively, and (i,j) their corresponding IPF maps (Reproduced with permission from[Citation183]).

Figure 24. A Schematic layout of typical microstructures obtained after LPBF and DED processing of various steels, ret.: retained, GB: grain boundary, α: ferrite – bcc, α: martensite – bcc/bct, γ: austenite – fcc (Adapted with permission from[Citation64]).

Figure 24. A Schematic layout of typical microstructures obtained after LPBF and DED processing of various steels, ret.: retained, GB: grain boundary, α: ferrite – bcc, α′: martensite – bcc/bct, γ: austenite – fcc (Adapted with permission from[Citation64]).

Figure 25. (a) A schematic layout of grain morphology evolution in a solidifying single melt pool (Adapted with permission from[Citation156]), (b) a lateral section of a direct-energy-deposited specimen of SS 316L (Reproduced with permission of [Citation215]), (c, d) solute banding lines (red arrows) in the LPBF SS 316L (Reproduced with permission from[Citation212]).

Figure 25. (a) A schematic layout of grain morphology evolution in a solidifying single melt pool (Adapted with permission from[Citation156]), (b) a lateral section of a direct-energy-deposited specimen of SS 316L (Reproduced with permission of [Citation215]), (c, d) solute banding lines (red arrows) in the LPBF SS 316L (Reproduced with permission from[Citation212]).

Figure 26. (a,b) IPF maps of AM SS 316 L printed under 1000 and 400 W beam powers, respectively,[Citation210] (c,d) effect of ultrasound treatment on ΔTC magnitude (CS zone), epitaxial columnar growth, cavitation effect, and columnar-to-equiaxed transition time (TE, TA: equilibrium liquidus-, and actual temperatures, ΔTC,ΔTn: constitutional-, and nucleation undercooling, t: time),[Citation219] (e–g) IPF maps of AM Fe-Ti alloy containing 2, 5, 7.5 wt.% Ti, respectively (Reproduced with permission from[Citation222]).

Figure 26. (a,b) IPF maps of AM SS 316 L printed under 1000 and 400 W beam powers, respectively,[Citation210] (c,d) effect of ultrasound treatment on ΔTC magnitude (CS zone), epitaxial columnar growth, cavitation effect, and columnar-to-equiaxed transition time (TE, TA: equilibrium liquidus-, and actual temperatures, ΔTC,ΔTn: constitutional-, and nucleation undercooling, t: time),[Citation219] (e–g) IPF maps of AM Fe-Ti alloy containing 2, 5, 7.5 wt.% Ti, respectively (Reproduced with permission from[Citation222]).

Figure 27. Role of Ṫ in determining (a) width, and (b) length of columnar grains in SS 316 L alloy produced by various AM techniques (Reproduced with permission from[Citation230]).

Figure 27. Role of Ṫ in determining (a) width, and (b) length of columnar grains in SS 316 L alloy produced by various AM techniques (Reproduced with permission from[Citation230]).

Figure 28. (a–d) Top and side views of the LPBF 18Ni-300 maraging steel produced via laser remelting under pure N2 atmosphere. Dark and white arrows respectively show the white parent powder embedded into the dark grey oxides, and TiN inclusions, (e) Top-view SEM micrograph of the LPBF 18Ni-300 maraging steel that shows heavily cracked inclusions. Single melting under oxygen rich N2 atmosphere was employed,[Citation232] (f) role of Ṫ in Sn-Mn inclusions size in conventionally cast, and LMD SS 316 L (Reproduced with permission from[Citation235]).

Figure 28. (a–d) Top and side views of the LPBF 18Ni-300 maraging steel produced via laser remelting under pure N2 atmosphere. Dark and white arrows respectively show the white parent powder embedded into the dark grey oxides, and TiN inclusions, (e) Top-view SEM micrograph of the LPBF 18Ni-300 maraging steel that shows heavily cracked inclusions. Single melting under oxygen rich N2 atmosphere was employed,[Citation232] (f) role of Ṫ in Sn-Mn inclusions size in conventionally cast, and LMD SS 316 L (Reproduced with permission from[Citation235]).

Figure 29. (a–c) Top and side views of a melt track showing variation in G, R, solidification mode, and grain size across fusion zone (Adapted with permission from[Citation236]), (d) cross-section micrographs of WAAM HSLA showing grain coarsening along the building direction (Q-PF: quasi-polygonal ferrite) (Adapted with permission from[Citation238]).

Figure 29. (a–c) Top and side views of a melt track showing variation in G, R, solidification mode, and grain size across fusion zone (Adapted with permission from[Citation236]), (d) cross-section micrographs of WAAM HSLA showing grain coarsening along the building direction (Q-PF: quasi-polygonal ferrite) (Adapted with permission from[Citation238]).

Figure 30. Effect of scanning speed, scan strategy, and preheating temperature on grain size and other microstructural features of the LPBF X30Mn21 AHSS (Adapted with permission from[Citation239]).

Figure 30. Effect of scanning speed, scan strategy, and preheating temperature on grain size and other microstructural features of the LPBF X30Mn21 AHSS (Adapted with permission from[Citation239]).

Figure 31. The microstructural features of the as-built and heat-treated LPBF 316 L samples (Reproduced with permission from[Citation251]).

Figure 31. The microstructural features of the as-built and heat-treated LPBF 316 L samples (Reproduced with permission from[Citation251]).

Figure 32. Microstructure of the LPBF 316 L at conditions of (a) as-built, (b,c) solution-annealing treatment at 800 °C and 1000 °C, respectively, (d,e) SEM images of the samples solution treated at 800 °C and 1000 °C, respectively, and (f) load-displacement curves plotted via automatic ball indentation (ABI) testing (Reproduced with permission from[Citation252]).

Figure 32. Microstructure of the LPBF 316 L at conditions of (a) as-built, (b,c) solution-annealing treatment at 800 °C and 1000 °C, respectively, (d,e) SEM images of the samples solution treated at 800 °C and 1000 °C, respectively, and (f) load-displacement curves plotted via automatic ball indentation (ABI) testing (Reproduced with permission from[Citation252]).

Figure 33. SEM micrographs of worn surfaces in (a) as-built sample, and (b, c) heat-treated samples tested at wear-test temperatures of 800 °C and 1000 °C, respectively (Reproduced with permission from[Citation252]).

Figure 33. SEM micrographs of worn surfaces in (a) as-built sample, and (b, c) heat-treated samples tested at wear-test temperatures of 800 °C and 1000 °C, respectively (Reproduced with permission from[Citation252]).

Figure 34. Orientation-imaging maps of the as-deposited and heat-treated (HT) GMA-AM 316 L: (a) as-deposited, (b) HT at 1000 °C for 1 h, WQ, (c) HT at 1100 °C for 1 h, WQ, and (d) HT at 1200 °C for 1 h (Reproduced with permission from[Citation253]).

Figure 34. Orientation-imaging maps of the as-deposited and heat-treated (HT) GMA-AM 316 L: (a) as-deposited, (b) HT at 1000 °C for 1 h, WQ, (c) HT at 1100 °C for 1 h, WQ, and (d) HT at 1200 °C for 1 h (Reproduced with permission from[Citation253]).

Figure 35. A comparison between cellular structures in AM 316 L steel: (a) as-built, (b) heat treated at 800 °C for 1 h, (c, d) SEM and bright field TEM images of the sample heat treated at 900 °C, (e,f) EBSD phase map and grain orientation map of the sample heat treated at 1100 °C, and (g) tensile stress–strain curves of the as-built and heat-treated (HT) samples (HT was annealing at 1100 °C for 1 h) (Reproduced with permission from[Citation254]).

Figure 35. A comparison between cellular structures in AM 316 L steel: (a) as-built, (b) heat treated at 800 °C for 1 h, (c, d) SEM and bright field TEM images of the sample heat treated at 900 °C, (e,f) EBSD phase map and grain orientation map of the sample heat treated at 1100 °C, and (g) tensile stress–strain curves of the as-built and heat-treated (HT) samples (HT was annealing at 1100 °C for 1 h) (Reproduced with permission from[Citation254]).

Figure 36. Cross-sectional SEM micrographs of (a) as-built, and (b) heat-treated samples, and (c, d) higher magnifications of (etched by ralph reagent) (Reproduced with permission from[Citation260]).

Figure 36. Cross-sectional SEM micrographs of (a) as-built, and (b) heat-treated samples, and (c, d) higher magnifications of Figure 36(b) (etched by ralph reagent) (Reproduced with permission from[Citation260]).

Figure 37. EBSD orientation maps of (a) as-built, (b) solution-annealed, (c) solution-annealed-aged samples, (d) distribution of grain boundary corresponding to the areas shown in grain boundary maps of the areas shown in (color Code for the grain boundaries are: Green: 2°–15°, red: 15°–50°, and blue: 50°–180°) (Reproduced with permission from[Citation269]).

Figure 37. EBSD orientation maps of (a) as-built, (b) solution-annealed, (c) solution-annealed-aged samples, (d) distribution of grain boundary corresponding to the areas shown in Figure 37(a,e) grain boundary maps of the areas shown in Figure 37(b) (color Code for the grain boundaries are: Green: 2°–15°, red: 15°–50°, and blue: 50°–180°) (Reproduced with permission from[Citation269]).

Figure 38. Fatigue life in the LPBF 17-4 PH samples at various heat treatment and manufacturing conditions (Reproduced with permission from[Citation264]).

Figure 38. Fatigue life in the LPBF 17-4 PH samples at various heat treatment and manufacturing conditions (Reproduced with permission from[Citation264]).

Figure 39. SEM Fractographs of the 15-5 PH samples after impact testing: (a) as-built, (b) solution-annealed, (c,d) H900, and (e,f) H1150 conditions (Reproduced with permission from[Citation271]).

Figure 39. SEM Fractographs of the 15-5 PH samples after impact testing: (a) as-built, (b) solution-annealed, (c,d) H900, and (e,f) H1150 conditions (Reproduced with permission from[Citation271]).

Figure 40. Bright-field TEM micrographs of (a,b) heat-treated LPBF 15-5 PH, (c) heat-treated wrought 15-5 PH, (d) HE-XRD profiles of heat-treated LPBF 15-5 PH samples at different applied strains (in these profiles, austenite and martensite are marked in green and red colors, respectively) (Reproduced with permission from[Citation272]).

Figure 40. Bright-field TEM micrographs of (a,b) heat-treated LPBF 15-5 PH, (c) heat-treated wrought 15-5 PH, (d) HE-XRD profiles of heat-treated LPBF 15-5 PH samples at different applied strains (in these profiles, austenite and martensite are marked in green and red colors, respectively) (Reproduced with permission from[Citation272]).

Figure 41. SEM micrographs of the LPBF CX parts in (a,b) as-built, (c) heat-treated conditions, (d,e) cyclic polarization measurements of heat-treated LPBF CX, and quench-tempered 420 stainless steels, respectively (Reproduced with permission from[Citation276]).

Figure 41. SEM micrographs of the LPBF CX parts in (a,b) as-built, (c) heat-treated conditions, (d,e) cyclic polarization measurements of heat-treated LPBF CX, and quench-tempered 420 stainless steels, respectively (Reproduced with permission from[Citation276]).

Figure 42. (a–d) inverse pole figure maps, (e–h) corresponding phase maps of the as-built and solution annealed samples, and (i) stress-strain curves of the as-built and annealed S31803 DSS samples (Reproduced with permission from[Citation279]).

Figure 42. (a–d) inverse pole figure maps, (e–h) corresponding phase maps of the as-built and solution annealed samples, and (i) stress-strain curves of the as-built and annealed S31803 DSS samples (Reproduced with permission from[Citation279]).

Figure 43. Contour curves of the LPBF MS1 samples mechanically tested (a) Rp0.2: yield strength; (b) Rm: tensile strength; (c) at: total % of elongation; (d) HV: Vickers hardness; and (e) ΔR: planar anisotropy estimated at 1.5% axial strain (Reproduced with permission from[Citation284]).

Figure 43. Contour curves of the LPBF MS1 samples mechanically tested (a) Rp0.2: yield strength; (b) Rm: tensile strength; (c) at: total % of elongation; (d) HV: Vickers hardness; and (e) ΔR: planar anisotropy estimated at 1.5% axial strain (Reproduced with permission from[Citation284]).

Figure 44. Atom probe tomography of the precipitates Ni3Ti (η type) and Fe7Mo6 (μ phase) formed in the LPBF MS1 maraging steel aged for 2 h at 510 °C (Reproduced with permission from[Citation80]).

Figure 44. Atom probe tomography of the precipitates Ni3Ti (η type) and Fe7Mo6 (μ phase) formed in the LPBF MS1 maraging steel aged for 2 h at 510 °C (Reproduced with permission from[Citation80]).

Figure 45. Plots of da/dN vs. ΔK at different (a) heat treatments, and (b) number of overload cycles (Reproduced with permission from[Citation290]).

Figure 45. Plots of da/dN vs. ΔK at different (a) heat treatments, and (b) number of overload cycles (Reproduced with permission from[Citation290]).

Figure 46. SEM micrographs of the samples subjected to different post treatments of (a) SR, (b) SR + HT, and (c) SR + HIP + HT, (d) large porosity in SR sample (Reproduced with permission from[Citation295]).

Figure 46. SEM micrographs of the samples subjected to different post treatments of (a) SR, (b) SR + HT, and (c) SR + HIP + HT, (d) large porosity in SR sample (Reproduced with permission from[Citation295]).

Figure 47. SEM micrographs of LPBF H13 samples after various heat-treatment processes of (a,b) tempering processes at 600 °C and 700 °C, respectively, and (c) austenitizing at temperature of 1040 °C followed by quenching in oil (Reproduced with permission from[Citation298]).

Figure 47. SEM micrographs of LPBF H13 samples after various heat-treatment processes of (a,b) tempering processes at 600 °C and 700 °C, respectively, and (c) austenitizing at temperature of 1040 °C followed by quenching in oil (Reproduced with permission from[Citation298]).

Figure 48. OM images of ODS samples produced at different laser powers of (a,b) 1200 W, (c,d) 1600 W, (e, f) 2000 W, and (g, h) 1200 W – HIP. TEM micrographs of the oxides in ODS alloy produced using various laser powers of (i) 1200 W, (j) 1600 W, (k) 2000 W, and (l) 1200 W – HIP (Reproduced with permission from[Citation305]).

Figure 48. OM images of ODS samples produced at different laser powers of (a,b) 1200 W, (c,d) 1600 W, (e, f) 2000 W, and (g, h) 1200 W – HIP. TEM micrographs of the oxides in ODS alloy produced using various laser powers of (i) 1200 W, (j) 1600 W, (k) 2000 W, and (l) 1200 W – HIP (Reproduced with permission from[Citation305]).

Figure 49. A scheme of cathodic, anodic, passive and transpassive regions used to identify the localized corrosion parameters (Reproduced with permission from[Citation321]).

Figure 49. A scheme of cathodic, anodic, passive and transpassive regions used to identify the localized corrosion parameters (Reproduced with permission from[Citation321]).

Figure 50. Cumulative charge of metastable pit vs. the frequency of the metastable pit (sample 4 has the highest fraction of porosity, while sample 7 has the lowest value. Wrought sample has no porosity) (Reproduced with permission from[Citation343]).

Figure 50. Cumulative charge of metastable pit vs. the frequency of the metastable pit (sample 4 has the highest fraction of porosity, while sample 7 has the lowest value. Wrought sample has no porosity) (Reproduced with permission from[Citation343]).

Figure 51. A scheme of the effect of gas porosity on the pitting mechanism in the LPBF 316 L stainless steel (Reproduced with permission from[Citation344]).

Figure 51. A scheme of the effect of gas porosity on the pitting mechanism in the LPBF 316 L stainless steel (Reproduced with permission from[Citation344]).

Figure 52. Corrosion mechanisms of stainless steels produced via powder metallurgy: (a) initial state with an intact passive layer on the pore’s surface, (b) beginning of the dissolution, (c) further dissolution as active-passive cells are formed, and (d) further progression of dissolution (Reproduced with permission from[Citation345]).

Figure 52. Corrosion mechanisms of stainless steels produced via powder metallurgy: (a) initial state with an intact passive layer on the pore’s surface, (b) beginning of the dissolution, (c) further dissolution as active-passive cells are formed, and (d) further progression of dissolution (Reproduced with permission from[Citation345]).

Figure 53. SEM micrographs of the surfaces in the LPBF 316 L stainless steel immersed into a solution at different times: (a) surface perpendicular to build direction (top sample) after a short time of immersion, (b) the surface parallel to build direction (side sample) after a short time of immersion, (c) the top sample after 20-day immersion, (d) the side sample after 20-day immersion, and (e) EDS maps of an area Pitted (Reproduced with permission from[Citation346]).

Figure 53. SEM micrographs of the surfaces in the LPBF 316 L stainless steel immersed into a solution at different times: (a) surface perpendicular to build direction (top sample) after a short time of immersion, (b) the surface parallel to build direction (side sample) after a short time of immersion, (c) the top sample after 20-day immersion, (d) the side sample after 20-day immersion, and (e) EDS maps of an area Pitted (Reproduced with permission from[Citation346]).

Figure 54. SEM micrographs of the inclusions before Exposing to 6 wt. % ferric chloride solution (left column) and after that (right column). the chemical composition of the inclusions is shown schematically in the middle column in which green color represents the manganese chromite, blue color is manganese silicate, yellow color is silicon oxide, and red color is MnS (Reproduced with permission from[Citation360]).

Figure 54. SEM micrographs of the inclusions before Exposing to 6 wt. % ferric chloride solution (left column) and after that (right column). the chemical composition of the inclusions is shown schematically in the middle column in which green color represents the manganese chromite, blue color is manganese silicate, yellow color is silicon oxide, and red color is MnS (Reproduced with permission from[Citation360]).

Figure 55. FIB cross sections that show the typical corrosion morphology at grain boundaries in (a–c) commercial, and (d–g) LPBF 316 L stainless steels. The high-magnified subfigures (b,c,f,g) are corresponding to the white rectangles marked in subfigures (a,d,e) (Reproduced with permission from[Citation361]).

Figure 55. FIB cross sections that show the typical corrosion morphology at grain boundaries in (a–c) commercial, and (d–g) LPBF 316 L stainless steels. The high-magnified subfigures (b,c,f,g) are corresponding to the white rectangles marked in subfigures (a,d,e) (Reproduced with permission from[Citation361]).

Figure 56. The EDS and SKPFM measurements of the LPBF 15-5 PH parts: (a) microstructure of the at sample, (b) higher magnification of the area marked in (a), (c) the SKPFM profile corresponding to b), and (d) the potential-distance data in different areas (Reproduced with permission from[Citation317]).

Figure 56. The EDS and SKPFM measurements of the LPBF 15-5 PH parts: (a) microstructure of the at sample, (b) higher magnification of the area marked in (a), (c) the SKPFM profile corresponding to b), and (d) the potential-distance data in different areas (Reproduced with permission from[Citation317]).

Figure 57. (a) SEM image of MPBs in the LPBF 316 L sample, and (b) a pit formed on a MPB (Reproduced with permission from[Citation369,Citation373]).

Figure 57. (a) SEM image of MPBs in the LPBF 316 L sample, and (b) a pit formed on a MPB (Reproduced with permission from[Citation369,Citation373]).

Figure 58. (a) Distinguishable areas of “A” and “B” at each side of MPBs, and (b) BSE image with higher magnification of cellular subgrains in the LPBF 316 L sample (Reproduced with permission from[Citation350]).

Figure 58. (a) Distinguishable areas of “A” and “B” at each side of MPBs, and (b) BSE image with higher magnification of cellular subgrains in the LPBF 316 L sample (Reproduced with permission from[Citation350]).

Figure 59. SEM images of corroded surfaces in (a–c) as-built sample, (d–f) laser polished sample, immersed into 0.4 Mol/L HCl solution (various magnifications), (g,h) corrosion mechanism in the as-built and laser polished samples, respectively (Reproduced with permission from[Citation311]).

Figure 59. SEM images of corroded surfaces in (a–c) as-built sample, (d–f) laser polished sample, immersed into 0.4 Mol/L HCl solution (various magnifications), (g,h) corrosion mechanism in the as-built and laser polished samples, respectively (Reproduced with permission from[Citation311]).

Figure 60. (a,b) BSE images of the columnar cells’ boundaries and the melt pool traces through the y-z plane, (c) corresponding Crystal orientation map along the z-axis, (d–f) {001} pole figures of the points C, D, and E in (b), respectively, where the arrows indicate the direction of cell orientation (green: ±45°-oriented cells, red: vertically-oriented cells), and (g) corresponding potentiodynamic polarization curves of the CLMs and the wrought counterpart (reference plate) immersed into 0.9 wt.% NaCl solution at 37 °C (Reproduced with permission from[Citation313]).

Figure 60. (a,b) BSE images of the columnar cells’ boundaries and the melt pool traces through the y-z plane, (c) corresponding Crystal orientation map along the z-axis, (d–f) {001} pole figures of the points C, D, and E in (b), respectively, where the arrows indicate the direction of cell orientation (green: ±45°-oriented cells, red: vertically-oriented cells), and (g) corresponding potentiodynamic polarization curves of the CLMs and the wrought counterpart (reference plate) immersed into 0.9 wt.% NaCl solution at 37 °C (Reproduced with permission from[Citation313]).

Figure 61. IPF maps of 316 L stainless steels at conditions of (a) Quenched, (b) SLM-120 W, (c) SLM-150 W, (d) SLM-195 W, (e) SLM-220 W, and (f) corresponding grain size distribution maps (Reproduced with permission from[Citation394]).

Figure 61. IPF maps of 316 L stainless steels at conditions of (a) Quenched, (b) SLM-120 W, (c) SLM-150 W, (d) SLM-195 W, (e) SLM-220 W, and (f) corresponding grain size distribution maps (Reproduced with permission from[Citation394]).

Figure 62. Mott-Schottky plot at frequency of 1 kHz measured in the cathodic direction. The area between two dotted lines shows a difference between the slope of plots through the passive region. Inset shows a difference in n-type region for different samples (Reproduced with permission from[Citation396]).

Figure 62. Mott-Schottky plot at frequency of 1 kHz measured in the cathodic direction. The area between two dotted lines shows a difference between the slope of plots through the passive region. Inset shows a difference in n-type region for different samples (Reproduced with permission from[Citation396]).

Figure 63. (a–c) Optical micrographs of three planes “A,” “B,” and “C” in the LPBF 316 L stainless steel along the building direction, (d–f) corresponding IPF maps of planes “A,” “B,” and “C,” respectively, (g) the grain distribution map, and (h) potentiodynamic polarization plots of planes “A,” “B” and “C.” (Reproduced with permission from[Citation397]).

Figure 63. (a–c) Optical micrographs of three planes “A,” “B,” and “C” in the LPBF 316 L stainless steel along the building direction, (d–f) corresponding IPF maps of planes “A,” “B,” and “C,” respectively, (g) the grain distribution map, and (h) potentiodynamic polarization plots of planes “A,” “B” and “C.” (Reproduced with permission from[Citation397]).

Figure 64. (a,b) Misorientation maps of grain boundary in the LPBF CX sample obtained from the top and side planes, respectively, (c) corresponding grain size distribution map, and (d) the CPP curves of the top and side planes (Reproduced with permission from[Citation78]).

Figure 64. (a,b) Misorientation maps of grain boundary in the LPBF CX sample obtained from the top and side planes, respectively, (c) corresponding grain size distribution map, and (d) the CPP curves of the top and side planes (Reproduced with permission from[Citation78]).

Figure 65. (a) Scheme of a PEMFC with a complex structure,[Citation398,Citation399] (b,c) bright-field TEM of the surface in the wrought 316 L stainless steel after hydrogen charging at 50 mA/cm2 for 4 h, (d) the electronic diffraction pattern of the area marked in (b), (e,f) bright-field TEM of the surface in the LPBF 316 L stainless steel after hydrogen charging at 50 mA/cm2 for 4 h (Reproduced with permission from[Citation398]).

Figure 65. (a) Scheme of a PEMFC with a complex structure,[Citation398,Citation399] (b,c) bright-field TEM of the surface in the wrought 316 L stainless steel after hydrogen charging at 50 mA/cm2 for 4 h, (d) the electronic diffraction pattern of the area marked in (b), (e,f) bright-field TEM of the surface in the LPBF 316 L stainless steel after hydrogen charging at 50 mA/cm2 for 4 h (Reproduced with permission from[Citation398]).

Figure 66. (a,b) Inverse pole figures of the LPBF stainless steel and particle-reinforced composite produced by the LPBF process, respectively, (c,d) the grain size distribution of the LPBF stainless steel and the LPBF composite, respectively, and (e) corresponding CPP curves of both Materials (Reproduced with permission from[Citation400]).

Figure 66. (a,b) Inverse pole figures of the LPBF stainless steel and particle-reinforced composite produced by the LPBF process, respectively, (c,d) the grain size distribution of the LPBF stainless steel and the LPBF composite, respectively, and (e) corresponding CPP curves of both Materials (Reproduced with permission from[Citation400]).

Figure 67. (a) Grain boundary map of a SCC in the LPBF 316 L steel, and (b) SCC crack growth of un-recrystallized (stress-relieved at 650 °C), partially-recrystallized (heat-treated at 955 °C), and fully-recrystallized (HIP + SA) LPBF components (all tests are through the x-z orientation with no primary cold working) (Reproduced with permission from[Citation402]).

Figure 67. (a) Grain boundary map of a SCC in the LPBF 316 L steel, and (b) SCC crack growth of un-recrystallized (stress-relieved at 650 °C), partially-recrystallized (heat-treated at 955 °C), and fully-recrystallized (HIP + SA) LPBF components (all tests are through the x-z orientation with no primary cold working) (Reproduced with permission from[Citation402]).

Figure 68. Bright-field transmission electron microscopy (TEM) images of (a) LPBF and (b) TM 304 L stainless steel after irradiation. Different depths of the cross-section are shown for the LPBF sample (a1–a3) and the TM sample (b1–b3). (c) Illustration of frank dislocation loops observed in LPBF and TM 304 L stainless steel following irradiation at the peak damage region. The size distribution of dislocation loops in LPBF and TM 304 L stainless steel, irradiated with 2 MeV protons to a dosage of 0.18 dpa at 360 °C, is presented (Reproduced with permission from[Citation403]).

Figure 68. Bright-field transmission electron microscopy (TEM) images of (a) LPBF and (b) TM 304 L stainless steel after irradiation. Different depths of the cross-section are shown for the LPBF sample (a1–a3) and the TM sample (b1–b3). (c) Illustration of frank dislocation loops observed in LPBF and TM 304 L stainless steel following irradiation at the peak damage region. The size distribution of dislocation loops in LPBF and TM 304 L stainless steel, irradiated with 2 MeV protons to a dosage of 0.18 dpa at 360 °C, is presented (Reproduced with permission from[Citation403]).

Figure 69. Surface morphology of the oxide scale formed on (a1,a2) irradiated LPBF samples, (b1,b2) irradiated TM samples, (c1,c2) unirradiated LPBF samples, and (d1,d2) unirradiated TM samples after corrosion for 454 h in simulated primary water of a pressurized water reactor (PWR) at 320 °C (Reproduced with permission from[Citation403]).

Figure 69. Surface morphology of the oxide scale formed on (a1,a2) irradiated LPBF samples, (b1,b2) irradiated TM samples, (c1,c2) unirradiated LPBF samples, and (d1,d2) unirradiated TM samples after corrosion for 454 h in simulated primary water of a pressurized water reactor (PWR) at 320 °C (Reproduced with permission from[Citation403]).