You are on page 1of 141

INVESTIGATION OF STRUCTURE-PROPERTY EFFECTS ON NANOINDENTATION

AND SMALL-SCALE MECHANICAL TESTING OF IRRADIATED ADDITIVELY

MANUFACTURED STAINLESS STEELS

Mohammad Jashim Uddin, B.S., M.S.

Dissertation Prepared for the Degree of

DOCTOR OF PHILOSOPHY

UNIVERSITY OF NORTH TEXAS

August 2022

APPROVED:

Hector R. Siller, Major Professor


Reza Mirshams, Co- Major Professor
Taeyul Theo Choi, Committee Member
Sheldon Shi, Committee Member
Bibhudutta Rout, Committee Member
Herman Shen, Chair of the Department of
Mechanical Engineering
Shengli Fu, Interim Dean of the College of
Engineering
Victor Prybutok, Dean of the Toulouse Graduate
School
Uddin, Mohammad Jashim. Investigation of Structure-Property Effects on

Nanoindentation and Small-Scale Mechanical Testing of Irradiated Additively Manufactured

Stainless Steels. Doctor of Philosophy (Mechanical and Energy Engineering), August 2022, 128

pp., 21 tables, 44 figures, 198 numbered references.

Additively manufactured (AM) 316L and 17-4PH stainless steel parts, concretely made by

laser powder bed fusion (L-PBF), are characterized and micro-mechanical properties of those

steels are analyzed. This study also explored and extended to proton irradiation and small-scale

mechanical testing of those materials, to investigate how irradiation affects microstructural

evolution and thus mechanical properties at the surface level, which could be detrimental in the

long term in nuclear applications. In-depth anisotropy analysis of L-PBF 316L stainless steel parts

with the variations of volumetric energy density, a combined study of nanoindentation with EBSD

(electron backscatter diffraction) mapping is shown to be an alternative methodology for enriching

qualification protocols. Each grain with a different crystallographic orientation was mapped

successfully by proper indentation properties. <122> and <111> oriented grains displayed higher

than average indentation modulus and hardness whereas, <001>, <101>, and <210> oriented

grains were found to be weaker in terms of indentation properties.

Based on an extensive nanoindentation study, L-PBF 17-4 PH stainless steels are found to

be very sensitive to high load rates and irradiation further escalates that sensitivity, especially after

a 0.25 s-1 strain rate. 3D porosity measurement via X-ray microscope ensures L-PBF stainless steel

parts are of more than 99.7% density and could be promising for many industrial applications.

High percentages of increment of nanohardness, maximum theoretical shear strength, and yield

strength were observed due to proton irradiation of 5 um damage depth on the surface of 17-4 PH

steel parts.
Small-scale mechanical testing of irradiated AM nuclear stainless steels such as 17-4 PH

was carried out and investigated by micro-compression of FIB fabricated pillars of different sizes

of diameter. Irradiated 17-4 PH materials have never been investigated by this kind of testing

procedure to asses the stress-strain characteristics of micro-scale volumes and to explore the

structure-property relationship. Both as-built and irradiated AM 17-4 PH micropillars exhibited

step-ups in the early stage of load-displacement curves with a varying number of slip bands

intermittently formed throughout the pillar volume while compressed by the uniaxial load. As for

the radiation-damaged zone, micropillars displayed lesser slip bands compared to as-built parts as

irradiation damage creates an obstacle to dislocations movement and hence hardening. It requires

higher loads to initiate plastic deformation as dislocation must overcome irradiation-induced

obstacles for the slip to occur and localization of strain without increasing the load for a certain

amount of time during the test. Proton irradiation effects on the compressive mechanical properties

of AM 17-4 PH stainless steel parts depending on the volumetric energy density (VED) used

during the parts’ fabrication process. On as-built parts, compressive yield strength varied from

107.27 MPa to 150.70 MPa and it was in the range of 133.43 MPa to 244.57 MPa under irradiated

conditions. All 2 μm pillars were fabricated as their height falls within the radiation damage depth

of 5 μm. It was expected to generate the highest yield strength and tensile strength due to the

radiation hardening effect as discussed earlier. Yield and tensile strength were found to be the

highest as expected as of 244.57 MPa and 375.08 MPa in irradiated 17-4 PH sample 1 (VED =

54.76 J/mm3). Samples with lower VED exhibited better micro-mechanical compressive responses

than higher VED AM 17-4 PH parts in both as-built and irradiated conditions.
Copyright 2022

by

Mohammad Jashim Uddin

ii
ACKNOWLEDGEMENTS

From the very start to the very end of my Ph.D. program in Mechanical Engineering at the

University of North Texas, I am so blessed to meet and work with so many outstanding professors,

colleagues, and friends. I would like to thank Dr. Hector Siller for being my thesis major professor

as well as to give me a chance to work in his research group where I received his continuous

guidance and mentorship throughout my research and dissertation. Additionally, I am so grateful

for his countless hours of reflecting, reading, and providing ideas for the resolution of problems

throughout the entire process and without his guidance, I would not be able to come this far. I

would also like to thank Dr. Reza Mirshams for being my co-major professor and to provide

continuous support and advice in every step of my study.

I would like to give special thanks to Dr. Taeyul Choi and Dr. Sheldon Shi, for being part

of my dissertation committee and to provide their valuable input and expertise throughout the

process. I convey special thanks to Dr. Teresa Golden and Dr. Bibhudutta Raut for their

collaboration, support, and helpful discussions.

I dedicate this dissertation to my family specially my parents Md. Majibar Rahman and

Momtaz Begum who encouraged and supported me. I also dedicate this dissertation to my lovely

wife Sabiha Tul Jannat for her unconditional inspiration, love, and support throughout the process.

iii
TABLE OF CONTENTS

Page

ACKNOWLEDGEMENTS ........................................................................................................... iii

LIST OF TABLES ......................................................................................................................... vi

LIST OF FIGURES ..................................................................................................................... viii

CHAPTER 1. INTRODUCTION ................................................................................................... 1


1.1 Additive Manufacturing .......................................................................................... 1
1.2 Radiation-Resistant Alloys Manufacturing by Laser Powder Bed Fusion (L-PBF)
................................................................................................................................. 8
1.3 Laser Powder Bed Fusion of 316L and 17-4 PH Stainless Steels ........................ 10
1.4 Thesis Objectives .................................................................................................. 13
1.5 Scientific Justification of the Research ................................................................. 17
1.6 Hypothesis............................................................................................................. 23
1.7 What’s Inside this Dissertation ............................................................................. 24

CHAPTER 2. MATERIALS AND EXPERIMENTAL PROCEDURES .................................... 25


2.1 Methodology ......................................................................................................... 25
2.2 Materials ............................................................................................................... 30
2.3 Process Parameters................................................................................................ 31
2.4 Sample Preparation ............................................................................................... 33
2.5 Proton Irradiation .................................................................................................. 34
2.6 Nanoindentation .................................................................................................... 36
2.7 Strain-rate Sensitivity of 17-4 PH Stainless Steels ............................................... 37
2.8 Small-scale Mechanical Testing ........................................................................... 40
2.9 X-ray Diffraction (XRD) and Electron Backscattered Diffraction (EBSD)
Microstructural Characterization .......................................................................... 41
2.10 Porosity Analysis .................................................................................................. 43
2.11 Fabrication of Micropillars by Focused Ion Beam (FIB) ..................................... 44

CHAPTER 3. RESULTS AND DISCUSSIONS.......................................................................... 45


3.1 Nanoindentation and Electron Backscatter Diffraction Mapping in Laser Powder
Bed Fusion of Stainless Steel 316L ...................................................................... 45
3.1.1 Energy Dispersive X-ray Spectroscopy (EDS) ......................................... 46

iv
3.1.2 Nanoindentation Results ........................................................................... 46
3.1.3 Grain Morphology and Microstructures ................................................... 49
3.1.4 Electron Backscattered Diffraction (EBSD) Maps ................................... 54
3.1.5 Validation of the Indentation Results ....................................................... 61
3.1.6 Misorientation Angle and Grain Size Estimation ..................................... 62
3.1.7 Conclusions on Nanoindentation, Anisotropy, and Mechanical Properties
................................................................................................................... 64
3.2 Effects of Proton Irradiation on Nanoindentation Strain-Rate Sensitivity and
Microstructural Properties in L-PBF 17-4 PH Stainless Steels ............................ 66
3.2.1 X-ray Diffraction (XRD) Analysis ........................................................... 67
3.2.2 Micro-Mechanical Properties by Nanoindentation ................................... 68
3.2.3 EBSD Microstructural Characterization ................................................... 73
3.2.4 Correlation of Martensite Fractions to Process Parameters of L-PBF 17-4
PH As-Built and Irradiated Stainless Steel Parts ...................................... 78
3.2.5 Porosity Analysis ...................................................................................... 81
3.2.6 Strain-Rate Sensitivity (SRS) and Effect of Proton Irradiation on the
Hardness .................................................................................................... 83
3.2.7 Maximum Shear and Yield Strength Approximation ............................... 86
3.2.8 Conclusions of Irradiation Effects on Mechanical Properties and
Microstructure ........................................................................................... 90
3.3 Small-Scale Mechanical Testing of Proton Irradiated L-PBF 17-4 PH Stainless
Steels ..................................................................................................................... 91
3.3.1 Micro-Compression Testing, Observation of Micropillar Fractures, and
Nanoindentation Response........................................................................ 91
3.3.2 Conclusions of Irradiation Effects on Small-Scale Mechanical Testing of
AM 17-4 PH Stainless Steel ................................................................... 102

CHAPTER 4. CONCLUSIONS AND FUTURE WORKS........................................................ 104

REFERENCES ........................................................................................................................... 110

v
LIST OF TABLES

Page

Table 1: A literature review on nanoindentation study of similar AM metallic alloys (MC =


microstructural characterization, OIP = Orientation based indentation properties, SSE =
indentation size effect, SSMT = small-scale mechanical test, SRS = strain-rate sensitivity, B =
Berkovich, S = Spherical, FT = Flat tip)....................................................................................... 18

Table 2: Literature review on small-scale mechanical testing of relevant metallic alloys (LP =
laser power, HD = hatch distance, LT = layer thickness, SS = scanning speed, SS = strain rate,
YS = yield strength, UTS = ultimate tensile strength) .................................................................. 21

Table 3: Elemental composition of LPW AISI 316L stainless steel powder particles ................. 30

Table 4: EOS elemental composition of L-PBF 17-4 PH powder particles and printed parts ..... 31

Table 5: L-PBF AISI 316L stainless steel manufacturing process parameters ............................ 32

Table 6: L-PBF process parameters for the fabrication of 17-4 PH stainless steel parts.............. 33

Table 7: XRM Scan parameters .................................................................................................... 44

Table 8: Elemental quantitative analysis of AISI 316L powder by energy dispersive spectroscopy
(EDS) ............................................................................................................................................ 46

Table 9: Analysis of variance (ANOVA) of the influence of volumetric energy density (VED) on
the indentation modulus (significance level, α =0.05) .................................................................. 47

Table 10: Analysis of variance (ANOVA) of the influence of volumetric energy density (VED)
on hardness values by nanoindentation (significance level, α =0.05) .......................................... 48

Table 11: Statistical One-way ANOVA of hardness data ............................................................ 70

Table 12: Statistical One-way ANOVA of indentation modulus data.......................................... 72

Table 13: Quantitative analysis of EBSD scans of as-built and irradiated L-PBF 17-4 PH
stainless steel parts ........................................................................................................................ 79

Table 14: Percentage of martensite phase according to experiment and Eq. 8 for L-PBF 17-4 PH
stainless steels from open literatures............................................................................................. 80

Table 15: Percentage of porosity determined by using different methods for L-PBF 17-4 PH
stainless steels extracted from open literature. ............................................................................. 82

Table 16: Strain-rate of L-PBF 17-4 PH stainless steel using different approaches .................... 85

Table 17: Maximum shear strength of as-built and irradiated L-PBF 17-4 PH steel parts at 0.01s-1
strain rate with loads associated with first displacement bursts. .................................................. 87

vi
Table 18: Yield strength of as-built and irradiated L-PBF 17-4 PH steel parts at 0.01s-1 strain
rate................................................................................................................................................. 88

Table 19: Ultimate tensile strength and yield strength determined from engineering stress-strain
curves of all the pillars in as-built and irradiated conditions. ..................................................... 100

Table 20: Comparison with other AM 17-4 PH stainless steel micropillar compression testing
literature (LP = laser power (Watts), HD = hatch distance (μm), LT = layer thickness (μm), SS =
scanning speed (mm/s), UTS = Ultimate tensile strength) ......................................................... 101

Table 21: Summary of major mechanical properties calculated from indentation and small-scale
mechanical testing (for 3 μm) methods on as-built and irradiated 17-4 PH stainless steels....... 107

vii
LIST OF FIGURES

Page

Figure 1: Schematic of laser powder bed fusion (L-PBF) process and parameters. ....................... 6

Figure 2: Summary of dissertation investigations from AM to small-scale mechanical testing.


....................................................................................................................................................... 14

Figure 3: Research map with the outline of this study.................................................................. 15

Figure 4: List of mechanical properties that are studied and analyzed in this research................ 23

Figure 5: Methodology of the proposal ........................................................................................ 26

Figure 6: Layer-by-layer micro-mechanical properties by small-scale mechanical testing ......... 29

Figure 7: Schematics of process parameters and scanning strategy for the fabrication of L-PBF
AISI 316L parts............................................................................................................................. 32

Figure 8: SRIM simulated (a) distribution of 1 MeV proton in the 17-4PH steel, (b) Vacancies
and displacements per atom (dpa) as a function of depth for a proton irradiation fluence of
1.0×1019 ions/cm2. ......................................................................................................................... 35

Figure 9: Nanoindentation experimental set-up for strain-rate sensitivity study on as-built and
irradiated L-PBF 17-4 PH stainless steel parts. ............................................................................ 37

Figure 10: Example of a load-displacement (P-h) curve of as-built L-PBF 17-4 PH part (VED =
61.11 J/mm3) at 0.1 s-1 strain-rate. ................................................................................................ 37

Figure 11: Micro-compression experimental set up (a) 10 um diameter flat indenter tip, (b) array
of micropillars in irradiated condition, and (c) array of micropillars in as-built condition of AM
17-4 PH stainless steel parts. ........................................................................................................ 41

Figure 12: Regression analysis (ANOVA) of nanoindentation hardness and modulus data with
various VED; (a) Interval plot of indentation modulus and (b) interval plot of hardness with a
confidence level of 95%................................................................................................................ 47

Figure 13: SEM images showing build direction (BD) and grain morphology of L-PBF AISI
316L parts with different volumetric energy density (VED); (a) and (b) VED = 45.24 J/mm3, (c)
and (d) VED = 50 J/mm3, (e) and (f) VED = 56.67 J/mm3. ......................................................... 49

Figure 14: SEM images showing planar and cellular dendritic grains with build direction (BD) of
L-PBF AISI 316L parts with volumetric energy density (VED)=50 J/mm3; (a) area of SEM scan,
(b) cellular dendritic structures with preferential 45o angle growth with pool boundary, and (c)
planar grains. ................................................................................................................................. 52

viii
Figure 15: Microstructures of L-PBF AISI 316L parts with different process parameters; (a)
VED = 50 J/mm3, (b) VED = 56.67 J/mm3. ................................................................................. 53

Figure 16: EBSD maps of L-PBF AISI 316L parts showing one of three nanoindentation regions
with the variation of volumetric energy density; (a) EBSD and IPF color map, (b) texture PF and
(c) texture IPF of parts (VED = 45.24 J/mm3), and (d) EBSD map, (e) texture, and (f) texture IPF
of parts (VED = 50 J/mm3). .......................................................................................................... 55

Figure 17: EBSD maps of L-PBF AISI 316L parts showing one of three nanoindentation regions
with the variation of volumetric energy density; (a) EBSD map, (b) texture PF and (c) texture
IPF of parts (VED = 56.67 J/mm3), and followed by (d) EBSD map and IPF color map, (e)
texture PF and (f) texture IPF of parts (VED = 66.67 J/mm3). ..................................................... 56

Figure 18: EBSD map of L-PBF AISI 316L parts showing one of three nanoindentation regions
with the variation of volumetric energy density; (a) EBSD map, (b) texture PF and (c) texture
IPF of parts (VED = 83.33 J/mm3), and (d) original image quality map of part with VED = 50
J/mm3, and (e) IPF color map. ...................................................................................................... 57

Figure 19: EBSD maps of all L-PBF AISI 316L parts marked as rectangles in Fig. 16, Fig. 17
and Fig. 18. All indents are labeled with indentation modulus (IM) and hardness (H) generated
by the nanoindentation experiment; (a) VED = 45.24 J/mm3, (b) VED = 50 J/mm3, (c) VED =
56.67 J/mm3, (d) VED = 66.67 J/mm3, (e) VED = 83.33 J/mm3, and (f) original Image Quality
(IQ) of part with VED = 66.67 J/mm3 and IPF color map............................................................ 58

Figure 20: (a) EBSD maps of conventionally manufactured AISI 316L stainless steel, (b) texture
PF, (c) texture IPF, (d) rectangular area with indent’s hardness (H) and indentation modulus
(IM), (e) original IQ, and, (f) IPF color map. ............................................................................... 60

Figure 21: Indentation modulus (IM) and hardness (H) within same crystallographic orientation
grains from EBSD maps of AISI 316L parts with (a) VED = 56.67 J/mm3 and (b) VED = 66.67
J/mm3. ........................................................................................................................................... 61

Figure 22: Effect of grain size and misorientation angle on indentation modulus and hardness
based on volumetric energy density (VED). ................................................................................. 63

Figure 23: X-ray diffraction measurements of L-PBF 17-4 PH as-built (AB) and irradiated (IR)
stainless steel parts. ....................................................................................................................... 67

Figure 24: Nanoindentation hardness plot against different strain rates (histogram) for as-built
and irradiated L-PBF (54.76 J/mm3 and 61.11 J/mm3) 17-4 PH steel parts. ................................ 69

Figure 25: Nanoindentation modulus plot against different strain rates (mean plot and standard
deviation as error) for (a) as-built and (b) irradiated L-PBF 54.76 J/mm3 and 61.11 J/mm3 17-4
PH steel parts. ............................................................................................................................... 71

Figure 26: EBSD scan of nanoindentation area (at 0.5 s-1 strain rate) of as-built L-PBF 17-4 PH
stainless steel part with VED = 54.76 J/mm3; (a) IPF map in the (X–Y) plane perpendicular to

ix
the build direction, (b) phase map, (c) texture pole figure, (d) inverse texture pole figure, and (e)
IPF color maps of phases. ............................................................................................................. 75

Figure 27: EBSD scan of nanoindentation area (at 1 s-1 strain rate) of irradiated L-PBF 17-4 PH
stainless steel part with VED = 54.76 J/mm3; (a) IPF map in the (X–Y) plane perpendicular to
the build direction, (b) phase map, (c) texture pole figure, (d) inverse texture pole figure, and (e)
IPF color maps of phases. ............................................................................................................. 76

Figure 28: EBSD scan of nanoindentation area (at 0.5 s-1 strain rate) of as-built L-PBF 17-4 PH
stainless steel part with VED = 61.11 J/mm3; (a) IPF map in the (X–Y) plane perpendicular to
the build direction, (b) phase map, (c) texture pole figure, (d) inverse texture pole figure, and (e)
IPF color maps of phases. ............................................................................................................. 76

Figure 29: EBSD scan of nanoindentation area (at 1 s-1 strain rate) of irradiated L-PBF 17-4 PH
stainless steel part with VED = 61.11 J/mm3; (a) IPF map in the (X–Y) plane perpendicular to
the build direction, (b) phase map, (c) texture pole figure, (d) inverse texture pole figure, and (e)
IPF color maps of phases. ............................................................................................................. 77

Figure 30: X-ray microscope scan of L-PBF 17-4 PH stainless steel including irradiated zone
(volumetric energy density of 54.76 J/mm3); (a) 3D scan volume (reconstructed), (b) single slice
(ImageJ), and (c) 3D reconstruction of all 2401 slices showing pores or voids within the scanned
volume........................................................................................................................................... 81

Figure 31: X-ray microscope scan of L-PBF 17-4 PH stainless steel including irradiated zone
(volumetric energy density of 61.11 J/mm3); (a) 3D scan volume (reconstructed), (b) single slice
(ImageJ), and (c) 3D reconstruction of all 2401 slices showing pores or voids within the scanned
volume........................................................................................................................................... 82

Figure 32: Load-displacement curves (P-h) for as-built L-PBF 17-4 PH stainless steel in different
strain rates, (a) VED = 54.76 J/mm3 and (b) VED = 61.11 J/mm3. .............................................. 84

Figure 33: Load-displacement curves (P-h) for irradiated L-PBF 17-4 PH stainless steel in
different strain rates, (a) VED = 54.76 J/mm3 and (b) VED = 61.11 J/mm3. ............................... 84

Figure 34: Strain-rate sensitivity (SRS) determined by plotting the log-log scale of average
hardness and strain rate values and plots are drawn considering the density of the parts
(normalized); (a) SRS in as-built and (b) SRS in irradiated L-PBF 17-4 PH steels parts. ........... 85

Figure 35: (a) First displacement pop-ins in load-displacement curves for as-built and irradiated
L-PBF 17-4 PH stainless steel parts and (b) effect of irradiation on maximum shear strength at
0.1 s-1 strain rate nanoindentation tests. ........................................................................................ 87

Figure 36: Hardness-displacement curves of as-built and irradiated 54.76 J/mm3 17-4 PH
stainless steel parts in different strain rates................................................................................... 88

Figure 37: Summary of micro-mechanical properties, quantitative analysis of EBSD maps, and
maximum shear strength of as-built (AB) and irradiated (IR) L-PBF 17-4 PH stainless steel parts
(based on EBSD scans performed on 1 s-1 strain rate nanoindentation areas).............................. 89

x
Figure 38: Initial (top row) and after micro-compression (bottom row) micropillars of as-built
AM 17-4 PH stainless steels with the volumetric energy density of 54.76 J/mm3 (sample 1); (a1)
and (a2) 3 μm pillar, and (b1) and (b2) 5 μm pillar. ..................................................................... 92

Figure 39: Initial (top row) and after micro-compression (bottom row) micropillars of as-built
AM 17-4 PH stainless steels with the volumetric energy density of 61.11 J/mm3 (sample 2); (a1)
and (a2) 3 μm pillar, and (b1) and (b2) 5 μm pillar. ..................................................................... 93

Figure 40: Initial (top row) and after micro-compression (bottom row) micropillars of irradiated
AM 17-4 PH stainless steels with the volumetric energy density of 54.76 J/mm3 (sample 1); (a1)
and (a2) 2 μm pillar, (b1) and (b2) 3 μm pillar , and (c1) and (c2) 5 μm pillar. ........................... 93

Figure 41: Initial (top row) and after micro-compression (bottom row) micropillars of irradiated
AM 17-4 PH stainless steels with the volumetric energy density of 61.11 J/mm3 (sample 2); (a1)
and (a2) 2 μm pillar, (b1) and (b2) 3 μm pillar, and (c1) and (c2) 5 μm pillar. ............................ 94

Figure 42: Load-displacement curves of as-built and irradiated AM 17-4 PH stainless steel parts
with two different volumetric energy densities (sample 1 = VED of 54.76 J/mm3 and sample 2 =
VED of 61.11 J/mm3). .................................................................................................................. 96

Figure 43: Engineering stress-strain curves of sample 1 (VED = 54.76 J/mm3) with different
sizes of micropillars of L-PBF 17-4 PH stainless steels in (a) as-built condition, (b) irradiation
condition, and (c) comparison between as-built and irradiation conditions. ................................ 97

Figure 44: Engineering stress-strain curves of sample 2 (VED of 61.11 J/mm3) with different
sizes of micropillars of L-PBF 17-4 PH stainless steels in (a) as-built condition, (b) irradiation
condition., and (c) comparison between as-built and irradiation conditions. ............................... 98

xi
CHAPTER 1

INTRODUCTION

1.1 Additive Manufacturing*

Additive manufacturing (AM) techniques are the processes to fabricate a wide range of

structures and complex geometries from 3-D solid models. These processes are usually fabricating

successive layers of materials that are formed on top of each other and built on support structures.

This technology was developed by Charles Hull in 1986 as a process known as stereolithography

(SLA), which later evolved into processes such as powder bed fusion, fused deposition modeling

(FDM), inkjet printing, and contour crafting (CC), and so on. An additive manufacturing process

involves numerous methods, materials, and equipment [1]. The growing consensus of adapting the

additive manufacturing techniques over traditional techniques is attributed to several advantages

including fabrication of complex geometry with high precision, maximum material savings,

flexibility in design, and personal customization. A wide range of materials is becoming available

to be fabricated using AM processes as traditional processes are more time-consuming, difficult,

and costly. An optimized pattern of additive manufacturing parts is important to control flaw

sensitivity and anisotropic behavior [2]. Also, changes in the printing environment influence the

quality of finished products [3]. AM is capable of fabricating parts of various sizes from the micro-

to macro-scale.

The revolution of additive manufacturing technology happened recently a few decades ago.

Some advantages over traditional manufacturing processes established the additive manufacturing

process as a unique field of fabrication of hundreds of materials and the numbers are increasing

* The latter portion of this section is reproduced from E. Ramirez-Cedillo, M.J. Uddin, J.A. Sandoval-Robles, R.A.
Mirshams, L. Ruiz-Huerta, C.A. Rodriguez, H.R. Siller, Process planning of L-PBF of AISI 316L for improving
surface quality and relating part integrity with microstructural characteristics, Surf. Coatings Technol. 396 (2020)
125956. https://doi.org/10.1016/j.surfcoat.2020.125956, with permission from Elsevier.

1
day by day. Some of the advantages are rapid prototyping, improved speed of production, ability

to create complex parts without any difficulty, cost-effective even for lower production volume,

high strength fabricated alloys parts, improved carbon footprint i.e. less waste in the production

process, and less energy usage, the requirement of less space to operate. Many industries such as

aerospace (engine and turbine parts), automotive (exhausts, mirrors, seat belt hook, headlight

heatsink, tire rims), clothing (eye wear, sports shoes, helmet, rackets), construction (entire

building), electronics (circuits, parts, mobile phones, antennas, sensors), medical (prosthetics,

body parts as in reconstructive surgery), technology (laptop and tablet cases, motherboards,

robotics) are implementing additive manufacturing processes and that number is increasing rapidly

[4–11]. From its roots in prototypes, metal AM has rapidly become a core technology that has a

practical application in several different industries. Often described as belonging to a group known

as “disruptive technologies” metal AM has the capacity to completely revolutionize how things

are made. It’s not currently used in high volume, mass productions, but instead offers the most

value where precision engineering is vital or where customization and flexibility are key. Fuel

nozzles in the aerospace industry are a great example of how this technology is used in practice.

The new designs created with the use of metal AM techniques are five times more durable and

25% lighter.

Different additive manufacturing processes are described in the standard ISO-ASTM

52900. Among them, the process of interest in this dissertation is Laser powder bed fusion

technologies (L-PBF), which have been extensively used in recent years since machines are more

affordable and the process is in continuous improvement [12–14]. With the advances in the

development of more efficient systems and lasers, and in situ monitoring, the productivity of L-

PBF machines has increased. The Wohlers report presented that revenue from AM metal machines

2
grew 29.9% in 2018 to an estimated 947.8 US million dollars [15]. Early adoption in the academy

and industry has increased the research and development of final products significantly. In the

academic field, according to the SCOPUS database, from 1975 to 2019, a total of 5889 documents

were published about L-PBF. In the last years, 2016 (624 documents), 2017 (865 documents), and

2018 (1226 documents) have been published, increasing the scientific and industrial knowledge

around this technology, where China, Germany, and the US are the countries with more

contributions. Another growing market in AM is the production of the powder. Commercially

available alloys commonly used for casting/powder metallurgy have been introduced to L-PBF

with successful results. Significant growth of 41.9% in 2018 of the revenue of different types of

commercial alloys in a total of 260.2 US million dollars. However, some critical issues have been

limiting the advantages of AM with these available materials, such as balling of the material during

the process, defects produced due to the thermal history of the material (solidification cracking/hot

tearing), and others [16]. Hence, the development of new alloys that can be weldable avoiding the

cracking tendency, has contributed to the exponential growth in the research on this topic. Even

though it is a recent technology, and there is a curve of knowledge for early adopters, with the

right guidelines, workflow, documentation, and characterization, optimized parameters can be

obtained. Thereby machines can be running production as some industries have reported. AM

processes provide unique challenges not experienced in traditional manufacturing methods. Due

to the nature of many AM processes, it is difficult to control the thermal history of a component,

resulting in heterogeneous microstructures and defect distributions [17]. This can be a result of the

size of the component, variations in thickness, or the number of components. As microstructure

plays a significant role in the material behavior in extreme environments, whether it be mechanical

3
properties or other environmental effects, it is vital that this process be better understood and

controlled.

The advantages of AM processes have emerged and continue to prevail through ongoing

research efforts to understand and eliminate constraints that inhibited the use of this technology.

Design tools to assess life-cycle costs i.e., AM-oriented computer-aided design (CAD) systems

with more user-friendly and advanced simulation capabilities are some of the key aspects that need

to be realized [18]. A distinguished advantage of AM techniques is mass customization i.e., the

production of a series of personalized goods such that each product can be different while

maintaining a low price due to mass production. These processes are devoid of the added cost due

to mold making and tooling for a customized product. Therefore, mass production of a number of

identical parts can be as cost-effective as the same number of different personalized goods. The

change between different designs is straightforward with negligible added cost and no need for

special preparation. AM processes also have the potential for mass production of complex

geometries such as lattice structures, where the application of traditional methods of manufacturing

such as casting is not straightforward and require further time-consuming tooling and post-

processing. However, improvements in the fabrication speed and cost reduction must be resolved

through the improvement of machine design. Also, the high costs and time-consumption of the

AM processes remain to be major hurdles that inhibit mass production [19].

Nuclear Regulatory Commission is inspiring scientific communities to perform

investigations to the better understanding of the behavior of irradiated additively manufactured

(AM) materials, like austenitic stainless steels that are used in the Transformational Challenge

Reactor Program [20,21] and Generation IV nuclear reactors. Presently, there are significant

efforts in several fields for thermo-mechanical processing to treat the intrinsic change of

4
mechanical properties and microstructural changes in additively manufactured components prone

to degradation, such as the evaluation of stress-corrosion cracking susceptibility, porosity,

crystallography, thermal diffusivity, anisotropy, and chemical heterogeneity, considering that AM

processes could bring profits to the fabrication of complex components in nuclear reactors and

vessels [22–26]. To mitigate the chance of materials failure or damage during service, AM

radiation-resistant engineering alloys must withstand levels of temperature and radiation existing

in operational conditions; hence the study of microstructure and micro-mechanical properties upon

irradiation, is mandatory toward effective materials qualification.

The reliance on witness specimens and printed coupons to determine the results of process

parameters are time-consuming and costly, and typically covers low-temperature applications

(witness specimens are scarcely been used for creep or high-temperature cyclic testing). There is

a large need for better AM models that more accurately reflect actual thermal histories, as well as

the resulting variations in microstructures, to reduce the need for experimental optimization of

processing parameters. The testing of AM material so far has largely, if not entirely, been

performed with uniaxial testing. There is a need to expand this work to examine the effects of

multiaxial loading [27,28], particularly for AM components that have exhibited anisotropic

mechanical strengths [12]. While necessary for all testing, it is expected that multiaxial testing

would be of particular importance when considering elevated temperature failure mechanisms. For

energy applications, however, there is a need to understand and qualify the behavior of AM

materials at elevated temperatures. Long-term and time-dependent failure mechanisms such as

creep and creep-fatigue damage control the design of high-temperature structural components [29].

These failure mechanisms are connected to time-dependent properties, such as creep, thermal

aging, cyclic softening, etc., and not time-independent properties, such as yield stress and ultimate

5
tensile stress. There is a significant gap in understanding the effects of AM properties on the

corrosion of a material. For many advanced energy systems, the environments have become

increasingly complex (molten salt, liquid sodium, super critical water, and super critical CO2). For

the use of AM within these and other environments, there needs to be an increased understanding

of the corrosion behavior of AM materials.

Figure 1: Schematic of laser powder bed fusion (L-PBF) process and parameters.

Understanding the basics of the processing parameters and variables that are inherent to

the process is fundamental to generate workflow or process guidelines. Firstly, processing

parameters have different names and nomenclatures. Figure 1 represents a summary of the most

important and well-known processing variables and their equivalences for different machines.

Laser Power (P), Layer thickness (LT), laser scanning speed (vs), and hatch distance (HD) are

considered to be the most important parameters due to their impact on the integrity of the parts

(Figure 1) [30]. Commercial machines have different laser systems that can be programmed in

different forms: pulse width modulated (PWM) and continuous (CW). When a laser is working

6
either in PWM, it can provide processing parameters related to the pulse frequency, duration, and

peak power [31]. Since L-PBF is an AM process, where layers (n) are added continuously, process

parameters such as the initial and rotation angle of layers can have an impact on the distribution

of stresses in all directions [32].

In order to relate parameters with the mechanical, geometric, surface, and density

properties, the volumetric energy density has been widely adopted as an indicator to select the

optimum range of parameters. Eq. (1) is the representation of VED in terms of the previously

presented most influential parameters P (watts), vs (mm/s), LT (μm), and HD (μm).


𝑃𝑃
𝑉𝑉𝑉𝑉𝑉𝑉 = 𝑣𝑣𝑠𝑠 ×𝐿𝐿𝐿𝐿×𝐻𝐻𝐻𝐻
(Eq. 1)

Some authors discussed if VED is a good approximation for parameter synthesis, drawing

several concerns due to the lack of information about the material consideration in the VED

equation [33], the lack of understanding of the main characteristics of the laser [34]; and the

uncertainty in reproducibility and feasibility. Other authors made profound conclusions about why

VED is a thermodynamic quantity without capturing complex physical phenomena such as the

Marangoni flow, hydrodynamic instabilities, and the recoil pressure, which have a direct impact

on the consistency of the tracks [35]. Still, VED is widely used to summarize process parameters,

in addition to characterization techniques such as porosity analysis, density measurements, and

tensile and hardness tests.

In summary, this dissertation is a package of a new strain-rate sensitivity model,

correlations between process-driven microstructures and micro-mechanical properties, and small-

scale mechanical testing of as-built and proton irradiated stainless steel materials fabricated by the

additive manufacturing process. The package, wholeheartedly, would be crucial for the overall

7
integrity of alternative materials qualification protocol for nuclear intricate material parts’

applications.

1.2 Radiation-Resistant Alloys Manufacturing by Laser Powder Bed Fusion (L-PBF)

Nuclear energy is looking to AM, along with other advanced manufacturing techniques, to

provide design flexibility, reduce costs, and shorten timelines for the production of components

that make up a nuclear power plant [36]. While other methods for nuclear energy components,

such as powder metallurgy (specifically hot isostatic pressing), are more developed, AM is a fairly

new area of interest for nuclear energy. AM has the benefit of allowing for constant process

monitoring of each layer of material as it is manufactured, allowing for improved manufacturing

control and the collection of detailed information about the properties of the final component.

Another benefit of AM is the ability to transition materials more seamlessly (gradual transitions

and narrow fusion zones with minimized heat-affected zones) than traditional joining (brazing,

welding, or soldering) or to create functionally graded compositions, or even fully 3D patterned

phase, grain size, and composition-distributions aimed at optimizing the microstructure-properties

relationship [37]. Material transitions are particularly common in nuclear plants, as there are large

variations in pressures, temperatures, and radiation doses, resulting in the need to employ a variety

of materials. AM has been demonstrated to create graded compositions of many types of materials,

such as Inconel to steel, copper to steel, titanium alloys to steel, as well as titanium alloys to carbide

materials [37–41]. Modeling and simulation efforts have helped progress this work and increase

the feasibility of functionally graded compositions, and even 3D printed micro-composites [42,43].

Most reactor internals is not pressure-retaining components and do not fall within the

restrictions of the ASME BPV Code, therefore, AM work has progressed more significantly in the

area of actual application in nuclear power plants. There is interest in AM for reactor core internals,

8
including manufacturing of fuel, cladding, control elements, etc. Westinghouse was one of the first

to consider installing actual AM components into a reactor, choosing to install a thimble plugging

device made of AM 316L steel. Significant work is being performed within research and test

reactors to better understand the performance of AM components. Oak Ridge National Laboratory

has explored the use of AM to manufacture control elements for the High-Flux Isotope Reactor

(HFIR) [44,45] and Idaho National Laboratory has used AM capsules for tests inside the Advanced

Test Reactor. Similar to the structural components, the benefits from AM include the ability to

create complex designs, rapidly prototype, and transition between dissimilar materials. For some

exotic materials, such as those used in nuclear fuels, there are additional benefits, such as using

the AM energy source not only for manufacturing the structure of the fuel but also for reducing or

modifying the fuel composition, such as in the conversion of UF4 to U3Si2, an accident tolerant

fuel [46]. In this case, there are additional benefits, such as using less equipment for fuel fabrication

from cradle to fuel product, as well as less raw material input and minimizing the facility footprint.

It also has the potential to improve the economics, as traditional manufacturing routes would

include an additional step of converting the UF4 to uranium metal before the formation of U3Si2.

AM has found applications in other areas related to nuclear energy outside of the plant itself. One

such application is the use of AM in the reprocessing of spent fuel. Spent nuclear fuel is a

significant challenge for the nuclear energy industry. One method for overcoming this problem is

to recycle the spent fuel and use it again in reactors. AM has been demonstrated to manufacture

complex, fluid devices with internal channels, as well as a full array of 1.25-cm diameter rotor

centrifugal contactors for small-scale demonstrations of separation techniques aiming to remove

minor actinides from used nuclear fuel [47]. In this case, AM was used to simplify the fabrication

processes and reduce the level of needed materials and human effort.

9
1.3 Laser Powder Bed Fusion of 316L and 17-4 PH Stainless Steels*

316L stainless steel alloy is a member of the austenitic steels that were developed more

than three decades ago for fast-breeder reactor applications within the EU countries. AISI 316L

grade has both good weldability and machinability and in combination with high corrosion

resistance, it can be used in a wide range of applications. Based on its excellent properties the

SS316L steel was selected as the main structure retaining material for the Vacuum Vessel and

other systems in the International Thermo-Nuclear Experimental Reactor (ITER) [48]. Researchers

and engineers have performed many experiments to establish the properties of 316L stainless

steels and tested various fabrication and joining methods used for the manufacturing of designed

components for nuclear fusion applications [49–51]. The aim was to confirm the steel behavior

under tough and very critical working environments. One major problem encountered has been the

complex geometry of the ITER in-vessel components, like the first wall (FW) panel, which needs

assembly of different sub-components to achieve the final complex part. This process is both

difficult and costly and the assembled larger component has a higher risk to fail at the joints. An

example of the presently favored technique is to make separate sub-components and use hot

isostatic pressing (HIP) to press them together. This procedure, however, involves many steps;

machining of the sub-plates, assembling of these sub-plates, pre-welding the sub-plates by laser or

electron beam, HIP at elevated temperature, post-machining, and final heat treatments. With such

a complex procedure a flaw formed in any of the initial steps might generate failure in the final

* This section is reproduced from a combination of 2 sources, both used with permission from Elseiver: (1) M.J.
Uddin, E. Ramirez-Cedillo, R.A. Mirshams, H.R. Siller, Nanoindentation and electron backscatter diffraction
mapping in laser powder bed fusion of stainless steel 316L, Mater. Charact. 174 (2021) 111047,
https://doi.org/10.1016/j.matchar.2021.111047 and (2) M.J. Uddin, H.R. Siller, R.A. Mirshams, T.A. Byers, B.
Rout, Effects of proton irradiation on nanoindentation strain-rate sensitivity and microstructural properties in L-PBF
17–4 PH stainless steels, Mater. Sci. Eng. A. 837 (2022) 142719. https://doi.org/10.1016/j.msea.2022.142719.

10
component. Therefore, new manufacturing methods are highly demanded for future fusion reactor

construction components.

The first material of interest in this research, austenitic stainless steel with denomination

AISI 316L, is widely employed in marine engineering, potable water systems, food preparation

equipment, pharmaceutical manufacturing, and medical implants. Particularly, the applications of

additively manufactured parts via laser powder bed fusion (L-PBF) are constantly expanding in

medical, nuclear, chemical, food, petrochemical, architecture, and automotive fields very rapidly,

because of its exceptional corrosion resistance and superior biocompatibility [52–61]. Several

limitations are preventing the full exploitation of L-PBF of AISI 316L; some of them are due to

the anisotropy and the heterogeneity in microstructure and mechanical properties as reported in

recent studies [62–66]. Pores, impurities, and unmelted powder particles are responsible for this

behavior, but there is also the influence of the nature of the microstructure: columnar and cellular

grain microstructures are formed together, a complex cyclic thermal history that includes

directional heat extraction, repeated melting, and rapid solidification of powder particles on the

same layer [67]. Additionally, the grain size and shape variation in different fabrication planes

parallel and perpendicular to the build direction also contribute to the anisotropy in mechanical

properties [63]. Another group of researchers verified that AISI 316L parts fabricated by L-PBF

exhibit lower resistance to deformation along the building direction, and anisotropy in

microstructure and mechanical properties which can be eliminated after recrystallization [62]. It is

also reported that post-processing of AM parts, though currently very costly, is mandatory to

obtain optimum dynamic properties like fatigue resistance in comparison to conventionally

produced parts [68]. One key aspect to consider for obtaining enhanced mechanical properties is

the control of grain aspect ratio and orientation since it was found in this study that material with

11
a lower grain aspect ratio (equiaxed shape) showed significantly increased mechanical properties

(YS=583 MPa, TS=685 MPa) compared to parts with a higher grain aspect ratio (columnar shape)

(YS=521MPa, TS=617MPa). To sum up, the influencing factors in the anisotropy and

heterogeneity in the microstructure and mechanical properties involve crystallographic texture,

lack-of-fusion defects, grain morphology, phase transformation, heterogeneous recrystallization,

layer building, and microstructural coarsening.

Different research studies addressed the relationship between anisotropy and process

parameters in the L-PBF of AISI 316L. According to recent scientific work, the most influential

parameter is the laser power, which affects part’s density, residual stresses, and tensile mechanical

properties [68]. Regarding the impact in terms of microstructure, a group of researchers found

that laser power of 380 W resulted in a <001> crystallographic texture and lower strains with larger

grain size and higher laser power (950 W) resulted in a <011> crystallographic texture and higher

strains with more dislocation density, which eventually make the latter stronger and tougher [69].

In parallel, high laser power of 1 kW with narrow Gaussian power distribution resulted in the

formation of long columnar grains along the building direction with variation in mechanical

properties and microstructures, whereas, low laser power of 400 W resulted in a finer sub-grain

and randomly oriented columnar grains with an increase in hardness and tensile strength [70].

From these findings, it can be stated that power by itself does not describe properly a clear

relationship and that the use of energy density could be more appropriate for understanding the

microstructural evolution and its connection with the mechanical response.

The second nuclear alloy in this research study is 17-4 precipitation hardening (PH) steels

which are considered difficult to be manufactured via conventional machining processes due to

their high strength and high hardness properties. PH steels are well suited to be fabricated by laser

12
powder bed fusion (L-PBF) technique due to better weldability and austenitic/martensitic

microstructure [71]. Nuclear, aerospace, marine, naval, and chemical industries use PH steels for

components and intricate parts. One of the most popular and widely used PH steel alloys is 17-4

PH stainless steel, and in the specific case of L-PBF 17-4 PH steels, the findings demonstrate their

corrosion resistance, a relatively high mechanical strength, good resistance against radiation, and

favorable fatigue behavior [72–74]. However, microstructural analysis and mechanical properties

have been reported and discussed in limited literature. Additively manufactured alloys are suitable

for applications under severe radiation environments and the application usually narrows the

materials qualification processes. Materials used under those harsh conditions require a standard

level of mechanical strength and stability, especially at elevated temperatures along with

corrosion-prone surroundings [75]. Recently, a group of researchers investigated the neutron-

irradiated conventional 304L austenitic stainless steels at a microstructural scale but did not study

the effect of irradiation on micro-mechanical properties such as nanohardness and indentation

modulus [76]. Therefore, to mitigate the chance of materials failure or damage, L-PBF radiation-

resistant engineering alloys must withstand successfully that level of temperature and radiation;

hence the study of microstructure and micro-mechanical properties of L-PBF 17-4 PH stainless

steels upon ion irradiation, seems to be mandatory toward effective materials selection in nuclear

industries.

1.4 Thesis Objectives

Laser powder bed fusion (L-PBF) processes are associated with pores, impurities, and

unmelted powder particles that are generally accountable for inferior performance of parts in the

field of applications, but there is also the influence of the nature of the microstructures as columnar

and cellular grain microstructures are formed together at complex cyclic thermal history, which

13
includes directional heat extraction, repeated melting, and rapid solidification of powder particles

on the same layer. As a whole, the manipulating factors in the microstructure disturbing

mechanical response comprise of crystallographic texture, lack-of-fusion defects, grain

morphology, phase transformation, heterogeneous recrystallization, layer building, and

microstructural coarsening among others. The different qualification methods in AM materials,

schematized in Figure 2, rely on the use of standardized specimens at different orientations on the

powder bed, and with different configurations of process parameters like laser power, laser

scanning speeds, layer thicknesses, hatching distances, and others.

Figure 2: Summary of dissertation investigations from AM to small-scale mechanical testing.

These tests can capture most of the mechanical characteristics but are unable to discover intrinsic

material properties of irradiated materials at low surface depths. The use of nanoindentation

(Figure 2) becomes an alternative and is complementary to elucidating the effect of proton and

neutron irradiation on a change in mechanical properties, that ultimately would be detrimental in

14
the long term. With the use of witness specimens replicating those geometries used in nuclear

reactors, the use of nanoindentation mapping, tensile testing, creep-fatigue testing, and

proton/neutron irradiation on AM materials, the development of microstructurally informed

models could be created as digital threads to increase reliability which is the main objective of this

dissertation.

Figure 3: Research map with the outline of this study

Besides, the research on the fundamental science of alternative qualification processes for

additively manufactured materials at the surface level is limited. Concretely, studies of

nanoindentation response, small-scale mechanical testing i.e. micropillar micro-compression,

strain-rate sensitivity (SRS), and microstructural and micro-mechanical characterizations on AM

nuclear alloys such as 316L and 17-4 PH stainless steels are very limited. This dissertation

emphasizes those scarcely explored research fields at the sub-micron level on L-PBF radiation-

resistant stainless steels and claims to have a multi-scale approach. The objective is to experiment

15
micro-mechanical behavior on L-PBF alloys in terms of micro-mechanical properties such as

indentation hardness, compressive strength, shear strength, and other critical properties needed to

be qualified in parts that will be subjected to extreme operational conditions. Additionally, to study

the sensitivity of nanoindentation behavior to loading rate in sub-micron level in which load-

displacement curves in different strain rates will provide insights into whether the radiation-

resistant alloys are susceptible to an increase in loading rate. Furthermore, the implementation of

focused ion beam (FIB) milling to create flat nanoindenter tip and micropillars is proposed as a

method to create samples to assess the size effects at the micro and small scale but particularized

to AM irradiated materials [77–82]. The research map is shown in Figure 3.

Studying strain-rate sensitivity (SRS) would also be helpful to determine the maximum

shear stress from the nanoindentation curve which is an estimate of the theoretical shear strength

of L-PBF radiation-resistant stainless steels. A mechanistic relationship is thereby possible

between size effects and micro-mechanical properties. Micro-compression tests with a range of

micropillars (diameter of 2 μm to 5 μm) are proposed for minimizing the size effects on the

accuracy of nanoindentation and to study how to determine yield strength in terms of a power-law

relationship. A large array of nanoindentation batch experiments would validate the hardness

measurements. On the other hand, radiation effects on the L-PBF parts in terms of hardness,

microstructures, and micro-compressive strength via micropillars both on the surface and along

the damage profile depth would be beneficial to characterize them in depth. Besides, process

parameters such as laser power, hatch distance, and layer thickness play an important role in

determining the micro-mechanical properties and crystallographic microstructures. There are

available resources of research on mechanical properties by nanoindentation of L-PBF radiation-

resistant stainless steels. But, the fields of exploring size effects, minimizing those effects, layer-

16
by-layer micro-mechanical properties along with the height of parts [83], the study of strain-rate

sensitivity by increasing the loading rate on those alloys, are very limited, and studying those fields

would be a significant contribution to the scientific community.

1.5 Scientific Justification of the Research*

Firstly, in-depth analysis of the microstructure and mechanical properties of additively

manufactured metals are less reported especially for radiation-resistant stainless steels. To realize

the full design potential that additively manufactured alloys can offer, more precisely for load-

bearing structural components, it is very important to provide a thorough understanding of the

anisotropic and heterogeneous microstructure and mechanical properties that often occur within

L-PBF parts during the complex solidification process. This dissertation outlines 316L and 17-4

PH stainless steel parts and reviews literature on the anisotropy and heterogeneity in microstructure

and mechanical. It can be highlighted that the contributing factors to the anisotropy and

heterogeneity within metal L-PBF parts were either their unique microstructural features or

manufacturing deficiencies. The correlation will be drawn by implementing quantitative analysis

tools provided based on EBSD (Electron backscatter diffraction) patterns between the grain

orientations and indent’s modulus of elasticity and hardness values determined by

nanoindentation.

Indentation techniques such as nanoindentation have a long practice in characterizing

mechanical property changes due to exposure in a nuclear environment and allow access to more

data from an instrumented indentation test. The utilization of nanoindentation has received great

* This section is reproduced from M.J. Uddin, H.R. Siller, R.A. Mirshams, T.A. Byers, B. Rout, Effects of proton
irradiation on nanoindentation strain-rate sensitivity and microstructural properties in L-PBF 17–4 PH stainless
steels, Mater. Sci. Eng. A. 837 (2022) 142719. https://doi.org/10.1016/j.msea.2022.142719, with permission from
Elsevier.

17
attention in ion beam irradiated materials [84]. Application of indentation in cross-section (large

array of indents in an area) to assess the ion-beam irradiation-related property changes has been

growing rapidly. Continuous stiffness mode indentation (where a low amplitude oscillation,

typically, 2 nm at 50 Hz, is overlaid on the load curve), is increasingly used because of very

shallow ion-beam irradiation, and more complex material properties such as strain rate sensitivity

and creep properties can be measured. A comparative literature review on nanoindentation

properties of various additively manufactured metallic alloys is shown in Table 1:

Table 1: A literature review on nanoindentation study of similar AM metallic alloys (MC =


microstructural characterization, OIP = Orientation based indentation properties, SSE = indentation
size effect, SSMT = small-scale mechanical test, SRS = strain-rate sensitivity, B = Berkovich, S =
Spherical, FT = Flat tip)
Tip used and
Mechanical Properties
Author Alloy (H=Nanohardness, IM= OIP SSE SSMT Irradiation SRS
Indentation Modulus in
GPa)
JS Weaver
SS 316 L B, S, H=3.59
[85]
X Wang
Ti-6Al-4V B, IM=112-116, H =6
[86]
B, H=2.6, H= 3.1
X Sun [87] SS 316 L (Irradiated Helium √ √
bubble interface)
A
Ti-6Al-4V, B, IM=112-144, H=2.9-
Reichardt
SS 304L 5.1
[88]
NT
B, H=1.82±0.01,
Aboulkhair AlSi10Mg
H=1.52±0.02 (HT)
[89]
H Wang B, H=3.58-4.70,
Inconel 718 √
[90] IM=160-200
B
B, FT, H=3.27±0.25,
AlMangour SS 17-4 PH √
H=3.87-4.07 (HT)
[91]
Ali
B, H=2.0±0.19,
Behroozfar Cu alloy √
IM=128.2±10.9
[92]
B, H=2.92-3.89 (variable
D. Gu [93] Ti alloy
scan speed)
W
B, H=2.8±0.5,
Żórawski Ti alloy
IM=98.8±12.9
[94]

18
Tip used and
Mechanical Properties
Author Alloy (H=Nanohardness, IM= OIP SSE SSMT Irradiation SRS
Indentation Modulus in
GPa)
Nicola M.
AlSi10Mg B, H=2.25±0.02
Everitt [95]
M Munther B, H=2.98-3.65,
Inconel 718
[96] IM=155-218
J. S.
Weaver Inconel 625 B, H=~5.5 √
[83]
R Erick, M
B, H= 3.93Y±1.19, IM=
J Uddin SS 316 L √
191.05±22.92
[12]
B, H=3.55±0.15,
SS 316 L,
M J Uddin IM=184.85±11.93,
Conventional √ √
[97] Conv: H=3.60±0.05,
316L
IM=189.15±5.88
B, H=4.02±1.25,
IM=170±70.89 (as-
M J Uddin built), H=7.47±3.15,
SS 17-4 PH √ √ √
[14] IM=196.15±70.352
(Irradiated) @variable
strain rates (0.01-1 s-1)
B, FT, At 0.01 s-1:
H=3.28±0.14,
M J Uddin, IM=185.32±5.27 (as-
SS 17-4 PH √ √ √ √
2022 built), H=6.46±0.15,
IM=224.18±6.82
(Irradiated)

Secondly, the term ‘small-scale materials’ testing relates to mechanical test specimens with

dimensions that require focused ion beam (FIB) to fabricate micropillars, a nanoindenter with a

flat indenter tip to perform the micro-compression test, and lastly, SEM or FIB to observe the

fracture type. Typically, the size scale is in the order of tens of nanometers to tens of micrometers,

therefore covering the lower end of the mesoscale spectrum. Small-scale mechanical testing

techniques have a long history in nuclear material research due to the limitations posed by nuclear

facilities [98]. The limited space in reactors and the issue that metal structures are often radioactive,

in addition to the increasing need to obtain mechanical properties from ion beam irradiated

samples, require small specimen mechanical testing. Since the beginning of the nuclear age, the

19
question of material degradation in a nuclear environment being influenced by radiation has always

been a key issue. The reliability of materials in nuclear environments is a big issue, in civil as well

as defense-related applications. Today, for all nuclear reactor concepts proposed or in service, it is

the material issues that pose the greatest challenge. If no material is available or qualified to be

used in a nuclear environment, it is impossible to realize future concepts with greater demands on

material performance. Therefore, it is vital to assess the mechanical properties, microstructural

changes, and the status of materials at specific locations in a component such as an end cap welds,

fuel/clad interfaces, or rather small parts such as the spacer grits and all these parts go under very

high dose of irradiation. Unfortunately, achieving these high doses requires extremely long

irradiation times in reactors and naturally, there are only very few materials irradiated to such high

doses. In fact, in this dissertation L-PBF materials have been irradiated to doses little above 20 dpa

with a limited penetration depth and therefore drastically limits the amount of material available

to be investigated. Overall, the higher penetration depth on the L-PBF parts and bigger area take a

very long time to be irradiated. This limited penetration depth of the ions, which is one of the

drawbacks of utilizing ion beam irradiations, makes small-scale material investigations essential

and is one of the main driving forces to enhance and deploy techniques such as nanoindentation.

In addition, oftentimes the amount of available testing material is limited due to reactor space

limitations. Statistical certainty of the data is another issue leading to the fact that a smaller sample

volume in small-scale mechanical testing is advantageous to obtain more data points on a given

volume of material. A comparative study on small-scale mechanical testing on similar AM alloys

to 316L and 17-4 PH stainless steels is in Table 2.

All irradiation procedures benefit from small-scale material testing techniques regardless

if using reactor neutron or being the ion-beam source. The necessity to link testing with bulk

20
properties is a big concern in the field to extract bulk engineering data. Therefore, the development

of small-scale mechanical testing in combination with the microstructural investigation is of great

interest to the nuclear material community for both, material development and monitoring

applications, respectively. One of the main focuses of this dissertation regarding irradiation is to

evaluate yield stresses, since the more interesting parameters such as total or uniform elongation,

and work hardening rate are more difficult to capture. It has to be noted that irradiation does cause

a noticeable difference in failure mode, where a sudden localized failure of the micropillars could

be observed.

Table 2: Literature review on small-scale mechanical testing of relevant metallic alloys (LP = laser
power, HD = hatch distance, LT = layer thickness, SS = scanning speed, SS = strain rate, YS = yield
strength, UTS = ultimate tensile strength)

AM LP, HD, PD, UTS, Indentation


Author AR SR, s-1 YS, MPa
alloys LT, SS μm MPa properties

B
As-built 200, 110, 688.52±36.
AlMangour --- ---
17-4 PH 40, 750 31
[99]
As-built 4 1:3 0.0001
B
17-4 PH 200, 110, 1130 ±
AlMangour 914±12.73 ---
shot 40, 750 17.68
[91]
peened
As-built
197-451 --- ---
316L 400, 250,
C Shiau,
LT [x], 5 1:2 0.001
2022 [100]
Irradiated 12.7
476-588 --- ---
316L
Sinle
400, HD crytal:
X Wang, As-built
[x], LT 8 1:2 0.0031 408±18, --- ---
2021[101] 316L
[x], 17 Bi-crystal:
502
As-built 107.27- 213.58 -

17-4 PH 220-230, 150.70 286.40
M J Uddin, 1: 2-
120, 50, 2 to 5 0.01
2022 2.5
Irradiated 600-700 133.43- 212.75-

17-4 PH 244.57 375.08

21
Cross-section (at a certain area) indentation enables a more precise way of assessing

hardness versus dose. The indentation experiment on irradiated as well as the unirradiated area

within the same field of indents allows a direct comparison between an unirradiated and irradiated

L-PBF 316L and 17-4 PH stainless steel parts. It is known that irradiated materials lead to different

behavior and irradiation damage increases the strength of a polycrystalline material due to the

presence of small but strong and densely spaced radiation defects (a particular region on the

irradiated area), creating a new small internal material length scale [98]. However, once the sample

size is getting small compared to the length scales represented by grain structure or radiation

damage, they are no longer the limiting factors but rather the stress to create or propagate a

dislocation takes over, returning the material behavior to the unirradiated (quasi single crystal)

material. It is expected that the hardness values generated by nanoindentation would exhibit an

increment over the irradiated area compared to as-built area. The microstructural changes are also

expected on the surface area of irradiation as well as beneath the surface.

Thirdly, strain-rate sensitivities are to be determined by choosing high loading or test rate.

High-resolution property mapping and detailed statistical studies from a large array of indents by

nanoindentation help to confirm the qualification process aimed for nuclear applications of

additively manufactured 316L and 17-4 PH stainless steels. The quality and accuracy of results

largely depend on the quality, calibration, right newer indenter tip, and adequacy of the models

used for data evaluation as well as a well-known example is the increase of accuracy in the

determination of hardness and elastic modulus, which could be attained by calculating the contact

stiffness via a power-law approximation of unloading curve (Oliver & Pharr, 1992) instead of the

originally suggested linear approximation [102]. It is generally accepted that the values of SRS are

strongly size-dependent: for face-centered-cubic (FCC) metals, SRS normally increases with

22
decreasing grain sizes, whereas the opposite holds for body-center-cubic (BCC) structures.

1.6 Hypothesis

Structure-property relationship of additively manufactured (AM) alloys could affect the

mechanical properties of nuclear alloys used in nuclear applications. Nanoindentation and small-

scale mechanical tests could provide a comprehensive set of mechanical properties (as shown in

Figure 4) and evaluation methods for qualification purposes, and could help to accelerate AM

product reliability and process integrity.

Figure 4: List of mechanical properties that are studied and analyzed in this research

The following assumptions could be derived from this hypothesis:

• A modification of the process parameters of laser powder bed fusion in terms of the
use of different energy densities could induce a change in microstructural behavior and
orientation that would affect mechanical properties at the surface level.

• The use of small-scale mechanical testing techniques can afford the possibility of
measuring constitutive materials properties such as compressive strength, yield
strength, maximum shear strength, and studying intrinsic size effects in terms of
deformation behavior.

• Strain-rate sensitivity study of L-PBF radiation-resistant alloys such as 316L and 17-4
PH stainless steel could assist in characterizing the deformation mechanics of those
alloys before and after the irradiation.

23
• The use of small-scale mechanical test techniques in the characterization of L-PBF
radiation-resistant alloys could improve the accuracy of the evaluation of surface
mechanics.

The set of experimentation techniques could clarify and help to understand in a better way:

 Compressive yield strength with the correlation with indentation hardness and modulus
of elasticity.

 Irradiation affects the L-PBF materials (microstructure, micro-mechanical properties,


strength) and changes in deformation mechanics in the damaged surface and sub-
surface up to 5 µm depth.

1.7 What’s Inside this Dissertation

This dissertation includes an introduction of key areas of the research, experimental

procedures of material’s microstructural and mechanical properties’ characteristics, results and

discussion of each experimental section in detail with experimental data, images, and analysis of

data, and conclusion with probable future works that can be accomplished, and numbered

references of open journals or books or dissertations that are used in those studies.

24
CHAPTER 2

MATERIALS AND EXPERIMENTAL PROCEDURES

2.1 Methodology*

L-PBF processes are utilized to fabricate nuclear alloys such as 316L and 17-4 PH stainless

steels. After cutting by electrical discharge machining (EDM), the parts are prepared for material

characterization beginning with nanoindentation and proton irradiation. A 10 by 10 indents array

has been selected to increase the validity of the results. Five different strain rates have been chosen

for measuring SRM to study material’s deformation behavior with increasing loading rate. The

SRS would be calculated by plotting the log scale of both hardness values and different strain rates.

The same setup is being used after the irradiation process. Indented parts would then be observed

under FEI Quanta ESEM to study grain morphology, energy dispersive spectroscopy (EDS) for

elemental composition analysis, and X-ray diffraction patterns (XRD) are being achieved to

confirm the phases present on the parts by using Rigaku Ultima III. The indentation areas would

then be inspected and electron backscattered diffraction (EBSD) patterns are acquired by using

FEI Nova NanoSEM to study the microstructures, grain size distribution, grains crystallographic

orientations, and texture intensity. The overall methodology has been illustrated in Figure 5.

The parts with different process parameters would be placed on a pin mount in FIB (FEI

NOVA FIB/SEM) to fabricate micropillars for micro-compression testing. A very small current

would be used to eliminate the possibility of FIB damage on the micron-sized pillars. Pillars of

size ranging from 2 to 5 μm with an aspect ratio of 2- 2.5 would be milled very carefully and later

* Sections 2.1 to 2.4 are reproduced from a combination of 2 sources, both used with permission from Elseiver: (1)
M.J. Uddin, E. Ramirez-Cedillo, R.A. Mirshams, H.R. Siller, Nanoindentation and electron backscatter diffraction
mapping in laser powder bed fusion of stainless steel 316L, Mater. Charact. 174 (2021) 111047,
https://doi.org/10.1016/j.matchar.2021.111047 and (2) M.J. Uddin, H.R. Siller, R.A. Mirshams, T.A. Byers, B.
Rout, Effects of proton irradiation on nanoindentation strain-rate sensitivity and microstructural properties in L-PBF
17–4 PH stainless steels, Mater. Sci. Eng. A. 837 (2022) 142719. https://doi.org/10.1016/j.msea.2022.142719.

25
on those micropillars would be compressed by using a 10 μm flat indenter using MTS nanoindenter

and stress-strain curves such as shown in Figure 6 are being expected. The L-PBF stainless steel

parts are being proton irradiated up to 5 μm depth and on an approximate area of 3.5 mm X 4 mm.

Micropillars will be fabricated in the irradiated area with before mentioned parameters via FIB to

compare the mechanical properties with those micropillars in the unirradiated area. Layer-by-layer

mechanical properties and layer strength are possible to determine as depicted in Figure 5.

Figure 5: Methodology of the proposal

Additive manufacturing processes based on the local fusion of a powder bed (L-PBF) are

a valid alternative to conventional technologies and a growing number of industrial sectors are

currently relying on these processes for the production of different components of nuclear alloys

such as 316L and 17-4 PH stainless steels. With the advances in the development of more efficient

systems and lasers, the productivity of SLM machines has increased. Also, the development of

new alloys has contributed to the exponential growth in the research on this topic. To work with

26
this technology, there is a curve of knowledge that has to be considered, but with the right

guidelines, workflow, documentation, and characterization, machines can be running production

as some industries have been reported.

L-PBF stainless steels would be cut by EDM in order to minimize the temperature and

residual stress effect on the microstructures. Additively manufactured parts would be put in

bakelite and polished with silica carbide papers with grate sizes ranging from 240 to 1200,

followed by polishing in micro polishing clothes with Al2O3 powders of grain sizes ranging from

5 μm to .05 μm and lastly with 0.03 μm colloidal silica suspension. The polished scratch-free

samples would then be carefully dried with acetone and air blow gun. An MTS Nanoindenter®

XP equipped with TestWorks® software and continuous Stiffness Measurement (CSM)

technology would be used in nanoindentation. For assessing micro-mechanical properties,

nanoindentation would be performed on the top surface of the additively manufactured alloys with

10 x 10 indents with a standard Berkovich tip with a radius of 50 nm and φ and β angles of 65.3°

and 12.95°, respectively. Besides, the Berkovich indenter, a flat indenter (FIB milled) will also be

used for the micro-compression test of micro-pillars. The samples will be collocated on the stage

assembly of the MTS Nanoindenter® XP for the nanoindentation test. The important parameters

that will be used for nanoindentation testing are the strain rate of 5×10−2 s-1, Poisson's ratio of 0.3,

and surface approach velocity of 10 nm/s. The instrument will be calibrated before each batch test

using a standard fused silica specimen. In an array of 10×10 indents with a spacing of 50 µm in

‘x’ and ‘y’ directions. The machine would make indents consecutively by minimizing the traveling

distance of the indenter head. The indentation depth limit for each indent will be in the range of

800 to 1500 nm to make it easier to find the indents on NanoSEM and also to get a better EBSD

map. Average values will be determined based on the values of all the indents' hardness and

27
modulus of elasticity. Statistical analysis will also be performed on each part by analyzing the data

provided by the nanosuite software.

The intention of using sharp Berkovich in this proposal is to study strain-rate sensitivity

and determine the indentation modulus of elasticity and hardness. Though, spherical

nanoindentation has recently become a tool to generate an effective stress-strain response which

can be compared with a uniaxial true stress-strain response by means of conversion and can give

an approximation of uniaxial yield strength.

Strain-rate sensitivity studies will be performed by choosing five different strain rates from

slow to fast loading rates of 0.01, 0.05, 0.1, 0.5, and 1 s-1. In this way, by implementing a simple

power-law relationship, the values of SRS (m) will be determined and will be correlated with

micro-mechanical properties, phases, microstructures, and micro-compression testing. This way

as discussed in the literature review section, it is possible to determine maximum shear strength

from the SRS study and yield strength from micropillar micro-compression.

In this study, microstructures will be analyzed using an optical microscope as well as

NanoSEM after the nanoindentation, micro-tension, and micro-compression tests. An FEI Nova

NanoSEM 230 with EDAX Octane Elite EDS and Hikari Super EBSD detector with an operating

voltage of 20 kV would be used to obtain an EBSD map of the nanoindentation region. Electron

Backscatter Diffraction (EBSD) provides (2D) maps of crystal orientation from polished cross-

sections. The latest version (v8) OIM Analysis™ has reset the standard for EBSD data analysis

capability with the addition of new functionality such as Neighbor Pattern Averaging and

Reindexing (NPAR™) and other features which enables users to achieve new insight into

microstructural characterization. The samples will be mirror polished and tilted 70˚ angle in order

to increase the diffraction efficiency during the EBSD acquisition. TEAMTM EBSD analysis

28
system will be used for EBSD data acquisition and TSL OIM Version 8 software for data analysis.

Typical scan area, step size, and average Confidence Index (CI) of EBSD acquisition would be

determined by surface texture and polishing quality. The additively manufactured high-

performance alloy samples crystal structures will be studied and phases will be determined to input

on the EBSD NanoSEM software. The selected phase/phases on the OIM software contained the

information necessary to model the electron backscatter patter (EBSP) produced by the expected

phase in the alloy’s samples. The hardness and modulus of elasticity of each indent will be

identified from the nanoindentation data and effort will be given to find out any correlation

between different crystallographic planes and mechanical properties layer by layer. 3D porosity

map would be achieved by using X-ray microscopy (AMMPI facility). This will facilitate the

micro-compression testing assuming that bigger micropillars would have more porosity compared

to smaller micropillars.

Figure 6: Layer-by-layer micro-mechanical properties by small-scale mechanical testing

29
2.2 Materials

The material used for the experimentation is LPW’s gas atomized stainless steel AISI 316L

powder particles of sizes ranging from 15 μm to 45 μm. The elemental chemical microanalysis

was performed using energy dispersive X-ray spectroscopy (EDS) in an FEI Quanta 200

Environmental SEM and the results are presented in Table 3. This instrument is capable of running

a low vacuum and thus able to reduce charging on an insulating sample and allowing imaging

without applying a conductive coating. The analysis is included in the experimental results in

Section 3.1. A conventional round rod of AISI 316L stainless steel was purchased for addressing

the difference between additively manufactured and conventionally manufactured parts, in terms

of microstructure and small-scale mechanical response.

Table 3: Elemental composition of LPW AISI 316L stainless steel powder particles
Element Minimum Maximum
Symbol Name wt% wt%

C Carbon 0.03
Cr Chromium 17.5 18
Cu Copper 0.5
Fe Iron Balance
Mn Manganese 2
Mo Molybdenum 2.25 2.5
N Nitrogen 0.1
Ni Nickel 12.5 13
O Oxygen 0.1
P Phosphorous 0.025
S Sulfur 0.01
Si Silicon 0.75

For this study, stainless steel 17-4 PH powder was procured from EOS® which is a leading

provider of metallic powder for additive manufacturing. According to the vendor, the parts built

from the powder should have chemical composition corresponding to ASTM F899 and certain

30
anisotropy which can be eased by solution annealing. The chemical composition of the powders is

listed in Table 4. The particle size is ranging from 16 µm to 63 µm and the as-built part density is

7.77 g/cm3. The elemental chemical composition analysis has been performed by implementing

energy dispersive X-ray spectroscopy (EDS) in an FEI Quanta 200 Environmental SEM. The

instrument can run at a low vacuum and thus able to minimize charging on the insulating sample

and allow imaging without applying a conductive coating. The analysis is included as the

experimental results and presented in section 2.5.

Table 4: EOS elemental composition of L-PBF 17-4 PH powder particles and printed parts
EOS 17-4 PH SS powder EDS of 17-4 PH SS
Min (%wt) Max (%wt) parts

Chromium (Cr) 15 17.5 16.30


Nickel (Ni) 3 5 4.40
Copper (Cu) 3 5 3.97
Silicon (Si) - 1 0.78
Manganese (Mn) - 1 1.55
Phosphorus (P) - 0.04 NA
Sulfur (S) - 0.003 0.30
Carbon (C) - 0.07 NA
Niobium + Tantalum (Ta) 0.15 0.45 0.33
Iron (Fe) Balance Balance

2.3 Process Parameters

In L-PBF, scanning strategy, processed parameters, and micro-mechanical properties are

required to be studied together for the ease of understanding the laser processing of powder metal

materials. There are unchangeable laser characteristics such as wavelength, beam energy

distribution during the fabricating process, and beam spot diameter for the end-user. On the other

hand, there is a set of process parameters in the L-PBF process that can be modified such as laser

power (P), hatch distance (HD), scan velocity (vs), and layer thickness (LT) with proper selection

31
of scanning strategy. The laser scanning strategy has been an important factor in determining better

surface quality and reducing residual stress effects in L-PBF parts. ‘Quad Islands’ was chosen as

a divisional scanning strategy. The schematics and the scanning strategy are illustrated in Figure

7. An experimental design was made by taking into consideration the parameters previously

mentioned along with some constraints for the determination of processing parameters. The laser

power was used in the range of 150 to 190 watts with VED ranging from 45.24 J/mm3 to 83.33

J/mm3. The other process parameters are: hatch distance (HD) = 120 μm, layer thickness (LT) =

50 μm and scanning speeds (vs) = ranging from 300 mm/s to 700 mm/s (Table 5). These parameters

were pre-defined based on previous investigations and on the machine manufacturer’s

recommendations to deliver functional parts with enhanced levels of surface quality (areal surface

roughness (Sa) between 14.24 µm and 24.34 µm).

Figure 7: Schematics of process parameters and scanning strategy for the fabrication of L-PBF AISI
316L parts.

Table 5: L-PBF AISI 316L stainless steel manufacturing process parameters


Sample P, Power HD, Hatch LT, Layer vs, Scanning VED
Number (Watts) Distance (μm) Thickness (μm) Speed (mm/s) (J/mm3)
1 190 120 50 700 45.24
2 180 120 50 600 50
3 170 120 50 500 56.67

32
Sample P, Power HD, Hatch LT, Layer vs, Scanning VED
Number (Watts) Distance (μm) Thickness (μm) Speed (mm/s) (J/mm3)
4 160 120 50 400 66.67
5 150 120 50 300 83.33

The geometry of the AM 316L stainless steel parts was 5 mm × 15 mm (height × diameter)

and the parts were manufactured using a laser powder bed fusion system from Aconity 3D

(AconityMIDI), with a maximum building volume of 170 mm of diameter and 400 mm of vertical

displacement. The radiation source was a single-mode fiber laser with a maximum power of P = 1

KW, a wavelength of λ= 1070 nm, and a spot size (diameter) in the range of 80-500 µm.

AM As-built 17-4 PH stainless steel parts have been produced as a cylinder with a

dimension of 5 mm × 20 mm (height × diameter). The laser powers have been used at 220 watts

and 230 watts from which volumetric energy density (VED) is calculated to be 54.76 J/mm3 and

61.11 J/mm3. The process parameters for additive manufacturing are shown in Table 6. A single-

mode fiber laser is being used as the radiation source with a maximum power of P = 1 KW along

with the above-mentioned parameters.

Table 6: L-PBF process parameters for the fabrication of 17-4 PH stainless steel parts

Sample P, Power HD, Hatch LT, Layer vs, Scanning


VED (J/mm3)
Number (Watts) Distance (μm) Thickness (μm) Speed (mm/s)

1 230 120 50 700 54.76


2 220 120 50 600 61.11

2.4 Sample Preparation

Both L-PBF AISI 316L and 17-4 PH stainless steels samples were prepared for the material

characterization. The first step was to carefully cut from the steel disc base by conventional

Electrical Discharge Machining (EDM) in order to minimize the temperature, unanticipated phase

33
formation, strain, and residual stress effects on the microstructures and overall on the micro-

mechanical properties. The samples were mounted with resin mixtures after applying the release

agent to mounting cups and to acquire a mirror-like surface, the grinding operation was carried out

in the sequence of silicon carbide papers with increasing grit sizes starting from 180 and then 320,

400, 600, 800, and 1200. Then the mounted samples were polished on polishing wheels embedded

with a fine Rayon polishing pad using a liquid suspension of alumina (Al2O3) powder particles

very carefully so that previous scratches would be removed starting with 5 µm, 1 µm then 0.05

µm, and lastly, 0.03 µm to get the mirror finish. Additionally, 0.03 µm colloidal silica was used

for a better result in the characterization processes. In the next step, the polished sample’s surface

was laved using an ultrasonic cleaning process and utilized to remove any remaining metal

particulate, dirt, and silica particles on the polished samples. This process uses ultrasound to agitate

solvent (distilled water/alcohol/acetone). The samples of conventional AISI 316 stainless steel

were prepared by following the same above procedure.

2.5 Proton Irradiation*

Before the proton irradiation, the range of the proton and created vacancies were estimated

utilizing a well-known ion-solid simulation code Stopping Range of Ions in Matter [SRIM/TRIM]

[103]. A detailed calculation with a full damage cascade was chosen within the TRIM code. The

1.0 MeV proton irradiation in the L-PBF 17-4 PH stainless steel target simulations was performed

using 160,000 ions. The target composition was taken as Cr (16 atomic %), Co (4 atomic %), Ni

(4 atomic %), Si (0.5 atomic %), C (0.035 atomic %), and Fe (75.465 atomic %) with a density of

* This section is reproduced from M.J. Uddin, H.R. Siller, R.A. Mirshams, T.A. Byers, B. Rout, Effects of proton
irradiation on nanoindentation strain-rate sensitivity and microstructural properties in L-PBF 17–4 PH stainless
steels, Mater. Sci. Eng. A. 837 (2022) 142719. https://doi.org/10.1016/j.msea.2022.142719, with permission from
Elsevier.

34
7.77 g/cm3, which is equivalent to 8.44×1022 atoms/cm3. The target layer was kept at a thickness

of 8 µm to speed up the calculations. In the SRIM calculation, different displacement energies of

25eV for the metal target atoms (Cr, Co, Fe, and Ni), 15 eV for silicon, and 28 eV for carbon were

considered to estimate the damages. The distribution of protons as a function of depth is given in

Figure 8(a). This plot shows that the maximum impact lies at a depth range of ~6.6 µm. The

variation in the number of vacancies generated and the displacements per atom (dpa) with target

depth when irradiated with a fluence of 1.0×1019 ions/cm2 are given in Figure 8(b). At the mean

depth of the 1 MeV proton beam (a depth of ~6.6 µm corresponding to ~33×10-4 vacancies/Ao-ion

and ~39 dpa. The samples were irradiated with 1.0 MeV protons with a fluences 1.0×1019 ions/cm2

from a 3MV Pelletron accelerator (NEC-9SH) [104]. The beams were irradiated at a target area of

0.4 cm × 0.3 cm with a beam flux of 2.6×1014 ions/cm2-sec. The proton irradiation occurred at a

vacuum of 2×10-8 torr. The samples were mounted on a stainless-steel holder with silver paste.

During the irradiation, the target temperature was periodically measured using an IR camera and

found to reach up to 60 oC.

Figure 8: SRIM simulated (a) distribution of 1 MeV proton in the 17-4PH steel, (b) Vacancies and
displacements per atom (dpa) as a function of depth for a proton irradiation fluence of 1.0×1019
ions/cm2.

35
2.6 Nanoindentation

An MTS Nanoindenter® XP equipped with TestWorks® software and Continuous

Stiffness Measurement (CSM) technology has been utilized at room temperature for

nanoindentation to determine micro-scale mechanical properties (indentation modulus and

hardness), strain-rate sensitivity (SRS) investigation, and theoretical shear strength of as-built and

irradiated parts. For the SRS study, five different strain rates (0.01, 0.05, 0.1, 0.5, and 1 s-1) have

been selected in five different regions of as-built 17-4 PH parts and an additional two strain rates

(0.25, 0.75 s-1) have been added for irradiated parts (4mm × 4mm area with 5μm depth) and it is

shown schematically in Figure 9. As-built (unirradiated) and irradiated areas are examined using

all the strain rates mentioned above. An indentation depth of 1500 nm is selected and a total of

100 (10× 10 array) indents in five (5) different regions have been made on each sample in as-built

and irradiated areas with 50 μm of spacing between two consecutive indents. A Berkovich tip (50

nm radius with φ and β angles of 65.3° and 12.95° respectively) is employed. The CSM has been

engaged with an amplitude of 2 nm and a frequency of 45 Hz. The CSM allows the measurement

of depth-dependent properties of any metal in a single step and it involves applying a dynamic

load on top of the static load while loading. It also helps in unloading stiffness continuously

through the indentation depth, by overlaying small oscillations over the primary loading signal.

The required inputs on the nanoindentation batch operation are being the surface approach velocity

of 10 nm/s, strain rate of 0.01 s-1 to 1 s-1, and the Poisson's ratio of 0.3. The instrument has been

carefully calibrated for each batch of the nanoindentation using a standard sample. The equipment

generates the values of hardness and indentation modulus as a spreadsheet by using the Nanosuite

software. Nanoindentation experiment has been designed in such a way so that leads to locating

easily any indents under Nano-SEM for microstructural observation and could be labeled with

36
measured micro-mechanical properties with different crystallographic orientations.

Figure 9: Nanoindentation experimental set-up for strain-rate sensitivity study on as-built and
irradiated L-PBF 17-4 PH stainless steel parts.

2.7 Strain-rate Sensitivity of 17-4 PH Stainless Steels

One of the useful features of nanoindentation is the capability of generating load-

displacement curves using a variety of methods such as displacement and load control modes using

continuous stiffness measurement (CSM).

Figure 10: Example of a load-displacement (P-h) curve of as-built L-PBF 17-4 PH part (VED = 61.11
J/mm3) at 0.1 s-1 strain-rate.

37
The loading portion of the load-displacement curve (Figure 10) can be taken as a power-law

relationship (Eq. 2) and contact stiffness, S can be defined as the slope of the unloading portion of

the curve as in Eq. 3 [105].

P = Chk (Eq. 2)

where P = load, C = loading curvature, and k = an exponent ranging from 1.5 to 2

dPu
S= at h = hm (Eq. 3)
dh

where S = contact stiffness, Pu = unloading force, h = displacement into the surface and hm =

maximum displacement into the surface

Strain-rate sensitivity (SRS) studies are generally based on conventional methods by

instrumented indentation for macroscopic materials testing. For better understanding in detail, a

textbook by Caliard [106] can be referred to. SRS is usually denoted as ‘m’ and the value of ‘m’

can be directly calculated from the power-law relationship (Eq. 5) between plastic stress, σ, and

strain rate, ε [107]. For instrumented indentation processes such as nanoindentation experiments,

stress can be replaced with nano-hardness since the stress is directly proportional to hardness with

a constraint factor C* (Eq. 4). The hardness values are determined by nanoindentation and plotted

against different strain-rates (0.01 s-1, 0.05 s-1, 0.1 s-1, 0.5 s-1, and 1 s-1) in log scale.

E∝H

H
E= C∗
(Eq. 4)

where C* = constraint factor (depends on the elastic-plastic deformation behavior of the material)

∂lnσ ∂ ln H
Strain − rate sensitivity, m = ∂ ln ε̇ ≅ ∂ ln ε̇

∂ ln H2 −∂ ln H1
= (Eq. 5)
∂ ln ε2̇ − ∂ ln ε1̇

= slope of the logarithmic load − displacement curve

38
Nanoindentation load-displacement curves are used to determine maximum shear stresses

with different strain rates by using displacement bursts (initial sudden fluctuations in

displacements) [108,109]. Incipient plasticity and elastoplastic deformation behavior of materials

are easy to apprehend with the help of these curves [110,111]. These curves from the very small

amount of strain rate are also very important to analyze dislocation activity and are related to the

first nucleation of dislocations. First displacements bursts are more significant compared to very

low strain-rate and become less prominent with the increase in strain rates. First displacement

bursts are not dependent on strain rates [108]. Elastoplastic behavior can be predicted from these

bursts which are separated by elastic regions. To determine maximum shear stress (Eq. 6), load-

displacement curves with 0.01 s-1 are selected for both L-PBF 17-4 PH parts before and after the

proton irradiation and only initial portions of the curves are considered. Using the loads

corresponding to displacement bursts, reduced modulus of elasticity, and Berkovich tip radius,

maximum shear stress has been calculated. The modulus of elasticity of L-PBF 17-4 PH is chosen

from a recent study [112]. The reduced modulus of elasticity (Er) is determined according to Eq.

7. Hardness values determined by nanoindenter can be converted to Vickers hardness as per

empirical relationships given in Eq. 8 and the yield strength of as-built and irradiated 17-4 PH

stainless steel parts has been calculated according to Eq. 9 which is applicable for BCC structured

iron-chromium alloys [113,114]. Factor 3.06 was determined empirically, especially for ferritic/

martensitic alloys in Eq. 9 those hardness values are highly affected by irradiation which is true

for L-PBF 17-4 PH stainless steels [115].


1
6PEr 2 3
Τmax = 0.31 � π3R2 � (Eq. 6)

where P = load correscponds to first displacement burts, Er = reduced modulus of elasticity, and R

= Berkovich indenter tip radius = 50 nm

39
−1
1−vsa 2 1−vin 2
Er = � + � (Eq. 7)
Esa Ein

where vsa and vin = Poisson rations of the sample and Berkovich indenter, Esa and Ein = modulus

of elasticity of the sample and Berkovich indenter (1000 Gpa) respectively

𝐻𝐻𝑣𝑣 = 94.495 𝐻𝐻 (Eq. 8)

𝜎𝜎𝑦𝑦 = 3.06 𝐻𝐻𝑣𝑣 (Eq. 9)

kg
where Hv = Vickers hardness �mm2� , H = nanohardness (GPa), σy = yield strength (MPa)

2.8 Small-scale Mechanical Testing

Small-scale mechanical testing has been widely utilized right after the commercialization

of nanoindentation in the 1980s and uniaxial micro-compression testing was carried out in 2004

right after it was possible to fabricate precisely smaller geometries by the implementation of the

focused ion beam (FIB) which opened the window to investigate in micro and nano-level

mechanical properties [116]. In this process, the bulk material is used in as-fabricated conditions

to fabricate or mill micro-sized pillar structures into the material surface and a very low current is

used to avoid FIB induced near surface damages [117].

Layer-by-layer micro-mechanical properties determination would be useful to study the

strength of the L-PBF layer on the fabricated parts as well as define the size effects with the

sample’s dimensions. As of now, there are no studies available in the literature for L-PBF

radiation-resistant alloys such as 17-4 PH stainless steels. To fabricate the pillars, different beam

currents have been used in rough, fine, and very fine milling in different stages during micropillar

fabrications. All the pillars of size ranging from 2 μm to 5 μm are punched by a 10 μm flat conical

diamond indenter in MTS XP nanoindenter at a strain rate of 0.01 s-1 and 2 μm pillars are milled

only on irradiated regions of 17-4 PH steels so that the height of those pillars (~5 μm) falls within

40
the radiation damage depth of around 5 μm. The micropillars arrays in as-built and irradiated

damaged areas with flat indenter are shown in Figure 11.

Figure 11: Micro-compression experimental set up (a) 10 um diameter flat indenter tip, (b) array of
micropillars in irradiated condition, and (c) array of micropillars in as-built condition of AM 17-4
PH stainless steel parts.

2.9 X-ray Diffraction (XRD) and Electron Backscattered Diffraction (EBSD) Microstructural
Characterization*

SEM with electron backscattered diffraction (EBSD) has been a useful technique for the

measurement and understanding of material’s textures, phase identification, grain size, and

misorientation angle distribution. The crystallographic orientations related to phases/grains are

provided by the EBSD at a precise location, allowing them to be related with each indent and their

corresponding measurement of hardness and indentation modulus. This capability is helpful in

explaining anisotropy within the scan area with indents having different crystallographic

orientations.

An FEI Nova NanoSEM 230 with EDAX Octane Elite EDS and Hikari Super EBSD

detector with an operating voltage of 20 kV with suitable current, was employed to generate EBSD

maps of smaller (~0.02 mm2) and larger nanoindentation areas (~2.3 mm2). EBSD creates 2D

maps of crystal orientation from cross-sections to find the relationship between the

* This section is reproduced from M.J. Uddin, E. Ramirez-Cedillo, R.A. Mirshams, H.R. Siller, Nanoindentation
and electron backscatter diffraction mapping in laser powder bed fusion of stainless steel 316L, Mater. Charact. 174
(2021) 111047. https://doi.org/10.1016/j.matchar.2021.111047, with permission from Elsevier.

41
crystallographic planes with proper indexing of nanoindents for the characterization of anisotropic

behavior and micro-mechanical properties at the sub-micron scale. The mirror-polished samples

were pre-tilted by 70º to escalate the diffraction efficiency during the acquisition. OIM Analysis™

(v8) software with Neighbor Pattern Averaging and Reindexing (NPAR™) was used to achieve

the best possible microstructural characterization and data analysis, and TEAMTM EBSD analysis

system was used for EBSD data acquisition. Harmonic textures were generated and harmonic

series expansion was used as the calculation method with a Gaussian smoothing angle of 5֯.

The area of interest was identified and focused on SEM. Custom auto-focus and auto-

brightness/contrast routines were initiated to boost the similarity of the image contrast from layer

to layer. These custom algorithms were implemented versus the vendor-provided algorithms

because they influence the information from previous layers to improve the reliability of the

results. For each focused full-frame region, secondary electron (SE) and backscattered electron

(BSE) images were collected from the same area. Both images were pixelated, covering an area of

µm2/mm2, acquired with an accelerating voltage of 20 kV, a working distance of 12 mm to 14

mm depending on the sample dimension, a pixel dwell time of 0.1 s, and a beam current of 6.1 nA.

Approximately, a quarter of the full region of interest on the BSE images shows that although there

is a contrast between the polycrystalline grains, it is preposterous to characterize the grain

boundary interfaces because of the high plastic strains evolved during the AM solidification

process. Moreover, the pores and defects locations are observed in the BSE image as dark regions.

The scan areas were varied from 165 µm × 135 µm (for smaller maps) to 1.72 mm × 1.35 mm (for

larger maps), the step size was in the range of 0.3 µm to 5 µm depending on the scan area and

condition of the sample’s surface, and the average Confidence Index (CI) of EBSD acquisition

was around 0.90 to 0.96 which confirms valid scans for further studies. The L-PBF AISI 316L

42
parts have a face-centered cubic (FCC) crystal structure or γ-phase (austenite). The selected

austenite phase on the OIM software contained the information necessary to model the EBSD

images produced by the expected phase in the sample. The hardness and indentation modulus of

each indent was identified from the nanoindentation data on the EBSD maps and the correlation

between different crystallographic planes, VED, and those micro-mechanical properties were

thoroughly studied to better explain the anisotropy and heterogeneity in microstructure and micro

mechanical behavior.

Optical microscopes were used after submerging the polished surface in an etchant to

observe the microstructures after proper calibration, and ESEM and NanoSEM were utilized for

the study of grain morphologies. Planar and cellular grain morphologies are very common in L-

PBF AM 316L due to thermal gradients that occurred during the fabrication process as the

combined results of rapid solidification, remelting of powder particles, and repeated heating-

cooling cycles.

2.10 Porosity Analysis*

A non-destructive microscopic technique Zeiss XRADIA Versa 520 X-ray microscope

(XRM) has been utilized to determine the volume percentage (%) pores (micro-cracks if there are

any) in L-PBF 17-4 PH as-built and irradiated stainless steel parts. A volume of 2.9 × 2.9 × 2.9

mm3 has been scanned with a voxel size of 1 μm and an objective lens of 4X. Pores vol.% is the

ratio of volume hold by pores and cracks on the scanned volume to the total scanned volume of

the sample. Primary scan parameters that are used to generate X-ray tomograms have been

* This section is reproduced from M.J. Uddin, H.R. Siller, R.A. Mirshams, T.A. Byers, B. Rout, Effects of proton
irradiation on nanoindentation strain-rate sensitivity and microstructural properties in L-PBF 17–4 PH stainless
steels, Mater. Sci. Eng. A. 837 (2022) 142719. https://doi.org/10.1016/j.msea.2022.142719, with permission from
Elsevier.

43
summarized in Table 7. Data and images are analyzed in Dragonfly and ImageJ software.

Table 7: XRM Scan parameters


Parameter Value/info
Imaging mode Tomography
Binning 2
Voltage 140 KV
Current 64.34 μA
Source filter HE4
Objective 4X
Sample theta -180֯ to 180֯
Exposure time 2 sec
Pixel size 3 μm
No. of projections 2401
Width 2.9 mm
Height 2.9 mm
Depth 2.9 mm
Volume 25.75 mm3
Voxel size 1 μm

2.11 Fabrication of Micropillars by Focused Ion Beam (FIB)

Micropillars were fabricated on AM 17-4 PH stainless steel parts in as-built and irradiated

conditions using FEI Nova FIB/SEM 230 by utilizing coarse and fine milling using a 30.0 kV

operating voltage and a milling current in the range of 20 nA to 0.3 nA depending on the materials

etching responses. Modifications to the operating condition of FIB were made if it was necessary.

Micropillars of various diameters (2 μm to 5 μm) were milled with a diameter to height ratio of 1:

2-2.5. A very fine milling procedure was performed while fabricating 2 μm micropillars.

44
CHAPTER 3

RESULTS AND DISCUSSIONS

3.1 Nanoindentation and Electron Backscatter Diffraction Mapping in Laser Powder Bed
Fusion of Stainless Steel 316L*

This research work presents the anisotropic behavior in microstructure and micro-

mechanical properties of L-PBF AISI 316L as well as conventional 316L stainless steels based on

the observation of electron backscatter diffraction (EBSD) and nanoindentation maps in relation

to various volumetric energy densities (VED) and grains at different crystallographic orientations.

It was noticed that the variations of VED, calculated from different process parameters

combinations, have an influence on crystallographically oriented grain growth evolution affecting

the nanoindentation performance. This progression was responsible for the variation in grain size

and average misorientation angles and was found to have an important role in determining the

micro-mechanical behavior. The highest nanoindentation response was obtained at a VED=56.67

J/mm3, with an average indentation modulus of 196.77 GPa and hardness of 3.68 GPa. From

Texture Pole Figures (PFs) and Inverse Pole Figures (IPFs) of EBSD maps, it was revealed that

the highest texture strength of <101> oriented grains along with the evolution and domain of

<111> oriented grains enhanced the micro-mechanical properties at that VED value. The other

causes of this behavior are the relatively lower grain size (11.62 μm) and a higher average

misorientation angle (11.74°) when comparing it with the rest of the L-PBF AISI 316L parts. The

additional but undesirable strong <001> texture somehow contributed to the slight mitigation of

the micro-mechanical properties in conventional AISI 316L stainless steels and these parts were

found to be comparatively less anisotropic than the parts obtained from the L-PBF process.

* This section is reproduced from M.J. Uddin, E. Ramirez-Cedillo, R.A. Mirshams, H.R. Siller, Nanoindentation
and electron backscatter diffraction mapping in laser powder bed fusion of stainless steel 316L, Mater. Charact. 174
(2021) 111047. https://doi.org/10.1016/j.matchar.2021.111047, with permission from Elsevier.

45
3.1.1 Energy Dispersive X-ray Spectroscopy (EDS)

The weight percentage of each possible element in Table 8 was analyzed by quantitative

analysis by FEI ESEM. Most of the elements were in the range of LPW’s range, nevertheless, the

EDS measurement was not able to quantify the percentage of Carbon (C), Nitrogen (N), and

Phosphorus (P), and collectively, these three elements were only accounted for 0.155 wt% and

mapping showed no segregation. Thus, EDS quantitative analysis was in good agreement with the

balance of the alloys and similar findings were reported in the literature [118]. Cu, Mn, Mo, and

Si weight percentages showed lower values than expected and in the case of O and S, the weight

percentage was found to be slightly higher than the supplier’s composition data.

Table 8: Elemental quantitative analysis of AISI 316L powder by energy dispersive spectroscopy
(EDS)
Element
wt%
Symbol Name
C Carbon NA
Cr Chromium 17.4
Cu Copper 0.24
Fe Iron 66.25
Mn Manganese 1.43
Mo Molybdenum 0.89
N Nitrogen NA
Ni Nickel 12.27
O Oxygen 0.69
P Phosphorous NA
S Sulphur 0.3
Si Silicon 0.58

3.1.2 Nanoindentation Results

One-way analysis of variance (ANOVA) of nanoindentation hardness and modulus results

by Minitab is shown in Figure 12, Table 9, and Table 10 with a confidence level of 95%. This

analysis is very useful when a categorial factor and a continuous response are assuming each data

46
has equal variance (Tukey comparison procedure). It helps to determine whether the means of two

groups of properties are different. The interval plots as shown in Figure 12 display the confidence

intervals for the differences between the means of indentation modulus and hardness for the

selected comparison method. It can be stated from Figure 12 that the hardness found by the

nanoindentation experiment is not statistically significant (since the α-value is more than 0.05

which is the significance level) compared to the indentation modulus which was found to be

statistically very significant. In Figure 12(b) interval plot of hardness, the means are seemed to be

displaced with different values of volumetric energy densities. However, by running a mean

comparison, a trend can be identified that the mean at VED of 56.67 J/mm3 is the highest in the

set of experiments.

Figure 12: Regression analysis (ANOVA) of nanoindentation hardness and modulus data with
various VED; (a) Interval plot of indentation modulus and (b) interval plot of hardness with a
confidence level of 95%.

Table 9: Analysis of variance (ANOVA) of the influence of volumetric energy density (VED) on the
indentation modulus (significance level, α =0.05)
Degrees of Adj Sum of Adj Mean
Source F-value P-value
Freedom Squares Squares
VED 4 8761 2190.2 4.01 0.004
Error 135 73683 545.8
Total 139 82444

47
Table 10: Analysis of variance (ANOVA) of the influence of volumetric energy density (VED) on
hardness values by nanoindentation (significance level, α =0.05)
Degrees of Adj Sum of Adj Mean
Source F-value P-value
Freedom Squares Squares
VED 4 1.747 0.4368 1.33 0.261
Error 130 42.6 0.3277
Total 134 44.347

The mean indentation modulus and hardness values were observed to be in the range of

175 GPa to 196.77 GPa and 3.40 GPa to 3.68 GPa respectively for L-PBF AISI 316L stainless

steels parts. The mean indentation modulus and hardness were found for the conventional AISI

316L sample to be 189.15 GPa and 3.60 GPa respectively. As per the ASM handbook [119], the

indentation modulus of conventional AISI Type AISI 316L stainless steel is 193 GPa, slightly

higher than the value found in this study. The L-PBF AISI 316L part that was fabricated with 56.67

J/mm3 VED exhibited the best micro-mechanical response with respect to indentation modulus

(196.77 GPa) and hardness (3.68 GPa). Further increase in VED after 56.67 J/mm3, a decrease in

indentation modulus as well as hardness was noticed. The lowest mean indentation modulus and

hardness were found in the parts with the highest VED of 83.33 J/mm3 valued as 175 GPa and

3.40 GPa respectively which were very close to the values found on VED 45.24 J/mm3 as 177.27

GPa (indentation modulus) and 3.43 GPa (hardness).

Nanoindentation load-displacement curves generally provide a mechanical fingerprint of

the response of a material to contact deformation. The unloading curves obtained in the

nanoindentation tests were found to be well behaved (exhibiting almost linear behavior) and those

curves did not display considerable elastic recovery. Thus, the nanoindentation test was confirmed

to obtain valid mechanical property data for L-PBF and conventional AISI 316L parts.

48
3.1.3 Grain Morphology and Microstructures

Detailed grain morphologies are shown in Figure 13 and Figure 14. Microstructural

characteristics located in layer-layer weld pool boundary and track-track weld pool boundary are

shown in Fig. 5. Different regions of microstructural characteristics were found in the

metallographic study and different crystallographic grain growth modes were observed in SEM,

which confirmed some distinctive morphologies, as reported in many recent studies of L-PBF of

AISI 316L [120–131].

Figure 13: SEM images showing build direction (BD) and grain morphology of L-PBF AISI 316L
parts with different volumetric energy density (VED); (a) and (b) VED = 45.24 J/mm3, (c) and (d)
VED = 50 J/mm3, (e) and (f) VED = 56.67 J/mm3.

49
SEM scans of L-PBF AISI 316L parts (VED of 45.24 J/mm3 to 56.67 J/mm3) and their

corresponding magnified scans are shown in Figure 13(a) to Figure 13(f). Features of each part are

outlined in the figures for a better understanding. Cellular morphologies along with dendritic

crystallographic (region 1) and planar grain region (region 2) were found to appear in every sample

either clearly visible on the figures or on higher magnification scans. Additionally, a variation in

the growth orientation of dendritic structures was observed. The finest cellular structures were

found in the parts with a volumetric energy density of 56.67 J/mm3 (Figure 13(c), and Figure 13(d))

and 50 J/mm3 (Figure 13(e), and Figure 13(f)). Planar interface growth was noticed as the principal

crystal growth pattern in the black dashed area (region 2). The growth of new crystalline layers of

planar grains formed epitaxially at the solidified cellular grains along the fusion line in the normal

direction shown in Figure 13(d) and Figure 13(f), which is consistent with the findings in our

previous study. Planar and finer cellular microstructures are very common microstructures found

in L-PBF techniques for a similar types of stainless-steel grades. A distinct type of solidification

process associated with inhomogeneous nucleation of grains in a new weld pool of L-PBF AISI

316L parts is responsible to create the parts having anisotropic and heterogenous microstructures.

Additionally, directional heat extraction, repeated melting, and rapid solidification on the same

layer of powder metal particles play an important role as reported in one study [67]. Variations in

grain size are found in different fabrication planes to build direction and hence, anisotropy in

mechanical properties is proven to be evident in another study [132]. In the present study, planar

interface growth was noticed as the principal crystal growth pattern in the black dashed area

(region 2). The growth of new crystalline layers of planar grains formed epitaxially at the solidified

cellular grains along the fusion line in the normal direction shown in Figure 13(d) and Figure 13(f),

which is true according to theory and literature [133,134].

50
The angle between the growth orientation of dendrites and the tangential of fusion may be

affected not only by heat flux direction but also by preferred growth orientation related to the

crystal structure. Researchers studied and determined the reason for the variation of the planar

grains in the weld pool [128]. They found out that there is a relationship between the preferred

growth direction and the preferred crystal orientation of the planar grains.

In another study, it was predicted that Marangoni flow could change the heat flux direction,

which could cause convective heat transfer and flow of fluid within a melt pool. Eventually, this

phenomenon could lead to altering growth orientation along the fusion line as shown in Figure

13(d) and Figure 13(f) with finer black dash lines. Similar findings were reported in a study [135].

The crystallographic growth mode was noticed to change from planar to cellular and finally to

dendritic solidification mode as illustrated in Figure 13(d) where planar grain growth was observed

along with line AP and dendritic grain growth was observed along with line AD in agreement with

[123]. These distinct solidification modes can be attributed to changes in temperature distribution

across AP and AD within the weld pool during the manufacturing process. Planar growth along

AP is favorable due to the higher thermal gradient; on the other hand, the thermal gradient is lower

along AB, which results in the formation of a constitutional undercooling zone and instigates a

dendritic growth mode. For higher volumetric energy density parts, coarser grains were found later

in the quantitative analysis of EBSD data. Theoretically, grains choose to develop in the

crystallographic orientation of the parent grains in stainless steels manufactured by L-PBF, and

simultaneously, grains prefer to grow along the heat transfer direction, arising the competitive

process among grains. Furthermore, part of the previously solidified layer is exposed to the melting

and cooling cycle during the L-PBF fabrication process. In the same fashion, the phenomenon of

remelting of the same solidified layers unquestionably refrains from nucleation for solidification

51
which sanctioned the commencing of epitaxial growth at the partially remelted grains and that

facilitates the growth of grains perpendicular to the curvature of the weld lines, as illustrated in

Figure 14(a) to Figure 14(c). It is worth to mention that crystallographic orientation of cellular

sub-grains was observed to develop at around 45o away from the building direction and met at

pool boundaries as visible from Figure 14(b).

Figure 14: SEM images showing planar and cellular dendritic grains with build direction (BD) of L-
PBF AISI 316L parts with volumetric energy density (VED)=50 J/mm3; (a) area of SEM scan, (b)
cellular dendritic structures with preferential 45o angle growth with pool boundary, and (c) planar
grains.

Microstructural characteristics were studied using an optical microscope. These

characteristics are shown in Figure 15(a) to Figure 15(b) with details. Polished samples were dried

52
by following standard procedure, where defects and pores were visible even with naked eyes, and

etched with Marble’s (4g CuSO4 and 20 ml HCl added to 20 ml distilled water) reagent as long as

the sample surface seemed to be lightly reacted with it. Successive weld pool layers are visible as

two lines (one dashed and one solid black lines) in Figure 15(a) and weld pools overlap partially

at a particular angle because of the scan rotation [120]. The weld pools are wider and less uniform

than the layer thickness (50 µm) which confirms the remelting of the previously solidified layer.

The overlapping nature of the weld pool is a confirmation of the successful fusion of powder

particles and strong bonding within successive layers. Coarsening of grains was observed with

parts with increasing volumetric energy density.

Figure 15: Microstructures of L-PBF AISI 316L parts with different process parameters; (a) VED =
50 J/mm3, (b) VED = 56.67 J/mm3.

One grain seems to extend over two consecutive layers as marked with the white arrow in

Figure 15(a) and this can be addressed as epitaxial grain growth. The grain orientations take place

according to solid nucleus surface orientation (surface of the previously applied layer which

determines the surface of the crystallization) during the solidification process in laser powder bed

fusion [127,129]. Grains were found to grow toward the weld pool center line and upon reaching

a certain size their surface met and formed grain boundaries. Moreover, the presence of heavy

53
elements such as Mo and the accumulation of these elements could also contribute to a high

solidification rate and corresponding constitutional undercooling [118].

3.1.4 Electron Backscattered Diffraction (EBSD) Maps

The electron backscattered diffraction (EBSD) maps along with harmonic texture i.e. pole

figure (PF) and inverse pole figure (IPF) are illustrated in Figure 16, Figure 17, and Figure 18 for

all L-PBF AISI 316L parts. In all these figures, volumetric energy density is shown to increase

downward. In Figure 16, EBSD maps of L-PBF AISI 316L parts with the lowest volumetric energy

density are displayed (VED= 45.24 J/mm3 (Figure 16(a)) and 50 J/mm3 (Figure 16(d)). These maps

were generated on the region of indents. The black triangular areas are shown in Figure 19 and

labeled with micro-mechanical response i.e. hardness and indentation modulus for better

visualization of indents together with the grain orientation. Scans were generated from a very small

area (0.02 mm2) to larger areas (2.13 mm2) to illustrate the surface texture of oriented grains, and

changes in grain orientation and to study the extent of anisotropy with the variation of volumetric

energy density. In order to quantify the effect of VED on the texture intensities of L-PBF AISI

316L parts, the texture index (PFs) and texture (IPFs) are also introduced. The pole figures quantify

the texture intensity for three crystallographic direction families <001>, <101>, and <111> with

respect to the front view of the sample (top surface). On the other hand, inverse pole figures are

2D representations of orientations. IPFs can be plotted as points and each point represents a sample

direction. Crystallographic directions in the sample coordinate system are plotted in a pole figure

whereas sample directions in the crystallographic coordinate system are plotted in an inverse pole

figure.

The colors on the EBSD maps are related to the orientation of crystal structures in the

samples fabricated perpendicular to the building direction. It is noticeable from the maps that the

54
grains are randomly oriented and consist of finer, coarse, and elongated grains. Elongated grains

were found to build up almost perpendicularly to the weld pool lines as shown in Figure 16(a) and

Figure 16(d). Some of the grains, marked in Figure 16 and Figure 18, were very large over 100

µm in size. It is noticeable that the very large grains were found to be mostly <101> oriented (green

color in IPF color map). The values of indentation modulus found in <101> oriented grains were

around 170 GPa to 180 GPa. A number of elongated grains in the scanned EBSD maps were

noticed to decrease with the increase in VED in all L-PBF parts. Parallelly, the increase in the

number of grains oriented in the <101> direction was observed with increasing VED.

Figure 16: EBSD maps of L-PBF AISI 316L parts showing one of three nanoindentation regions with
the variation of volumetric energy density; (a) EBSD and IPF color map, (b) texture PF and (c)
texture IPF of parts (VED = 45.24 J/mm3), and (d) EBSD map, (e) texture, and (f) texture IPF of
parts (VED = 50 J/mm3).

55
Figure 17: EBSD maps of L-PBF AISI 316L parts showing one of three nanoindentation regions with
the variation of volumetric energy density; (a) EBSD map, (b) texture PF and (c) texture IPF of parts
(VED = 56.67 J/mm3), and followed by (d) EBSD map and IPF color map, (e) texture PF and (f)
texture IPF of parts (VED = 66.67 J/mm3).

The maximal texture indexes (PFs) for parts with VED of 45.24 J/mm3 and 50 J/mm3 were

calculated as 3.886 and 2.917 respectively and a strong <110> texture was found in both parts. For

isotropic materials, the texture index is equal to unity. The tendency to evolve towards <001> and

<111> poles from strong <110> texture was observed with increasing VED. The IPFs indicated

that the <101> highly texture strength was found in all the parts regardless of the extent of VED

for the front view of the samples [001]. The parts with a VED of 56.67 J/mm3 were highly <101>

and <001> textured (Figure 17(c)) compared to other parts and possibly could be the reason for

exhibiting higher micro-mechanical responses. The EBSD map generated for the part with a VED

of 83.33 J/mm3 (Figure 18) was identical to the lowest VED parts except the strongest <110>

textured with a maximal texture index of 4.177 and the highest texture strength of 3.269 among all

56
the parts. This part displayed the most anisotropic behavior with relatively lower indentation

modulus and hardness when considering smaller area scans. Texture strength did not show any

particular type of trend to relate with VED in small area scans illustrated in Figure 16 to Figure

18. It is noticeable from these figures that there is a change in color of crystallized grains with the

variations in VED.

The increase in VED was responsible for evolving to more <101> and <001> oriented

grains in the microstructures. Considering larger area scans, that are added as supplementary

material in Appendix, clearly revealed a strong relationship between harmonic textures and

volumetric energy density along with a better understanding of the certain crystallographic

orientation of grains.

Figure 18: EBSD map of L-PBF AISI 316L parts showing one of three nanoindentation regions with
the variation of volumetric energy density; (a) EBSD map, (b) texture PF and (c) texture IPF of parts
(VED = 83.33 J/mm3), and (d) original image quality map of part with VED = 50 J/mm3, and (e) IPF
color map.

57
Figure 19: EBSD maps of all L-PBF AISI 316L parts marked as rectangles in Fig. 16, Fig. 17 and Fig.
18. All indents are labeled with indentation modulus (IM) and hardness (H) generated by the
nanoindentation experiment; (a) VED = 45.24 J/mm3, (b) VED = 50 J/mm3, (c) VED = 56.67 J/mm3,
(d) VED = 66.67 J/mm3, (e) VED = 83.33 J/mm3, and (f) original Image Quality (IQ) of part with VED
= 66.67 J/mm3 and IPF color map.

There was a similarity in textures and microstructures observed in parts with VED of 45.24

J/mm3, 50 J/mm3, and 56.67 J/mm3. The planar grains were more present and visible in these parts

than in the parts with higher VED (66.67 J/mm3 and 83.33 J/mm3). It is known that the preferential

growth direction for FCC structured materials such as AISI 316L is <001>. The grains started to

grow from the weld line area in the closer region of <001> direction indicated on the IPF color

58
map, and each layer melted at least two times and contributed to the reduction in grain width as

well as to a change in the final texture of the EBSD cross-section. Identical results were discovered

in recent literature [120]. As reported in another study, a strong epitaxy is noticed with planar grain

in <001> and <011> directions [136]. In this study, it was observed that the dendritic cells

nucleating epitaxially tend to grow most rapidly and thus eliminating other orientations gradually

since those cells require less nucleation energy. The elongated grains became wider and larger in

size on the surface of L-PBF AISI 316L fabricated parts with increasing VED and it was due to

the directional solidification and thus related to the epitaxial growth. Newly nucleated grains of

small planar shapes in zigzag orientations were spotted around the intersections of the weld pool

boundaries. In contrast, dominant orientations of the grains were not so evident on the parts with

higher volumetric energy density. The elastic modulus and hardness of each indent in one region

of the nanoindentation array of each L-PBF part, were labeled and identified on the EBSD maps

and these are shown together in Figure 19.

Notable anisotropic behavior in hardness and indentation modulus (IM) depending on the

crystallographic orientation of grains was observed in all the parts regardless of the amount of the

volumetric energy density. Gradual decrease in preferential orientation <101> (green color) and

gradual increase in preferential orientation <111> and <100> (increase in blueish and reddish

color) were noticed. Grains oriented in mixed yellow-green colored in IPF (the mid-region of

<011> and <001> line on the IPF) showed higher values in hardness and indentation modulus as

in Figure 19(a) as compared to <133> (around mid-region of <101> to <111>). The indent with

the highest indentation modulus (198.2 GPa) was found to be intersected with a few grains (left

bottom indent in Figure 19(a)). As the VED increases, the fraction of blueish and greenish color-

oriented grains started to evolve and dominate in EBSD maps as it has been already discussed.

59
Figure 20: (a) EBSD maps of conventionally manufactured AISI 316L stainless steel, (b) texture PF,
(c) texture IPF, (d) rectangular area with indent’s hardness (H) and indentation modulus (IM), (e)
original IQ, and, (f) IPF color map.

The purple color-oriented grains appeared more in 56.67 J/mm3 parts when compared to 50 J/mm3

and 66.67 J/mm3 parts and could be one of the possible reasons for exhibiting enhanced micro-

mechanical responses compared to all the parts regardless of the extent of VED. The indent in the

middle of the last row in Figure 19(c) displayed the highest IM = 226.6 GPa. The blue and purple

color-oriented grains were transformed to reddish and green color-oriented grains with increasing

VED (as shown in Figure 18(a) and Figure 19(e)) and those particular grain orientations eventually

contributed to a decrease in the values of IM and hardness in parts with VED = 66.67 J/mm3.

Completely <001> (red color in IPF) oriented grains displayed IM of around 145 GPa to 150 GPa

and yellow-colored grains exhibited slightly higher IM values in 150s GPa as it is clearly

noticeable from Figure 19(d) and Figure 19(e). Retained purple and blueish color-oriented grains

60
continued to show higher IM and hardness. As a whole, it can be said that an overall decrease in

micro-mechanical behavior was noticed. With the increase in VED, EBSD maps displayed almost

all red and green colored grains on the top surface of the L-PBF AISI 316 L part and those grains

collectively contributed and lowered the average values of IM and hardness.

3.1.5 Validation of the Indentation Results

The indentation micro-mechanical properties within a particular crystallographic

orientation of a single larger grain were studied and illustrated in Figure 21.

Figure 21: Indentation modulus (IM) and hardness (H) within same crystallographic orientation
grains from EBSD maps of AISI 316L parts with (a) VED = 56.67 J/mm3 and (b) VED = 66.67 J/mm3.

For this purpose, two L-PBF 316L parts with VED of 56.67 J/mm3 and 66.67 J/mm3 were selected.

Grains 1, 2, and 3 were chosen from the EBSD scan as shown in Figure 17(a) and grains 4, 5, 6,

7, 8, and 9 were chosen from the EBSD scan as shown in Figure 17(d). The average hardness

values, indentation modulus, and standard deviations (SD) were shown along with crystallographic

orientation in Figure 21. For the grain with <112> direction (grain 1), four indents were made and

61
average hardness and indentation modulus were found to be 2.89 GPa and 171.95 GPa

respectively. As discussed in earlier sections, and validating previous results, the evolution of

blue/dark blue/sky blue grains in parts with 56.67 J/mm3 exhibited higher hardness and indentation

modulus when compared with other crystallographic orientated grains. <122> oriented grains

showed the highest average hardness at 3.66 GPa and indentation modulus at 210.70 GPa. Pink

and red colored grains with orientations <114> and <001> respectively showed lower values in

indentation micro-mechanical properties. It was observed that <001> oriented grain (9) displayed

the lowest average values indentation modulus determined by nanoindentation as 161.67 GPa with

a standard deviation of 9.83. Whereas, the green-colored grain (<101>) exhibited the lowest

average hardness values of 2.8 GPa with a standard deviation (SD) of 0.61.

3.1.6 Misorientation Angle and Grain Size Estimation

The misorientation angles distribution can also influence the micro-mechanical properties

as high angle grain boundaries (HAGBs) pose larger resistance to start a dislocation motion than

low angle grain boundaries (LAGBs) and therefore, materials with a higher number fraction of

HAGBs can lead to relatively higher average indentation modulus and hardness values generated

by nanoindentation experiment. Simplifying the misorientation angle distribution to the average

misorientation angle could be a simple way to explain if there is any relationship between the

misorientation angle and micro-mechanical properties. In Figure 22, micro-mechanical behaviors

of all L-PBF AISI 316L parts are summarized with average grain size and average misorientation

angles. These quantitative analysis values were derived from larger EBSD maps for better

reliability, and those maps are provided in the Appendix as supplementary data.

Generally, low angle grain boundaries (LAGBs) or sub-grain boundaries are those with a

misorientation angle of less than 5°. Larger misorientation angle such as more than 10° yields high

62
angle grain boundaries. Misorientation angles were found to increase with an increase of the VED

as shown in Figure 22 for samples with VED between 45.24 J/mm3 and 56.67 J/mm3. Nevertheless,

misorientation angles were found to decrease after that range for the parts with VED of 66.67

J/mm3 and 83.33 J/mm3, a behavior that requires further studies. The average misorientation angle

of conventional AISI 316L stainless steel was computed as 13.23° and it was higher than all the

AISI 316L parts fabricated by L-PBF. In summary, the sample with a VED of 56.67 J/mm3

exhibited the highest misorientation angle of 11.74° and displayed enhanced micro-mechanical

properties among all of the L-PBF parts.

The average cell size of cellular grains was determined approximately from higher

magnification images of all the L-PBF 316L parts, considering the cellular grains are circular. The

grain morphology was found to be not uniform throughout all the parts with different VEDs. The

average cellular cell size of parts with VED of 45.24 J/mm3, 50 J/mm3, 56.67 J/mm3, 66.67 J/mm3,

and 83.33 J/mm3 was found to be 1.08 μm, 0.34 μm, 0.75 μm, 1.23 μm, and 1.02 μm respectively.

Figure 22: Effect of grain size and misorientation angle on indentation modulus and hardness based
on volumetric energy density (VED).

In case of grain size analysis from OIM quantitative analysis of EBSD scans, the parts

(except parts with VED of 45.24 J/mm3) having higher average indentation and hardness exhibited

the finer grains with proper distribution of all types of grains discussed in the previous section.

63
Similar to the misorientation angle case, the grain size in diameter was noticed to increase with

the increase of VED starting from parts having a VED of 45.24 J/mm3 to 56.67 J/mm3. The size

of grains usually increases at higher temperatures (less grain boundary and HAGBs areas) and

mostly the finer grain structure is preferable to achieve slightly enhanced mechanical properties.

The finest grains were found on parts fabricated with the lowest VED of 45.24 J/mm3 which size

was 9.22 µm in average, and parts fabricated with a VED of 66.67 J/mm3 displayed the coarsest

grains (13.77 µm). The upward trend was found consistent when increasing the VED and the

average grain sizes were 17.21 µm, 18.24 µm, and 20.75 µm for the parts with VED 45.24 J/mm3,

56.67 J/mm3, and 83.33 J/mm3 respectively. Compared to L-PBF AISI 316L parts, conventionally

manufactured 316L exhibited finer and similar-sized grains with 5.97 µm (average in diameter)

due to the absence of a complex solidification process followed by directional grain growth. In

this study, it would be worth stating that the L-PBF parts with higher average misorientation angles

along with relatively smaller average grains displayed enhanced micro-mechanical properties i.e.

higher indentation modulus and hardness value.

3.1.7 Conclusions on Nanoindentation, Anisotropy, and Mechanical Properties

The use of nanoindentation in combination with electron backscatter diffraction mapping

is useful to characterize the anisotropy in micro-mechanical properties and microstructure in

additively manufactured parts made of AISI 316L alloy. After performing this characterization

methodology, the key findings can be summarized as follows:

The mode of crystallization and preferential grain growth exert an influence on the micro-

mechanical properties in L-PBF processed parts. L-PBF parts with VED 56.67 J/mm3 displayed a

similar fraction of <101> oriented grains when compared with other L-PBF parts with a variation

of VED but, the evolution of grains oriented along the line <001> to <111> (blue/blueish to light

64
purple in IPF color map) was highly visible and unique among all the other L-PBF parts. The

indents on <111> grain orientations displayed relatively higher values of indentation modulus and

hardness in the nanoindentation experiment and generated a higher value of misorientation angles

as well. The evolution of certain grain orientations was observed and a change in micro-mechanical

properties was reported with the variation of volumetric energy density. The evolution of particular

crystallographic orientations of grains in parts with VED of 56.67 J/mm3 (not observed in other L-

PBF parts) was most probably responsible for exhibiting better micro-mechanical properties. The

higher angle grain boundary region acts as a strong barrier against dislocation movement.

Dislocations pile-up when that region has gone under stress and prevent crack propagation at a

certain limit. The relatively lower grain size accompanied by a higher average misorientation

angle of L-PBF AISI 316L parts with VED of 56.67 J/mm3 enhanced the micro-mechanical

properties when compared to other fabricated parts with different VED.

The conventional AISI 316L stainless steel showed the lowest texture strength, but the

indentation modulus and hardness were lower than that of L-PBF AISI 316L parts with a VED of

56.67 J/mm3. Strong <101> and <001> oriented textures were visible from the texture IPFs. There

is a possibility that this additional but undesirable strong <001> texture somehow contributed to

the mitigation on the micro-mechanical properties. In general, parts with microstructures of the

finer grains are supposed to exhibit superior mechanical properties. The average grain size of L-

PBF AISI parts with VED of 45.24 J/mm3 was calculated to be the lowest among all L-PBF parts

at 9.22 μm and unexpectedly, those parts did not display the best micro-mechanical responses. It

also requires further investigations in future studies.

In summary, micro-mechanical anisotropy was found in all the L-PBF AISI 316L parts

regardless of the amount of volumetric energy density and the extent of anisotropy was addressed

65
by the grain morphology and the quantitative means of the analysis of data, generated by the

electron backscattered diffraction maps such as texture PFs, texture IPFs, misorientation angles,

and grain size. In-depth cell size analysis on different grain, morphologies should be evaluated

further, to fully correlate and compare their micro-mechanical characteristics with the help of

nanoindentation.

The combined study of nanoindentation with EBSD mapping is an alternative methodology

for enriching qualification protocols, of additively manufactured metallic components, in addition

to existing macro-mechanical evaluation via tensile testing. Future works should include additional

experimental procedures to correlate indentation response with macro-mechanical behavior and

porosity characterization.

3.2 Effects of Proton Irradiation on Nanoindentation Strain-Rate Sensitivity and


Microstructural Properties in L-PBF 17-4 PH Stainless Steels*

Laser powder bed fusion has been proven to be an effective manufacturing process to

fabricate 17-4 PH stainless steel parts with acceptable quality outputs. In this study, the effects of

1 MeV proton irradiation (with a fluence of 1×1019 ions/cm2) damage of depth 5 μm on the surface

of L-PBF 17-4 PH steels are investigated in terms of micro-mechanical properties, microstructure,

and strain-rate sensitivity. Strain-rate sensitivity of as-built (unirradiated) and irradiated 17-4 PH

stainless steel were studied in different strain rates of 0.01, 0.05, 0.1, 0.25, 0.5, 0.75, and 1 s-1, and

for each strain rate 100 indentations were made. Micro-mechanical properties, yield, and

maximum shear strength of proton-irradiated L-PBF 17-4 stainless steel parts were substantially

affected by means of a significant change in strain rates, concretely at strain rates higher than 0.25

* This section is reproduced from M.J. Uddin, H.R. Siller, R.A. Mirshams, T.A. Byers, B. Rout, Effects of proton
irradiation on nanoindentation strain-rate sensitivity and microstructural properties in L-PBF 17–4 PH stainless
steels, Mater. Sci. Eng. A. 837 (2022) 142719. https://doi.org/10.1016/j.msea.2022.142719, with permission from
Elsevier.

66
s-1 when compared to as-built parts at two different volumetric energy densities. These final

findings are relevant due to the applications of these alloys in the fabrication of components

subjected to irradiation and extreme working conditions and are factors to consider when

evaluating the use of additive manufacturing to fabricate intricate nuclear parts.

3.2.1 X-ray Diffraction (XRD) Analysis

XRD analysis confirms the presence of mostly martensitic phase both in as-built and

irradiated 17-4 PH stainless steel parts. XRD X-ray could not detect the austenitic phase on the

materials and it could not be able to distinguish between the body-centered cubic (BCC) ferrite

phase and body-centered tetragonal (BCT) martensite phase since low carbon (≤ 0.02 wt.%)

stainless steels exhibit a very small magnitude of lattice distortion in BCT martensitic phase [137].

Three prominent peaks have been found at 2θ angles of 44.341, 64.508, and 81.630 degrees as

shown in Figure 23, and XRD data agree with the literature found regarding additively

manufactured 17-4 PH stainless steels [138–143].

Figure 23: X-ray diffraction measurements of L-PBF 17-4 PH as-built (AB) and irradiated (IR)
stainless steel parts.

67
3.2.2 Micro-Mechanical Properties by Nanoindentation

Nanoindentation hardness and modulus data are analyzed for as-built and irradiated L-PBF

17-4 PH stainless steel parts fabricated in two different volumetric energy densities (i.e. 54.76

J/mm3 and 61.11 J/mm3). In this study, the focus has been given mainly on the analysis of

indentation hardness data since strain-rate sensitivity (SRS) is directly related to hardness data

from nanoindentation. Variations in hardness data due to proton irradiation against different strain

rates are shown in Figure 24 and effects of irradiation on indentation modulus data are shown in

Figure 25. Statistical one-way ANOVA analysis has been performed by Origin® and primary

important statistical data are enlisted in Table 11 and Table 12. To find the strain-rate sensitivity,

two additional strain rates have been selected for both irradiated parts within the designated

irradiated area. The easiest features on hardness vs. strain rates plots in the case of both L-PBF 17-

4 PH stainless steel parts are increment of hardness with higher strain rates and irradiation shift of

hardness level within the irradiation zones. Furthermore, irradiated parts display higher reading of

the indentation hardness with increasing strain rates. Different mechanisms of dislocations play

important role in the plastic deformation of irradiated stainless steels under nanoindentation

experiment. The gliding of dislocations in plastically deformed regions could be affected by

defects created by irradiation: DLs (dislocation loops), SSDs (statistically stored dislocations),

GNDs (geometrically necessary dislocations), solute atoms, lattice friction, and clusters of defects

made of alloy atoms [144]. In irradiated stainless steels, there are three hardening mechanisms

namely DL hardening, dislocation hardening, and intrinsic hardening. Hardness can also vary

under various irradiation doses. Lower doses result in a lower reading of hardness for the same

material, which is supposed to be higher in extent when compared to the same as-built material

[145,146]. The unique solidification process involved with laser powder bed fusion of 17-4 PH

68
stainless steels results in anisotropy in microstructure and micro-mechanical properties

[71,97,112,140–149], which agrees with the results found in this study. Mean hardness and

indentation modulus is observed to vary for each strain rate for both L-PBF as-built and irradiated

17-4 PH stainless steel parts.

Figure 24: Nanoindentation hardness plot against different strain rates (histogram) for as-built and
irradiated L-PBF (54.76 J/mm3 and 61.11 J/mm3) 17-4 PH steel parts.

It is observed from the hardness plot in different strain rates of 17-4 PH steel parts that at

higher strain rates the parts are more sensitive and hardness values rise rapidly for both irradiated

parts, especially after 0.5 s-1 strain rate. For example, in L-PBF 17-4 PH steel parts with VED of

54.76 J/mm3, the mean hardness jumps 137.2% to 175.6% (4.27±1.09 GPa to 10.74±1.98 GPa)

and 72.4% to 102.2% (5.21±0.75 GPa to 10±2.19 GPa) for VED 61.11 J/mm3 parts in strain rate

of 1 s-1 due to irradiation hardening effects. Whereas, at a low strain rate of 0.01 s-1, mean hardness

increases 90% to 120% (3.14±0.84 GPa to 6.31±1.26 GPa) for VED 54.76 J/mm3parts and 70.6%

to 132.5% (3.42±0.91 GPa to 6.62±0.78 GPa) for VED 61.11 J/mm3 parts.

69
Table 11: Statistical One-way ANOVA of hardness data

Strain Standard SE of R- Coeff. of


Mean F-vale Prob>F
rate (s-1) Deviation Mean square variance
0.01 3.140 0.840 0.091
VED = 54.76 0.05 3.481 0.828 0.092
J/mm3 (as- 0.1 3.622 0.476 0.062 0.185 0.232 21.260 8.19E-16
built) 0.5 4.013 0.873 0.096
1 4.270 1.092 0.131
0.01 6.310 1.259 0.127
0.05 6.541 0.923 0.095
VED = 54.76 0.1 6.606 0.852 0.086
J/mm3 0.25 6.585 0.678 0.083 0.647 0.151 197.140 2.83E-142
(irradiated) 0.5 6.984 0.622 0.064
0.75 8.548 0.903 0.090
1 10.740 1.981 0.199
0.01 3.423 0.912 0.092
VED = 61.11 0.05 4.014 0.670 0.071
J/mm3 (as- 0.1 4.095 0.614 0.064 0.457 0.170 86.048 5.56E-53
built) 0.5 4.975 0.593 0.071
1 5.281 0.747 0.092
0.01 6.616 0.779 0.080
0.05 6.625 0.531 0.054
VED = 61.11 0.1 6.824 0.641 0.065
J/mm3 0.25 6.766 0.536 0.056 0.536 0.147 124.487 2.51E-104
(irradiated) 0.5 7.392 0.998 0.110
0.75 8.165 0.955 0.100
1 10.002 2.185 0.220

The one-way ANOVA model fits moderately well for all the hardness data of as-built and

irradiated parts except as-built 54.76 J/mm3 due to a lower value of the coefficient of determination

(R2) at a significance level of α = 0.05. Additionally, a sufficiently large F-value or P-value smaller

than the significance value of α = 0.05 means the model is significant for all as-built and irradiated

L-PBF 17-4 PH stainless steel parts.

Indentation modulus plot as in Figure 25 that is plotted against strain rates exhibit the

reverse behavior of hardness plot as in Figure 24. Higher strain rates influence heavily modulus

values in all as-built and irradiated L-PBF 17-4 PH stainless steel parts. At a small loading rate

70
that impact is not very noteworthy. Interestingly, irradiation enhances and increases indentation

modulus in all cases. As an example, the mean indentation modulus for the as-built VED 54.76

J/mm3 part measured with 0.01 s-1 strain rate is calculated to be 180.04±31.36 GPa, and the same

experimental setup yields a mean of 217.36±31.74 GPa in the irradiated part. In the case of L-PBF

VED 61.11 J/mm3 17-4 PH stainless steel, indentation modulus is determined as 190.60±33.59

GPa in the as-built part and 231±19.03 GPa in the irradiated part. At a strain rate of 1 s-1, ANOVA

analysis (Table 12) generates a modulus mean of 100.08±20.26 GPa for as-built and 128.02±26.76

GPa for irradiated 54.76 J/mm3 part. Very high values of F-values indicate the models are

significant in all samples, and consequently, P-values are smaller than the significance level of

=0.05 which is an indicator that the influence of strain rates on indentation modulus is

statistically significant at different VED values.

Figure 25: Nanoindentation modulus plot against different strain rates (mean plot and standard
deviation as error) for (a) as-built and (b) irradiated L-PBF 54.76 J/mm3 and 61.11 J/mm3 17-4 PH
steel parts.

71
Table 12: Statistical One-way ANOVA of indentation modulus data

Strain Standard SE of R- Coeff. of


Mean F-vale Prob>F
rate (s-1) Deviation Mean square variance
0.01 180.044 31.362 3.382
VED = 54.76 0.05 177.786 22.690 2.521
J/mm3 (as- 0.1 198.025 16.462 2.143 0.654 0.143 176.88 7.12E-85
built) 0.5 163.907 21.615 2.373
1 100.077 20.261 2.422
0.01 217.363 31.740 3.206
0.05 219.637 21.722 2.229
VED = 54.76 0.1 221.708 21.236 2.145
J/mm3 0.25 212.887 15.846 1.936 0.694 0.124 244.35 2.15E-162
(irradiated) 0.5 189.077 19.768 2.028
0.75 152.066 22.276 2.228
1 128.016 26.765 2.690
0.01 190.594 33.585 3.393
VED = 61.11 0.05 210.503 25.298 2.667
J/mm3 (as- 0.1 210.101 20.339 2.132 0.581 0.157 141.90 5.97E-76
built) 0.5 165.690 35.144 4.231
1 114.539 28.141 3.464
0.01 231.001 19.026 1.942
0.05 233.129 13.979 1.419
VED = 61.11 0.1 230.441 13.616 1.390
J/mm3 0.25 230.224 15.862 1.663 0.798 0.102 424.49 2.38E-220
(irradiated) 0.5 195.948 23.005 2.541
0.75 158.889 26.906 2.805
1 125.796 26.703 2.684

The influence of higher strain rates on the modulus values of L-PBF 17-4 stainless steel

parts are most probably due to a few factors. Firstly, anisotropy in mechanical property behavior

of ductile BCC crystal structures of stainless steels such as 17-4 PH (0.07% Carbon), and 316L

(0.03% Carbon). Secondly, higher strain rates facilitate to arise an effect which is known as ‘Pile-

up effect’ [150]. Nanoindenter channel which computes the indenter contact area on the sample

surface under predicts the true contact area when pile-up begins and therefore, it is plausible that

the channel over predicts the indentation modulus values. In instrumented indentation techniques,

the contact area is calculated indirectly as a function of modulus and stiffness:

72
𝜋𝜋𝑆𝑆 2
𝐴𝐴𝐸𝐸 =
4𝐸𝐸𝑅𝑅2

where AE = contact area, S = stiffness, and Er = reduced modulus of elasticity

Thirdly, the viscoelasticity phenomenon drives a material to behave like a liquid as the

velocity of displacement raises. At lower strain rates, L-PBF 17-4 PH stainless steel parts have

enough time to stretch with a lower indentation load (around 200 mN) before reaching maximum

indentation depth (in this case 1500 nm), whereas, at higher stain rates, it requires higher

indentation load (more than 300 mN) because of insufficient time to make the indent on the surface

of the parts.

3.2.3 EBSD Microstructural Characterization

The microstructural study of L-PBF 17-4 PH stainless steel parts has been performed and

reported in many recent open literature. However, no study reported the microstructural

characterization on irradiated 17-4 PH parts so far. Directional and rapid cooling characteristics

make the L-PBF a unique and non-equilibrium solidification process with the presence of a very

high-temperature gradient. Rapid cooling rate such as 105-106 K/s makes the equilibrium phase

diagram invalid and pre-solidified metals undergo cyclic heating and cooling during the L-PBF

printing process resulting in non-equilibrium microstructural characteristics in 17-4 PH stainless

steel parts [151]. According to the equilibrium phase diagram of 17-4 PH stainless steel, during

cooling from melted steel at a high temperature above the liquidus line, the ferrite (BCC) phase

forms first and with further cooling, the austenite (FCC) phase starts to form through pretectic

reaction of high-temperature ferrite and present liquid phase. Additional solidification at a rapid

cooling rate initiates the formation of martensitic (BCC) transformation and the martensite phase

continues to form [152]. The microstructure and texture of the as-built L-PBF 17-4 PH stainless

73
steel parts rely on the powder particle’s initial chemical composition [140]. In the second stage,

austenite transforms to martensite as a melted layer of 17-4 PH powder particles cool down, but,

not all of the austenite phase can transform into martensite as the solidification process continues.

Thus, L-PBF 17-4 PH stainless steel part contains retained austenite (FCC and metastable phase

at room temperature) grains are found to be extremely small and this phenomenon inhibits the

kinetics of the phase transformation [112,153,154]. This presence of retained austenite is also

suspected to possibly inhibit the precipitation of copper-rich particles during the aging process that

hardens and strengthens the alloy [155]. Processing atmosphere conditions are also noted to affect

the produced alloy phase composition. Both nitrogen and argon are commonly used as

atmospheres for producing powders via gas atomization and shielding gases in LPBF processing

of 17-4 powders. In this present study, Argon gas has been used as a built atmosphere and produces

a mostly martensitic microstructure, which is consistent with the literature [139,156]. The

martensitic body-centered tetragonal (BCT) phase is problematic to distinguish in EBSD scanning

due to the c/a ratio close to unity but, it is quite easier to identify the face-centered cubic (FCC)

austenite phase. That is the reason, martensite is referred to as BCC/BCT in inverse pole figure

(IPF) in Figure 26 to Figure 29 [156,157]. It is obligatory to be noted that the BCC phase can be

referred to as either the martensite or ferrite, since these two phases are not possible to distinguish

by the XRD as well as the EBSD technique. The martensite in stainless steels with carbon content

less than 0.2%, which is the case of 17-4 PH steel (~0.07 % C), has a BCC structure similar to

ferrite.

Similar to literature, equiaxed austenite (FCC) grains are found to be very fine compared

to coarse martensite grains and located mostly near grain and melt pool boundaries where the

atomic arrangement is more irregular and martensite can no longer grow [158]. Elongated grains

74
are found in all samples and these grains tend to grow from melt pool boundaries with a favored

growth along <100> orientation as shown in Figure 26 to Figure 29. It is also clear that the area

where indents are made on the surface austenite phase reveals because indents penetrate around

1500 nm and expose more grain boundaries as shown in phase maps (Figure 26(b), Figure 27(b),

Figure 28(b), and Figure 29(b)). Texture pole figures do not show a notable difference in intensity

on any particular crystal orientation in as-built and irradiated 17-4 PH parts and show an extent of

anisotropy in microstructure as well as in micro-mechanical behavior similar to L-PBF 316L

stainless steel parts which have been reported previously [12].

Figure 26: EBSD scan of nanoindentation area (at 0.5 s-1 strain rate) of as-built L-PBF 17-4 PH
stainless steel part with VED = 54.76 J/mm3; (a) IPF map in the (X–Y) plane perpendicular to the
build direction, (b) phase map, (c) texture pole figure, (d) inverse texture pole figure, and (e) IPF
color maps of phases.

75
Figure 27: EBSD scan of nanoindentation area (at 1 s-1 strain rate) of irradiated L-PBF 17-4 PH
stainless steel part with VED = 54.76 J/mm3; (a) IPF map in the (X–Y) plane perpendicular to the
build direction, (b) phase map, (c) texture pole figure, (d) inverse texture pole figure, and (e) IPF
color maps of phases.

Figure 28: EBSD scan of nanoindentation area (at 0.5 s-1 strain rate) of as-built L-PBF 17-4 PH
stainless steel part with VED = 61.11 J/mm3; (a) IPF map in the (X–Y) plane perpendicular to the
build direction, (b) phase map, (c) texture pole figure, (d) inverse texture pole figure, and (e) IPF
color maps of phases.

76
Figure 29: EBSD scan of nanoindentation area (at 1 s-1 strain rate) of irradiated L-PBF 17-4 PH
stainless steel part with VED = 61.11 J/mm3; (a) IPF map in the (X–Y) plane perpendicular to the
build direction, (b) phase map, (c) texture pole figure, (d) inverse texture pole figure, and (e) IPF
color maps of phases.

Figure 26(c) to Figure 29(c) also represent texture pole figures corresponding to the (X–Y)

plane of as-built and irradiated L-PBF 17-4 PH stainless. A strong fiber texture is detected with a

<100> direction favorably aligned with the build direction Z for as-built parts and in irradiated

parts, strong texture along <100> and <111> directions is observed. This fiber texture is the

common solidification texture for the cubic crystals such as FCC and BCC metals. Even though

stainless steels (less than 0.2% C) generally undergo the phase transformation BCC (ferrite) to

FCC (austenite) to BCC (martensite), ensuing in a more random texture [159]. Hence, the strong

fiber texture observed contradicts the occurrence of this transformation sequence. On the contrary,

77
the microstructure revealed is the one obtained directly after solidification, with no further solid-

state phase transformation which validates that the ferrite (BCC) phase has bypassed the austenite

(FCC) phase during L-PBF thermal cycles [151].

3.2.4 Correlation of Martensite Fractions to Process Parameters of L-PBF 17-4 PH As-Built and
Irradiated Stainless Steel Parts

Several open literatures have reported that the major martensitic phase, as well as other

phases (austenite and ferrite), present on laser powder bed fusion, processed 17-4 PH stainless

steels parts; although reported percentages of martensite are not consistent. Andelle et.al. [160]

reported a dual-phase austenite-ferrite phase. The effect of hot isostatic pressing (HIP) on the

densification, mechanical properties, and microstructures using gas atomized powders has been

investigated and it is found that density increases from 90% to 97% [149]. At lower VED (64 and

80 J/mm3) like in this study, long-columnar and fine equiaxed grains are observed; around 95%

martensite is reported which contributes to high hardness. Another study published recently

reported, confirms to find a mixed microstructure of austenite and martensite and austenite phase

was detected mostly at the grain boundaries of the melt pool and martensite inside the melt pools

[161]. It was not possible to distinguish between ferrite (BCC) and martensite (BCT). Lashgari et.

al. [162] investigated the effects of scanning pattern, build orientation, and single vs. double scan

on the microstructure of L-PBF 17-4 PH stainless steel parts and found the volume fraction of

retained austenite varied from 0.7% to 4.4% and austenite phase appeared in the grain boundaries

due to the large strains on those regions, high dislocations density, and finer grain size. L-PBF 17-

4 PH parts with gas atomized powders are found to decrease in grain size with increasing VED

from 64 to 84 J/mm3 [163]. Parts that are built vertically oriented have been found to contain

around 97% martensite in another study [112].

Quantitative analysis data of EBSD microstructural characterization of L-PBF 17-4 PH

78
parts are presented in Table 13 below with corresponding to hatch distance and scanning speed

used during the printing process. It is seen from the table that irradiation could not change the

phases present on the microstructures significantly, but a slight decrease in average grain size. This

could be due to irradiation-induced twins and plasticity during proton irradiation [164]. These

phenomena helped to form strain-induced martensite and mechanical twins to accommodate

additional strains caused by irradiation. These mechanisms are dependent on the stacking fault

energy and this energy usually increases with decreasing in grain sizes until the phase of martensite

is stabilized. A small fraction of the microstructure is found to be an austenitic phase. In Figure

26(b) to Figure 29(b), green and red colors represent martensite and retained austenite phase

respectively on the phase map generated from EBSDs. The lower VED (54.76 J/mm3) as-built part

is observed to have only around 1.8% austenitic phase and mostly a little over 98% martensitic

phase as shown in Figure 26(b). Whereas, 0.9 to 1.9% austenite and 98.5 to 99.1% martensite have

been determined in as-built 17-4 PH steel parts with VED = 61.11 J/mm3 as illustrated in Figure

28(b).

Table 13: Quantitative analysis of EBSD scans of as-built and irradiated L-PBF 17-4 PH stainless
steel parts

Hatch Scanning Phase %


VED Grain Misorientation
Distance Speed
(J/mm3) Size (μm) Angle (o)
(μm) (mm/s) % %
martensite austenite
54.76 700 4.9-8.35 98.1-98.2 1.8-1.9 19.54-21.17
As-built 120
61.11 600 5.35-6.87 98.5-99.1 0.9-1.5 17.54-20.3
54.76 700 4.1-5.2 97.5-98.5 1.5-2.5 17.3-23.12
Irradiated 120
61.11 600 4.97-5.0 98.5-98.6 1.4-1.5 20.05-21.44

According to the literature data, Hsu et. al. [165], percentages of martensite present in AM

as-built 17-4 PH parts increase with a decrease in scanning speed and an increase in volumetric

energy density with the same hatch distance and this is also true for this study. In that study, around

79
91% and 92.5% of martensite are found in 17-4 PH parts with 55 J/mm3 and 60 J/mm3 volumetric

energy density respectively (hatch distance = 100 μm). In this study, 91.1-98.2% and 98.5-99.9%

of martensite have been found with a volumetric energy density of 54.76 J.mm3 and 61.11 J/mm3

(hatch distance = 120 μm). Therefore, energy density, scanning speed, and hatch distance are vital

to process parameters as far as the fabrication of alloys in the L-PBF process is concerned. To

validate the percentages of martensite that have been derived from EBSD quantitative analysis in

this study, an empirical relationship (Eq. 10) can be considered from the literature [112]. Fractions

of martensite of L-PBF 17-4 PH stainless steels determined by experiment and as per Eq.10 are

shown in Table 14 below.

h
Mf = 0.142 �E + 56 �50 − 1�� + 76 (Eq. 10)

J
�Mf = martensite fraction, E = energy density � � , and h = hatch distance (μm)�
mm3

Table 14: Percentage of martensite phase according to experiment and Eq. 8 for L-PBF 17-4 PH
stainless steels from open literatures

Vol. Energy Hatch


Experimental % % of Martensite
AM Alloy Density Distance References
of Martensite according to Eq. 7
(J/mm3) (μm)
50 100 91 91.05
60 100 92.5 92.47 [165]
160 50 98 98.72
54 100 71 91.62
L-PBF 17-4 [166]
61 100 74 92.61
PH SS
68 50 66.27 85.66 [167]
65 100 75 93.18 [140]
106.67 50 93 91.15 [112]
85.7 70 93.8 91.35 [142]
As-built L- 54.76 120 98.15 94.91
PBF 17-4 PH
SS 61.11 120 98.8 95.81
This study
Irradiated L- 54.76 120 98 94.91
PBF 17-4 PH
SS 61.11 120 98.55 95.81

80
3.2.5 Porosity Analysis

Reconstructed X-ray microscope (XRM) 3D scans for L-PBF 17-4 PH stainless steel parts

including the irradiated zone are shown in Figure 30 and Figure 31 for two different volumetric

energy densities. It is very clear from the scans that both of the parts do not have any

cracks/microcracks (microcracks having a dimension of fewer than 1 μm are not considered since

the voxel size was set at 1 μm) but possesses a small volume fraction as pores i.e. very high density

that is randomly distributed within the parts. This ensures the credibility and exceptionally good

printing ability of the additively manufacturing technique that has been used to fabricate alloys

such as 17-4 PH stainless steels. Micro-mechanical property and microstructural characterizations

data exhibit very good strength and uniform finer grains microstructures. Unfortunately, any effect

of proton irradiation could not be detected by XRM since the irradiation depth was very small (~5

μm), and setting pixel size less than 1 μm did not generate good scans.

Figure 30: X-ray microscope scan of L-PBF 17-4 PH stainless steel including irradiated zone
(volumetric energy density of 54.76 J/mm3); (a) 3D scan volume (reconstructed), (b) single slice
(ImageJ), and (c) 3D reconstruction of all 2401 slices showing pores or voids within the scanned
volume.

Pore % volume of 54.76 J/mm3 parts has been determined to be around 0.2755 %. A slice

of this part is shown in Figure 30(b) and random pore distribution has been observed as in Figure

30(c). For a slightly higher volumetric energy density of 61.11 J/mm3 17-4 PH parts, the pore %

81
volume has been calculated to be 0.1904 % (Figure 31). Therefore, it can be said that high density

L-PBF 17-4 PH stainless steel parts having no larger cracks are now easily manufacturable via

laser powder bed fusion.

Figure 31: X-ray microscope scan of L-PBF 17-4 PH stainless steel including irradiated zone
(volumetric energy density of 61.11 J/mm3); (a) 3D scan volume (reconstructed), (b) single slice
(ImageJ), and (c) 3D reconstruction of all 2401 slices showing pores or voids within the scanned
volume.

Table 15: Percentage of porosity determined by using different methods for L-PBF 17-4 PH stainless
steels extracted from open literature.

Volumetric
Measuring
AM alloy energy density % of Porosity Reference
technique
(J/mm3)

106.67 0.55-0.73 X-ray CT [112]


41 1.3
54 0.4 Archimedes [166]
61 0
L-PBF 17-4 68 < 0.0183 X-ray CT [167]
PH SS 72.8 0 X-ray Micro CT [168]
43.4 - 60.76 ≤1 X-ray Micro CT [169]
54.17 0.01 X-ray Micro CT [170]
54.76 0.2755 X-ray Microscope
This study
61.11 0.1904 (XRM)

A list of some recent open literature regarding the porosity studies of L-PBF 17-4 PH parts

is presented in Table 15. Gu et. al. [171] studied the porosity of L-PBF 17-4 PH stainless steels in

82
the Archimedes method and found that % porosity decreases as volumetric energy density

increases. Similar to this study, they calculated 0.4 % and 0 % porosity in 54 J/mm3 and 61 J/mm3

volumetric energy density 17-4 PH parts fabricated by the same laser powder bed fusion

technology.

3.2.6 Strain-Rate Sensitivity (SRS) and Effect of Proton Irradiation on the Hardness

Measuring local micro-mechanical properties at different strain rates is pivotal to

investigate deformation mechanisms in materials regardless of the manufacturing or fabrication

methods. Nanoindentation has been used for local micro-mechanical property measurements,

especially in strain rates not more than 0.1 s-1. The usage of high indentation strain rates proved to

be very helpful to investigate rate-dependent properties recently [172–175]. In the current study,

strain-rate dependency on as-built and proton irradiated L-PBF 17-4 PH stainless steel parts have

been investigated and at the same time, an effort has been made to correlate the effects of

irradiation on microstructure, porosity, and micro-mechanical properties.

Load-displacement curves are represented in Figure 32 for as-built and in Figure 33 for

irradiated 17-4 PH stainless steels under various strain rates and these curves are generated in

continuous stiffness measurement technique at room temperature. It is notable that before holding

the stage, the load on samples increases with increasing strain rates at the same displacement into

the surface (1500 nm), and this phenomenon confirms that there is an effect of strain rate on the

hardness of L-PBF as-built and irradiated17-4 PH stainless steels [176]. As observed from the

curves, applying high strain rates move the load-displacement curves upward for both samples.

Due to the irradiation, nanoindentation hardness increases with an increase in strain rates thus

requiring more load to make the indenter impression or to reach the depth limit of 1500 nm into

the surface. A load of around 200 mN is observed for the deformation required to reach that depth

83
for as-built parts, whereas, more than 300 mN load is required to do so for proton irradiated parts.

No scientific reference has been found on nanoindentation deformation behavior via load-

displacement curves for additively manufactured 17-4 PH steels.

Figure 32: Load-displacement curves (P-h) for as-built L-PBF 17-4 PH stainless steel in different
strain rates, (a) VED = 54.76 J/mm3 and (b) VED = 61.11 J/mm3.

Figure 33: Load-displacement curves (P-h) for irradiated L-PBF 17-4 PH stainless steel in different
strain rates, (a) VED = 54.76 J/mm3 and (b) VED = 61.11 J/mm3.

To measure the strain-rate sensitivity (SRS, m), hardness data of 100 indents (10 by 10

arrays as shown in EBSD maps) was considered in each strain rate. To determine m, the plot as

shown in Figure 34 is plotted in a log-log scale according to the power-law relationship (Eq. 4).

84
Figure 34: Strain-rate sensitivity (SRS) determined by plotting the log-log scale of average hardness
and strain rate values and plots are drawn considering the density of the parts (normalized); (a) SRS
in as-built and (b) SRS in irradiated L-PBF 17-4 PH steels parts.

Table 16: Strain-rate of L-PBF 17-4 PH stainless steel using different approaches
Strain rates range
Strain-rate sensitivity (m) Approach
(s-1)
0.01 - 10 0.0199 - 0.0578 Tension test [175]
0.001 - 1000 0.0084 - 0.0183 Tension test [177]
0.072 - 0.094 (as-built) Nanoindentation
0.01 - 1
0.0086 - 0.597 (irradiated) [this study]

The hardness values on Y-axis are normalized by considering the actual density of L-PBF 17-4

PH stainless steel parts determined by XRM. To derive a second pattern for irradiated parts (linear

trend line is not observed similar to as-built part), additional two strain rates have been selected

(0.25 s-1 and 0.75 s-1). The strain-rate sensitivity, m, values for as-built 54.76 J/mm3 and 61.11

J/mm3 AM 17-4 PH stainless steel parts are calculated to be 0.072 and 0.094 respectively.

Irradiated 17-4 PH steel parts become harder (as per hardness values by nanoindentation) and m

values are calculated to be 0.014 for 54.76 J/mm3 parts and 0.0086 for 61.11 J/mm3 parts in lower

strain rates (0.01 s-1 to 0.25 s-1). In higher strain rates (0.5 s-1 to 1 s-1), m is calculated to be 0.597

for 54.76 J/mm3 and 0.424 for 61.11 J/mm3 L-PBF 17-4 PH parts. A couple of scientific studies

have been found on strain-rate sensitivity of L-PBF 17-4 PH stainless steel and the authors

85
investigated it in terms of tensile studies (not via instrumented indentation approach) [152,175]

and Table 16 represents comparative m values from those studies for same alloys fabricated by the

same technique.

3.2.7 Maximum Shear and Yield Strength Approximation

The maximum shear strength of as-built and irradiated L-PBF 17-4 parts have been

determined from the nanoindentation approach using the first displacement bursts methodology.

The first pop in the phenomenon is where the elastic to elastic-plastic deformation transition starts.

The P-h curves that are displayed in Figure 32 and Figure 33 are supposed to be fully elastic before

the pop-ins. The shift of pop-in event results with an increase in loading rate (i.e. increment in load

with time) and release of local stress in load-displacement curves [178]. These pop-ins are

observed when loading is intimately associated with the nucleation of dislocation sources [179].

Primary displacement bursts are more significant compared to later bursts, independent of strain

rates, and loads associated with these are used to analyze maximum shear strength in instrumented

indentation; hence, are more appropriate to investigate and explore dislocation activity with only

selecting a low strain rate nanoindentation experiments [108,180]. In Figure 32(a), displacement

bursts for as-built and irradiated 17-4 PH parts are shown extracted from the very initial portion

of load-displacement curves from Figure 32 and Figure 33 at 0.01 s-1 strain rate and impact of

irradiation on maximum shear and yield strength of 17-4 PH parts are shown in Figure 35(b). Table

17 is representing the calculated maximum shear strength of as-built and irradiated parts at the

lowest strain rate nanoindentation experiments according to Eq. 7 and Eq. 8 and Table 18 shows

the yield strength values of those parts and calculated according to Eq. 9 and Eq.10.

It is noteworthy to state that if the primary initial portion of the load-displacement curves

is plotted like Figure 35(a), some discrepancies are observed due to the course of elastoplastic

86
deformation in different random penetration depths into the surface of 17-4 PH parts by revealing

multiple displacement bursts in each curve as shown in highlighted ellipses [179]. This

phenomenon is intimately related to the plastic deformation mechanism (i.e. dislocation nucleation

and propagation) and each displacement burst takes place at a different load (i.e. stress)

contributing to increasing the maximum indentation load. Besides, as-built parts exhibit mini

displacement (in black and red ellipse regions) bursts, and irradiated parts display comparatively

bigger displacement bursts (blue and green ellipse regions). Minor bursts affirm that nucleated

dislocations suppress dislocation activity happening due to barriers to the movement of

dislocations [180]. Therefore, minor displacement pop-ins in irradiated parts refer to more

resistance to plastic deformation and resulted in comparatively higher maximum shear strength.

Figure 35: (a) First displacement pop-ins in load-displacement curves for as-built and irradiated L-
PBF 17-4 PH stainless steel parts and (b) effect of irradiation on maximum shear strength at 0.1 s-1
strain rate nanoindentation tests.

Table 17: Maximum shear strength of as-built and irradiated L-PBF 17-4 PH steel parts at 0.01s-1
strain rate with loads associated with first displacement bursts.
Modulus of Reduced Max. shear Shear
First displacement
Samples elasticity modulus strength modulus
pop-in load (mN)
(GPa) (GPa) (GPa) (GPa)
54.76 J/mm3 (As-built) 0.01685 ± 0.002 10.43
54.76 J/mm3 (Irradiated) 0.01825 ± 0.001 10.71
187.3 [112] 170 7.2
61.11 J/mm3 (As-built) 0.02023 ± 0.004 11.09
61.11 J/mm3 (Irradiated) 0.02440 ± 0.003 11.8

87
Table 18: Yield strength of as-built and irradiated L-PBF 17-4 PH steel parts at 0.01s-1 strain rate.

Volumetric energy Average Vickers hardness Yield strength


density (VED) nanohardness (GPa) (kg/mm2) (MPa)

54.76 (As-built) 3.14 296.71 907.95


54.76 (Irradiated) 6.31 596.29 1824.65
61.11 (As-built) 3.42 323.41 989.65
61.11 (Irradiated) 6.62 625.20 1913.12

Figure 36: Hardness-displacement curves of as-built and irradiated 54.76 J/mm3 17-4 PH stainless
steel parts in different strain rates.

The calculated maximum shear strength from P-h curves from nanoindentation

experiments can be referred to as the theoretical shear strength of any alloy. As-built parts display

maximum shear strength as 10.43 GPa (54.76 J/mm3) and 11.09 GPa (61.11 J/mm3). As

mentioned before, irradiation effects slightly increase the shear strength of 17-4 parts. Modulus of

elasticity of L-PBF 17-4 PH stainless steel has been taken from recent literature using the same

additive manufacturing techniques as this study [112]. Irradiation affects heavily on the yield

strength of the stainless steel and Dolph et. al. [181] reported that this increase is attributed mostly

to the nucleation and growth of dislocation loops and pored under irradiation for iron-chromium

alloys.

88
The resulting hardness is plotted against the indentation depth into the L-PBF 17-4 PH SS

parts (VED = 54.76 J/mm3), which is defined and determined continuously during each separate

indentation experiment in Figure 36. Values of hardness are observed as shown in the plots to vary

very little after 400 nm displacement into the surface of the parts. This last statement is so true for

lower strain rates experiments; whereas, for increasing higher strain rates the hardness values are

stabilized later at 600 nm or 800 nm depth. These phenomena are true for 17-4 PH parts before

and after the proton irradiation. Hardness values are found to increase almost by a factor of 2 due

to the irradiation for each strain rate nanoindentation experiment. As for the highest strain rate (1

s-1), the hardness increases continuously up to 400-500 nm depth into the parts and stabilizes

within a range of hardness values. The resulting average hardness values for the parts with VED

54.76 J/mm3 are 4.27 GPa before the irradiation and 10.62 GPa after the irradiation for a strain

rate of 1 s-1. One of the reasons for this sudden jump in hardness values is irradiation damages by

proton irradiations which in turn work as obstacles for dislocations movement or sliding as

discussed in an earlier section [182].

Figure 37: Summary of micro-mechanical properties, quantitative analysis of EBSD maps, and
maximum shear strength of as-built (AB) and irradiated (IR) L-PBF 17-4 PH stainless steel parts
(based on EBSD scans performed on 1 s-1 strain rate nanoindentation areas).

89
To summarize, Figure 37 shows the change in hardness, indentation modulus, grain size,

and maximum shear stress due to proton irradiation from as-built additively manufactured 17-4

PH stainless steel parts in one plot. Based on EBSD microstructural characterization, average grain

size (μm) decreases because of proton irradiation in both 54.76 J/mm3 (8.35 μm to 4.1 μm) and

61.11 J/mm3 (5.35 μm to 5 μm) VED parts. Nanohardness, indentation modulus, and maximum

shear stress are found to increase with the increase in VED for as-built parts and irradiation shifts

those micro-mechanical properties further upward as the material becomes harder and more

sensitive to strain rate.

3.2.8 Conclusions of Irradiation Effects on Mechanical Properties and Microstructure

Firstly, this study validates that the additive manufacturing (laser powder bed fusion, L-

PBF) process is capable of printing 17-4 PH stainless steels very effectively and with 100%

integrity with less than 0.7% porosity (based on the map of X-ray tomography) and very fine grain

sizes of martensite.

Secondly, L-PBF 17-4 PH stainless steel parts are very sensitive to high strain rates and

irradiation further escalates the sensitivity notably after a strain rate of 0.25 s-1. Irradiated parts

made with 54.76 J/mm3 VED display a drastic rise in strain-rate sensitivity (m) from 0.014 to 0.597

and 0.0086 to 0.424 in the case of 61.11 J/mm3 parts inputting a maximum strain rate of 1 s-1.

Thirdly, experimentally higher martensite percentage (97.5 - 99.1%) has been found than

theoretically calculated (94.91 - 95.81%) and with constant hatch distance, higher VED contributes

to an increase in martensite percentage.

Fourthly, irradiation affects the average grain size of L-PBF 17-4 PH stainless steels and it

makes slightly finer grains when compared to as-built 17-4 PH stainless steels. It also makes the

90
alloy harder in term of nanohardness and contribute to an increment of the maximum shear strength

and affects the yield strength.

3.3 Small-Scale Mechanical Testing of Proton Irradiated L-PBF 17-4 PH Stainless Steels

Assessing the stress-stress characteristics of micro-scale volumes has been a challenge in

the field of engineering. The ability of nanoindenter with submicron level displacement and

accurate microscope-indenter calibration makes it easier for quite so many years. Small-scale

mechanical testing procedures have the capabilities to assess proficiently the effect of radiation in

micron-level structures (such as micropillars fabricated by focused ion beam (FIB)) on irradiated

L-PBF nuclear stainless steels such as 316L, 17-4 PH. This study provides uniaxial mechanical

responses of low volumes of as-built (AM) and irradiated 17-4 PH stainless steels and direct

correlation of the load-displacement i.e. stress-strain data to individual plastic deformation events

in terms of tension properties with the relation of micro-mechanical property responses generated

by the same nanoindenter with a different set-up.

3.3.1 Micro-Compression Testing, Observation of Micropillar Fractures, and Nanoindentation


Response

Micro-compression testing has gained momentum right after the introduction of the

focused ion beam (FIB) in the research industry as FIB sample manufacturing opened a new

window with its capability of fabricating well-defined geometries and low volume samples in sub-

micron level to any type of metallic alloys where hardness of that material is not a concerning

factor. Those low-volume micro-structures can lead to more sophisticated nano/micro-mechanical

behaviors if a proper experimental procedure is chosen and investigated systematically. FIB

processing techniques have minimized sample preparation effects of previously available sample

preparation techniques with the application of a controlled low-energy ion beam.

91
In this study, pillars are made with an aspect ratio of 2-2.50 (slight tapper at the bottom

which is considered while doing the calculation and making plots) and different diameters of 2

μm, 3 μm, and 5 μm. 2 μm micropillars are made only in an irradiated area so that the height falls

within the radiation damage depth of 5 μm. The micro-compression tests are done at a strain rate

of 0.1 s-1 and compressed micropillars are confirmed by observing them on the FIB followed by

load-displacement data extraction from the nanoindenter which would be converted to stress-strain

values, and thus, analysis of converted stress-strain curves would provide tension properties (such

as compressive yield strength,

Micropillars with different sizes on as-built AM 17-4 PH stainless steel parts (sample 1

with the volumetric energy density of 54.76 J/mm3 and sample 2 with 61.1 J/mm3 in Table 6) are

shown in Figure 38 and Figure 39. Micropillars with different sizes on irradiated AM 17-4 PH

stainless steel parts (sample 1 with the volumetric energy density of 54.76 J/mm3 and sample 2

with 61.1 J/mm3 in Table 6) are shown in Figure 40 and Figure 41.

Figure 38: Initial (top row) and after micro-compression (bottom row) micropillars of as-built AM
17-4 PH stainless steels with the volumetric energy density of 54.76 J/mm3 (sample 1); (a1) and (a2)
3 μm pillar, and (b1) and (b2) 5 μm pillar.

92
Figure 39: Initial (top row) and after micro-compression (bottom row) micropillars of as-built AM
17-4 PH stainless steels with the volumetric energy density of 61.11 J/mm3 (sample 2); (a1) and (a2)
3 μm pillar, and (b1) and (b2) 5 μm pillar.

Figure 40: Initial (top row) and after micro-compression (bottom row) micropillars of irradiated AM
17-4 PH stainless steels with the volumetric energy density of 54.76 J/mm3 (sample 1); (a1) and (a2)
2 μm pillar, (b1) and (b2) 3 μm pillar , and (c1) and (c2) 5 μm pillar.

93
Figure 41: Initial (top row) and after micro-compression (bottom row) micropillars of irradiated AM
17-4 PH stainless steels with the volumetric energy density of 61.11 J/mm3 (sample 2); (a1) and (a2)
2 μm pillar, (b1) and (b2) 3 μm pillar, and (c1) and (c2) 5 μm pillar.

When compared with the micro/nanoindentation test, micro-compression has the

noticeable lead of a relatively uniform stress/strain field. However, since the micro-compression

test is almost a couple of decades older technology to measure the strain-strain relation, there is no

standard yet. Some experimental variables may affect accurate measurements of strain and stress

[183]. These variables are the aspect ratio (the ratio of height h and diameter d of any micropillar),

dimensions of the substrate below the pillar, taper angle θ (>0) (the angle between the tangent of

wall and axis of the pillar), fillet angle, misalignment between axis of the pillar and the direction

of uniaxial compression load, and lastly, the strength of the substrate. The aspect of some of these

variables can be instinctively understood. For example, with a little amount of substrate volume,

the pillar would sink upon compression and the major portion of the deformation will be acting

upon the substrate instead of the pillar, which would lead to erroneous measurement of the load-

displacement data collection. In this study, this phenomenon has been observed. This sink-in effect

94
may be magnified for pillars with a large aspect ratio and suppressed for pillars with a large taper

angle. Pillars with larger aspect ratios could undergo early buckling upon applying compression

load. This intuitive argument indicates that these variables may be coupled together to influence

the accuracy of the micro-compression experimental measurement.

Deformation in single-crystal micron-scale pillars under compression has been

characterized in depth in many past studies [184,185]. Whereas consistently varying flow curves

are anticipated for bulk deformation in polycrystalline metallic alloys, intermittency and stochastic

flow are the trademark behaviors exhibited in single-crystal micro compression. Specimens

undergo elastic or near-elastic loading zone which creates step-ups in micro-compression loading

curves. Both as-built and irradiated AM 17-4 PH stainless steel micropillars in this study exhibited

this behavior with many small slip steps intermittently forming throughout the pillar volume during

plastic deformation, especially in the early period of the micro-compression testing procedure.

With the introduction of proton irradiation damage, however, obstacles to dislocation motion in

the form of a dense population of nanometer-scale Frank loops induce hardening by impeding

dislocation motion [186]. Higher stresses are necessary to initiate plastic deformation, as

dislocations must overcome irradiation-induced obstacles for the slip to occur. In studies

examining polycrystalline ion-irradiated austenitic stainless steels, this has been observed to result

in the sudden formation of defect-free channels, in which dislocations clear out irradiation defects

in a narrow channel, creating an easy path for additional dislocations to glide along. The result is

heterogeneous plastic deformation, in which strain is localized in a series of channels spaced 0.001

mm apart in each grain [187]. The proton-irradiated micro-compression specimens of this study

showed similar localization of strain, with often only two to three major slip steps forming in each

pillar of different sizes in diameter (Figure 40 and Figure 41). During the compression of an

95
irradiated pillar, each slip step forms unexpectedly and conveys a large amount of strain. As the

pillar is deformed more, strain likes to stay within these pre-existing slip steps, rather than

introducing new ones. This ultimately leads to significantly larger step heights on the irradiated

pillar surfaces than observed on the unirradiated pillars.

Figure 42: Load-displacement curves of as-built and irradiated AM 17-4 PH stainless steel parts with
two different volumetric energy densities (sample 1 = VED of 54.76 J/mm3 and sample 2 = VED of
61.11 J/mm3).

Load-displacement curves of as-built and irradiated AM 17-4 PH stainless steel parts with

two different volumetric energy densities are illustrated in Figure 42. As it seems from the plot,

the bigger the micropillar gets, it requires more load to be fractured [188]. The observed

differences in slip behavior can be quantified by first tallying the number of load drops observed

96
in the load-displacement curve for each pillar. Tests in this study were performed in displacement-

controlled mode using a fixed displacement rate and data acquisition rate. Following the initiation

of a slip event, strain rates are typically so high that only one or two data points are collected during

this time, resulting in a flat plateau in the load curve connecting the sparse data. However, the

indenter itself is load-controlled, such that when a sudden deformation event occurs, a feedback

loop detects the change in displacement and adjusts the applied indenter force accordingly. For an

event that occurs more quickly than the feedback loop can respond, such as the formation of a new

slip step, this then manifests as a steep load drop in the curve. Thus, the number of load drops in a

given load-displacement curve can be taken as an estimate of the number of slip steps in the pillar.

Figure 43: Engineering stress-strain curves of sample 1 (VED = 54.76 J/mm3) with different sizes of
micropillars of L-PBF 17-4 PH stainless steels in (a) as-built condition, (b) irradiation condition, and
(c) comparison between as-built and irradiation conditions.

97
Figure 44: Engineering stress-strain curves of sample 2 (VED of 61.11 J/mm3) with different sizes of
micropillars of L-PBF 17-4 PH stainless steels in (a) as-built condition, (b) irradiation condition., and
(c) comparison between as-built and irradiation conditions.

It is important to note that since load-displacement curves would be converted into stress-

stress curves to analyze as well as more insight will be presented on this research work in the

future, however, additional adjustments are required to do so as the strain was not measured

directly in micro-compression testing [189]. Engineering stress-strain curves of 2 μm to 5 μm

micropillars micro-compression tests are shown in Figure 43 and Figure 44 for sample 1(VED =

54.76 J/mm3) and sample 2 (VED of 61.11 J/mm3) respectively. To convert the load-displacement

curves equations are taken from some literatures [190–192]. All the pillars are connected to the

substrate (AM 17-4 PH SS) hence total displacement would be the summation of displacement of

98
the micropillar, indenter, and the substrate, and since, the displacement of the pillar can only be

applicable while converting to stress-strain curve, elastic deformation due to indenter and the

substrate must be eliminated.

The effect of irradiation on the compressive yield and ultimate tensile strength is very

clearly visible in AM 17-4 PH stainless steel parts fabricated by 54.76 J/mm3 (Figure 42 and Figure

43) and both of these properties have been observed to increase due to the irradiation effect. 2 μm

micropillar in the irradiated condition of sample 1 exhibited the highest compressive strength with

~ 244.57 MPa yield (approximate) and 375.08 MPa tensile strength. The percentage increases in

compressive yield strength were 97.31% (107.27 MPa to 211.65 MPa), and 36.8% (150.70 MPa

to 206.16 MPa) for 3 μm, and 5 μm, micropillars respectively in sample 1 because of radiation

damage; whereas 47.01% (213.58 MPa to 317.00 MPa) and 10.86% (286.40 MPa to 317.50 MPa)

increments were observed in compressive tensile strength for 3 μm and 5 μm micropillars from as-

built to irradiated areas.

Data derived from the stress-strain curves of samples 1 and 2 are listed in Table 19. In both

the samples, compressive yield and tensile strength are affected by the irradiation damage as the

hardness of the material is impacted due to radiation-induced strain hardening as reported in our

previous study [14]. Micro-compression on pillars regardless of diameters on sample 1 exhibited

superior compressive yield and tensile strength compared to sample 2. Diameter of 5 μm pillars

on sample 2, yield strengths were found to be unchanged from as-built (133.62 MPa) to irradiated

condition (133.43 MPa), though tensile strength increased from 233.54 MPa to 250 MPa. All 2

μm pillars were fabricated as their height falls within the radiation damage depth of 5 μm. It was

expected to generate the highest yield strength and tensile strength due to the radiation hardening

effect as discussed earlier. Yield and tensile strength were found to be the highest as expected as

99
244.57 MPa and 375.08 MPa in irradiated 17-4 PH sample 1. The results listed in Table 19 are

showing the tendency of variations relating to indentation hardness values reported in our previous

publication [14]. In as-built conditions, AM 17-4 PH stainless steel sample 1 generated lower

values in nanohardness compared to sample 2, and yield strengths of micropillars built on sample

1 exhibited higher values. In irradiated conditions, indentation hardness values were highly

affected (maximum average hardness of 10.71 GPa) by strain hardening effect due to irradiation

in sample 1 compared to sample 2 (10.06 GPa). That characteristic was probably responsible for

the resulting higher values in compressive yield strengths in 17-4 parts made with a VED of 54.76

J/mm3. Additionally, finer grains on irradiated sample 1 (4.1-5.2 μm) because of irradiation

damage could play a part in increasing the values of that mechanical property than that of sample

2 (4.97-5 μm).

Table 19: Ultimate tensile strength and yield strength determined from engineering stress-strain
curves of all the pillars in as-built and irradiated conditions.

Yield
Micropillar
3 strength Ultimate tensile
Sample # VED (J/mm ) Conditions diameter
(MPa) strength (MPa)
(μm)
(Approx.)
3 107.27 213.58
As-built
5 150.70 286.40
3
1 54.76 J/mm 2 244.57 375.08
Irradiated 3 211.65 314.00
5 206.16 317.50
3 146.67 237.09
As-built
5 133.62 233.54
3
2 61.11 J/mm 2 171.20 271.64
Irradiated 3 150.88 212.75
5 133.43 250.00

In a couple of the relevant open literature for direct metal laser sintering (DMLS) processed

17-4 PH stainless steel as-built parts average compressive yield strength was calculated to be in

the range of 688.52 ± 36.31 MPa - using the same nanoindenter equipment used in this study at

100
room temperature but with a lower strain rate of 0.0001 s-1 (in this study 0.01 s-1 strain rate was

used) [91]. Micropillars were fabricated with a 1:3 aspect ratio and all the micropillars were 4 μm

in diameter. Their processed 17-4 PH stainless steels have consisted of 82% retained austenite,

whereas, our study suggested that our processed 17-4 PH steel parts were made of 94.91% to

98.8% martensite. In their continuous study with the same alloy but shot-peening, average yield

strength was calculated to be 914 ± 12.73 MPa, and average UTS was found as of 1130.1 ± 17.68

MPa in the as-built condition which is way higher than what is found in this study [99]. It has been

declared in these studies that exceptional work-hardening was observed due to the presence of

retained austenite. A comparison of those findings with our study has been shown in Table 20:

Table 20: Comparison with other AM 17-4 PH stainless steel micropillar compression testing
literature (LP = laser power (Watts), HD = hatch distance (μm), LT = layer thickness (μm), SS =
scanning speed (mm/s), UTS = Ultimate tensile strength)
LP, HD, Micropillar Aspect Strain Yield strength,
AM 17-4 PH UTS, MPa
LT, SS size, μm ratio rate, s-1 MPa
As-built [91] 688.52±36.31 ---
200, 110,
4 1:3 0.0001
As-built shot 40, 750
914±12.73 1130 ± 17.68
peened [99]
As-built [This 213.58 -
220-230, 107.27-150.70
study] 286.40
120, 50, 2 to 5 1: 2-2.5 0.01
Irradiated [This
600-700 133.43-244.57 212.75-375.08
study]

No literature has been found on irradiated AM 17-4 PH stainless steel parts till the last was

reported. Orientation-based micropillar study on AM 316L materials reported a range of yield

strength from 198 MPa to 451 MPa in as-manufactured and from 476 MPa to 578 MPa in irradiated

conditions (32 dpa) [100]. Another study for proton irradiation of conventional 304 stainless steel

informed average yield strength of 272 ± 34 MPa in as-built condition, 834 ± 39 MPa in 10 dpa

damage, and 1070 ± 14 MPa in 100 dpa damage regions [193].

The lower values in compressive yield and ultimate tensile strength of AM 17-4 PH

stainless steel in this study could correspond to quite a few issues regarding micro-compression

101
testing of micropillars by the MTS nanoindenter. First of all, the strain rate chosen in this study as

0.01 s-1 is quite higher than the literature references mentioned above, since micro-mechanical

responses were not found satisfactory when choosing a much lower strain rate where the

nanoindenter was displacement controlled rather than load controlled. As in our previous study of

AM 17-4 PH stainless steel, it has been shown that the parts fabricated are very strain-rate

sensitive. That could be the second reason for lower values of compressive yield and tensile

strength. Additionally, slip steps, and load-displacement curves are machine-dependent material

responses while interpreting load drops, hardening, and elastoplastic fracture properties [194]. In

micro-compression micropillar testing, usually, the load-displacement curve is converted to an

engineering stress-strain curve by normalizing the load by the average cross-sectional area of the

micropillars and the displacement of those pillars’ height. Three unanticipated observations were

made in the literature due to sample size effects; a significant increase in yield strength at smaller

volume pillars, a stochastic variation in yield stress between the same diameter pillars, and load

drops in displacement controlled test [115,116,195,196]. Moreover, misalignment of flat punch

with respect to the sample top surface area obviously could corrupt load-displacement i.e. stress-

strain data as well as the deformation mode, and could lead to lower values in calculated

compressive yield and ultimate tensile strength [197,198]. Few factors such as non-uniaxial

loading during micro-compression, change in pillar cross-section throughout the pillar height, only

pillar volume can contain plastic deformation of the pillars could also be responsible for

inconsistent or unexpected micro-mechanical responses as the presence of these factors cannot be

discarded in this study [194].

3.3.2 Conclusions of Irradiation Effects on Small-Scale Mechanical Testing of AM 17-4 PH


Stainless Steel

One of the aims of this study was to understand and compare the synergy between

102
indentation properties i.e. approximate yield strength and small-scale micro-compression yield

strength which are relatable to radiation hardening effects. A strong agreement has been observed

between the heterogeneous plastic deformation behavior i.e. localization of strain of as-built and

irradiated micropillars in terms of major slip bands occurrence for the dislocation movement.

Higher stresses were required for micropillars fabricated in irradiated areas than in as-built areas.

For an example, 2 μm micropillars fabricated in irradiation conditions absorbed the highest amount

of uniaxial compressive load to start yielding with a compressive yield strength of 244.57 MPa

and a compressive UTS of 375.08 as expected since the whole height of the pillars was well within

the depth of radiation damage hence accompanied with higher hardness values when measured

with nanoindentation due to strain hardening effects of radiation. Overall proton irradiation

influenced (increment) the measured micropillar compressive yield and UTS values in both AM

17-4 PH stainless steel parts. Sample size effects were clearly visible in Figure 42 and Figure 43

for all the micropillars fabricated in irradiation conditions in both the 17-4 PH parts. As the

micropillars get smaller, the strength was observed to increase with one exception.

103
CHAPTER 4

CONCLUSIONS AND FUTURE WORKS*

Additively manufactured nuclear alloys that are used as materials in this study are getting

more attention and finding newer applications as this manufacturing process develops

permanently. The comprehensive package of mechanical properties and evaluation methods

presented in this dissertation is a mandatory source of data for AM 17-4 PH and 316L stainless

steel’s qualification processes, product reliability, and overall process integrity. In sequence, the

additive metallic powder materials are analyzed, then appropriate methodologies have been

developed, and mechanical and microstructural properties are reported in this study.

The use of nanoindentation in combination with electron backscatter diffraction mapping

has proven to be very useful to characterize the anisotropy in micro-mechanical properties and

microstructure in additively manufactured parts made of AISI 316L alloy. After performing this

characterization methodology, the key findings can be summarized as follows:

L-PBF parts with VED 56.67 J/mm3 exhibited a similar fraction of <101> oriented grains

similar to other parts with different VEDs but, a clear appearance of grains oriented along the line

<001> to <111> (blue/blueish to light purple in IPF color map) was observed and that was one

distinctive feature among all the other L-PBF parts. These grains were responsible for generating

higher indentation property values. The indentation properties on <111> grain orientations were

enhanced in the nanoindentation experiment and generated a higher value of misorientation angles

as well. The evolution of certain grain orientations was noticeable and variations in micro-

* The first half of this chapter is reproduced from a combination of 2 sources, both used with permission from
Elseiver: (1) M.J. Uddin, E. Ramirez-Cedillo, R.A. Mirshams, H.R. Siller, Nanoindentation and electron backscatter
diffraction mapping in laser powder bed fusion of stainless steel 316L, Mater. Charact. 174 (2021) 111047,
https://doi.org/10.1016/j.matchar.2021.111047 and (2) M.J. Uddin, H.R. Siller, R.A. Mirshams, T.A. Byers, B.
Rout, Effects of proton irradiation on nanoindentation strain-rate sensitivity and microstructural properties in L-PBF
17–4 PH stainless steels, Mater. Sci. Eng. A. 837 (2022) 142719. https://doi.org/10.1016/j.msea.2022.142719.

104
mechanical responses were reported. The higher angle grain boundary regions most probably

behaved as a strong barrier against dislocation movements and thus, increments over the

indentation hardness and modulus were taken place. Dislocations pile-up occurred when those

regions have gone under plastic deformation and prevented crack propagation at a certain limit.

Collectively the relatively lower grain size accompanied by a higher average misorientation angle

of L-PBF AISI 316L parts with VED of 56.67 J/mm3 also influenced the calculation of hardness

and modulus values.

The conventional AISI 316L stainless steel was noticed to have the lowest texture strength,

but the indentation modulus and hardness were lower than most L-PBF AISI 316L parts with a

VED of 56.67 J/mm3. Strong <101> and <001> oriented textures were visible from the texture

IPFs. A strong possibility is there that this additional but unwanted strong <001> texture someway

subsidized the mitigation of the values of indentation properties. Generally, parts with

microstructures of the finer grains are supposed to reveal superior mechanical properties. The

average grain size of L-PBF AISI parts with VED of 45.24 J/mm3 was calculated to be the lowest

among all L-PBF parts as 9.22 μm and unexpectedly, those parts did not display the best micro-

mechanical responses. It also requires further investigations in future studies.

In summary, micro-mechanical anisotropy was found in all the L-PBF AISI 316L parts

regardless of the amount of volumetric energy density and the extent of anisotropy was addressed

by the grain morphology and the quantitative means of the analysis of data, generated by the

electron backscattered diffraction maps such as texture PFs, texture IPFs, misorientation angles,

and grain size. In-depth cell size analysis on different grain morphologies should be evaluated

further, to fully correlate and compare their micro-mechanical characteristics with the help of

nanoindentation.

105
The combined study of nanoindentation with EBSD mapping is an alternative methodology

for enriching qualification protocols, of additively manufactured metallic components, in addition

to existing macro-mechanical evaluation via tensile testing. Future works should include additional

experimental procedures to correlate indentation response with macro-mechanical behavior and

porosity characterization.

This study authenticates that AM process is fully capable of fabricating nuclear alloys such

as 17-4 PH stainless steels very efficiently and with almost 100% integrity that only contains less

than 0.7% porosity (based on the map of X-ray tomography) with very fine martensitic grains

confirmed by the XRD. L-PBF 17-4 PH stainless steel parts are found to be very sensitive to high

strain rates i.e. high load rates and the fact that irradiation damage additionally intensifies the

sensitivity limit notably after a strain rate of 0.25 s-1. A severe rise in strain-rate sensitivity (m)

from 0.014 to 0.597 was noticed in parts with irradiated conditions made with 54.76 J/mm3 VED

and an increment from 0.0086 to 0.424 was in the case of 61.11 J/mm3 parts where a maximum

strain rate of 1 s-1 was applied during nanoindentation experiment. Additionally, the experimental

percentage of martensite (97.5 - 99.1%) was found to be higher than theoretically calculated (94.91

- 95.81%) values. With the same process parameters of AM such as hatch distance, an increase in

VED influenced the martensite percentage in all the parts. Lastly, a slight effect of irradiation was

observed over average grain size calculation. It made the 17-4 PH stainless steel harder in terms

of nanohardness as shown in the analysis. The behavior at the micro-scale seen in nanoindentation,

shows significant departures from the classical elastic–plastic models, mainly due to the size effect.

The study of this phenomenon on additively manufactured radiation-resistant alloys has been

limited, and layer-by-layer nanoindentation size effects and strain-rate sensitivity, with an increase

in load, have scarcely been explored on small-scale tests geometries. We developed new models

106
of strain rate sensitivity in AM proton irradiated alloys (1 MeV proton irradiation with a fluence

of 1×1019 ions/cm2), where we found that stainless steel parts are very sensitive to high strain

rates and irradiation further escalates the sensitivity notably after strain rate of 0.25 s-1 in parts

made at different processing volumetric energy densities. The irradiation makes the alloy harder

in terms of nano-hardness and contributes to an increment of the maximum shear strength affecting

the yield and shear strength, as shown in our recent publication.

Table 21: Summary of major mechanical properties calculated from indentation and small-scale
mechanical testing (for 3 μm) methods on as-built and irradiated 17-4 PH stainless steels
As-built Irradiated As-built Irradiated
54.76 J/mm3 54.76 J/mm3 61.11 J/mm3 61.11 J/mm3
Indentation Hardness
3.14 6.31 3.4225 6.61
(GPa)
Indentation Modulus
180.044 217.363 190.59 231
(GPa)
Indentation Yield
907.95 1824.65 989.65 1913.12
Strength (MPa)
Indentation Maximum
10.43 10.71 11.09 11.8
Shear Strength (GPa)
Micropillar
Compressive Yield 107.27 211.65 146.67 237.09
Strength (MPa)
Micropillar
Compressive Ultimate
213.58 314 150.88 212.75
Tensile Strength
(MPa)

In micro-compression testing of AM 17-4 PH stainless steel alloys, it was found that

micropillars fabricated in as-built conditions moving dislocations (slip system) were the major

deformation carrier with the presence of a little amount of very fine fabrication-induced pores.

Strain-hardening due to proton irradiation obviously impacted the micropillar compressive

properties as the irradiated AM 17-4 PH stainless steels have high indentation hardness compared

to as-built conditions. Fewer slip bands were observed in irradiated micropillar under uniaxial

compressive load. A strong agreement was observed between micro-mechanical properties

107
determined in the indentation method than micro-compression micropillar method considering

both as-built and irradiated conditions.

The development of qualification protocols for the evaluation of additively manufactured

parts with irradiation effects is an alternative to other testing protocols requiring extensive

experimentation, from which the extraction of important mechanical properties can be achieved.

Table 21 summarizes the most significant mechanical properties estimated from this study, taking

into account the consistency of the data and the reliability of the testing instruments and devices.

For this table, the values extracted from compression testing of 3 μm micropillars and

nanoindentation samples with a strain rate of 0.01 s-1 were considered.

It can be noted that the collection of procedures will be determinant to establish a clear

correlation between volumetric energy densities and irradiation effects. For the case of indentation

hardness, the highest VED produced hardening effects, and the combination of irradiation gives a

significant change in this mechanical property. In the case of Indentation Modulus, Indentation

Yield, and Maximum Shear Strength the trend is similar. For Micropillar Compressive Yield the

result shows that the highest energy density produced better results as well and irradiation is

increasing those values interestingly. However the consistency among nanoindentation data and

micropillar compression is seen in some instances and not through the whole set of experiments,

but the irradiation effects are very clear in the whole experimentation, which is a strong argument

that the qualification procedure presented here is valid. For the Micropillar Compressive Tensile

Strength, the behavior is the opposite, the lower energy density produced better results. More

investigation is needed to evaluate this mechanical property with micro-tensile samples fabricated

by focused ion beam to integrate a more comprehensive protocol for future designing and

manufacturing of irradiation resistant additively manufactured alloys.

108
Indentation and small-scale mechanical test yield properties of AM 316L stainless steels

can be studied in the future as this study could facilitate the qualification processes for various

intricate applications as well as a direct comparison could be made with AM 17-4 PH stainless

steels. Strain-rate sensitivity studies of AM 316L parts could also be crucial for load-sensitive

applications in various industries. Increasing the number of experimentations of micropillar micro-

compression tests of AM 17-4 PH stainless steel is required for a more accurate assessment of the

micro-mechanical properties. Micro-compression could be performed at a very lower strain rate

(0.0005 or lower) in the future with a different testing approach using the same nanoindenter

equipment so that it would be well compared with the literature available. Fabricating micropillars

in grains with specific crystallographic orientations could help elaborate the plastic deformation

mechanism on a broad scale for both process optimization and performance evaluation of

additively manufactured stainless steel alloys for nuclear applications. Introducing TEM studies

along with mechanical and microstructural characterizations could help in determining Schmid

factors for any fabricated micropillars with reduced data variations in different grain orientations

which is critical to calculate critical resolved shear stress and yield stress more accurately. The

strength of various microstructural phases such as retained austenite and martensite could also be

computed quantitatively in future works. In-situ micro-compression tests on small volumes can be

carried out in the future to convey more capability and control on the testing procedure.

109
REFERENCES

[1] T.D. Ngo, A. Kashani, G. Imbalzano, K.T.Q. Nguyen, D. Hui, Additive manufacturing
(3D printing): A review of materials, methods, applications and challenges, Compos. Part
B Eng. 143 (2018) 172–196. https://doi.org/10.1016/j.compositesb.2018.02.012.

[2] L. Cheng, X. Liang, J. Bai, Q. Chen, J. Lemon, A. To, On utilizing topology optimization
to design support structure to prevent residual stress induced build failure in laser powder
bed metal additive manufacturing, Addit. Manuf. 27 (2019) 290–304.
https://doi.org/10.1016/j.addma.2019.03.001.

[3] O. Ivanova, C. Williams, T. Campbell, Additive manufacturing (AM) and


nanotechnology: promises and challenges, Rapid Prototyp. J. 19 (2013) 353–364.
https://doi.org/10.1108/RPJ-12-2011-0127.

[4] S. Ganesh Sarvankar, S.N. Yewale, Additive Manufacturing in Automobile Industry, Int.
J. Res. Aeronaut. Echanical Eng. 7 (2019) 1–10.

[5] D. Delgado Camacho, P. Clayton, W.J. O’Brien, C. Seepersad, M. Juenger, R. Ferron, S.


Salamone, Applications of additive manufacturing in the construction industry – A
forward-looking review, Autom. Constr. 89 (2018) 110–119.
https://doi.org/10.1016/j.autcon.2017.12.031.

[6] D. Chantzis, X. Liu, D.J. Politis, O. El Fakir, T.Y. Chua, Z. Shi, L. Wang, Review on
additive manufacturing of tooling for hot stamping, Int. J. Adv. Manuf. Technol. 109
(2020) 87–107. https://doi.org/10.1007/s00170-020-05622-1.

[7] Z. Liu, B. He, T. Lyu, Y. Zou, A Review on Additive Manufacturing of Titanium Alloys
for Aerospace Applications: Directed Energy Deposition and Beyond Ti-6Al-4V, Jom. 73
(2021) 1804–1818. https://doi.org/10.1007/s11837-021-04670-6.

[8] E.M. Parsons, Lightweight cellular metal composites with zero and tunable thermal
expansion enabled by ultrasonic additive manufacturing: Modeling, manufacturing, and
testing, Compos. Struct. 223 (2019) 110656.
https://doi.org/10.1016/j.compstruct.2019.02.031.

[9] M. Salmi, Additive manufacturing processes in medical applications, Materials (Basel).


14 (2021) 1–16. https://doi.org/10.3390/ma14010191.

[10] P. Szymczyk-Ziółkowska, M.B. Łabowska, J. Detyna, I. Michalak, P. Gruber, A review


of fabrication polymer scaffolds for biomedical applications using additive
manufacturing techniques, Biocybern. Biomed. Eng. 40 (2020) 624–638.
https://doi.org/10.1016/j.bbe.2020.01.015.

[11] M. Javaid, A. Haleem, Current status and applications of additive manufacturing in


dentistry: A literature-based review, J. Oral Biol. Craniofacial Res. 9 (2019) 179–185.
https://doi.org/10.1016/j.jobcr.2019.04.004.

110
[12] E. Ramirez-Cedillo, M.J. Uddin, J.A. Sandoval-Robles, R.A. Mirshams, L. Ruiz-Huerta,
C.A. Rodriguez, H.R. Siller, Process planning of L-PBF of AISI 316L for improving
surface quality and relating part integrity with microstructural characteristics, Surf.
Coatings Technol. 396 (2020) 125956. https://doi.org/10.1016/j.surfcoat.2020.125956.

[13] M.J. Uddin, E. Ramirez-Cedillo, R.A. Mirshams, H.R. Siller, Nanoindentation and
electron backscatter diffraction mapping in laser powder bed fusion of stainless steel
316L, Mater. Charact. 174 (2021) 111047.
https://doi.org/10.1016/j.matchar.2021.111047.

[14] M.J. Uddin, H.R. Siller, R.A. Mirshams, T.A. Byers, B. Rout, Effects of proton
irradiation on nanoindentation strain-rate sensitivity and microstructural properties in L-
PBF 17–4 PH stainless steels, Mater. Sci. Eng. A. 837 (2022) 142719.
https://doi.org/10.1016/j.msea.2022.142719.

[15] O. T. Wohlers, I. Campbell, O. Diegel, Wohlers Report 2018, Fort Collins: Wohlers
Associates. (2018), Addit. Manuf. 22 (2018) 844–851.
https://doi.org/10.1016/j.addma.2018.06.025.

[16] K.G. Prashanth, S. Scudino, H.J. Klauss, K.B. Surreddi, L. Löber, Z. Wang, A.K.
Chaubey, U. Kühn, J. Eckert, Microstructure and mechanical properties of Al–12Si
produced by selective laser melting: Effect of heat treatment, Mater. Sci. Eng. A. 590
(2014) 153–160. https://doi.org/https://doi.org/10.1016/j.msea.2013.10.023.

[17] C. Sun, Y. Wang, M.D. McMurtrey, N.D. Jerred, F. Liou, J. Li, Additive manufacturing
for energy: A review, Appl. Energy. 282 (2021).
https://doi.org/10.1016/j.apenergy.2020.116041.

[18] H. Attar, S. Ehtemam-Haghighi, N. Soro, D. Kent, M.S. Dargusch, Additive


manufacturing of low-cost porous titanium-based composites for biomedical applications:
Advantages, challenges and opinion for future development, J. Alloys Compd. 827
(2020) 154263. https://doi.org/10.1016/j.jallcom.2020.154263.

[19] J. den Boer, W. Lambrechts, H. Krikke, Additive manufacturing in military and


humanitarian missions: Advantages and challenges in the spare parts supply chain, J.
Clean. Prod. 257 (2020) 120301. https://doi.org/10.1016/j.jclepro.2020.120301.

[20] B.R. Betzler, B.J. Ade, A.J. Wysocki, P.K. Jain, P.C. Chesser, M.S. Greenwood, K.A.
Terrani, Transformational Challenge Reactor preconceptual core design studies, Nucl.
Eng. Des. 367 (2020) 110781. https://doi.org/10.1016/j.nucengdes.2020.110781.

[21] M. Hiser, A. Schneider, M. Audrain, A. Hull, Regulatory Research Perspective on


Additive Manufacturing for Nuclear Component Applications, J. Nucl. Mater. 546 (2021)
152726. https://doi.org/10.1016/j.jnucmat.2020.152726.

[22] M.W. Graham, J.C. King, T.R. Pavlov, C.A. Adkins, S.C. Middlemas, D.P. Guillen,
Impact of neutron irradiation on the thermophysical properties of additively

111
manufactured stainless steel and inconel, J. Nucl. Mater. 549 (2021) 152861.
https://doi.org/10.1016/j.jnucmat.2021.152861.

[23] W. Zhong, N. Sridharan, D. Isheim, K.G. Field, Y. Yang, K. Terrani, L. Tan,


Microstructures and mechanical properties of a modified 9Cr ferritic-martensitic steel in
the as-built condition after additive manufacturing, J. Nucl. Mater. 545 (2021) 152742.
https://doi.org/10.1016/j.jnucmat.2020.152742.

[24] M. McMurtrey, C. Sun, R.E. Rupp, C.H. Shiau, R. Hanbury, N. Jerred, R. O’Brien,
Investigation of the irradiation effects in additively manufactured 316L steel resulting in
decreased irradiation assisted stress corrosion cracking susceptibility, J. Nucl. Mater. 545
(2021). https://doi.org/10.1016/j.jnucmat.2020.152739.

[25] M. Saleh, Z. Zaidi, M. Ionescu, C. Hurt, K. Short, J. Daniels, P. Munroe, L. Edwards, D.


Bhattacharyya, Relationship between damage and hardness profiles in ion irradiated
SS316 using nanoindentation - Experiments and modelling, Int. J. Plast. 86 (2016) 151–
169. https://doi.org/10.1016/j.ijplas.2016.08.006.

[26] B.P. Eftink, J.S. Weaver, J.A. Valdez, V. Livescu, D. Chen, Y. Wang, C. Knapp, N.A.
Mara, S.A. Maloy, G.T. Gray, Proton irradiation and characterization of additively
manufactured 304L stainless steels, J. Nucl. Mater. 531 (2020) 152007.
https://doi.org/10.1016/j.jnucmat.2020.152007.

[27] S. Qin, Z. Wang, A.M. Beese, Orientation and stress state dependent plasticity and
damage initiation behavior of stainless steel 304L manufactured by laser powder bed
fusion additive manufacturing, Extrem. Mech. Lett. 45 (2021) 101271.
https://doi.org/10.1016/j.eml.2021.101271.

[28] W. Macek, R. Branco, J. Trembacz, J.D. Costa, J.A.M. Ferreira, C. Capela, Effect of
multiaxial bending-torsion loading on fracture surface parameters in high-strength steels
processed by conventional and additive manufacturing, Eng. Fail. Anal. 118 (2020)
104784. https://doi.org/10.1016/j.engfailanal.2020.104784.

[29] S. Spigarelli, C. Paoletti, M. Cabibbo, E. Cerri, E. Santecchia, On the creep performance


of the Ti‐6Al‐4V alloy processed by additive manufacturing, Addit. Manuf. 49 (2022)
102520. https://doi.org/10.1016/j.addma.2021.102520.

[30] T. Kurzynowski, E. Chlebus, B. Kuźnicka, J. Reiner, Parameters in Selective Laser


Melting for processing metallic powders, Proc. SPIE. 8239 (2012) 823914.
https://doi.org/10.1117/12.907292.

[31] J. Kim, S. Ji, Y.-S. Yun, J.-S. Yeo, A Review : Melt Pool Analysis for Selective Laser
Melting with Continuous Wave and Pulse Width Modulated Lasers, Appl. Sci. Converg.
Technol. 27 (2018) 113–119. https://doi.org/10.5757/ASCT.2018.27.6.113.

[32] J.H. Robinson, I.R.T. Ashton, E. Jones, P. Fox, C. Sutcliffe, The effect of hatch angle
rotation on parts manufactured using selective laser melting, Rapid Prototyp. J. 25 (2019)
289–298. https://doi.org/10.1108/RPJ-06-2017-0111.

112
[33] E. Liverani, A. Fortunato, A. Leardini, C. Belvedere, S. Siegler, L. Ceschini, A. Ascari,
Fabrication of Co–Cr–Mo endoprosthetic ankle devices by means of Selective Laser
Melting (SLM), Mater. Des. 106 (2016) 60–68.
https://doi.org/10.1016/j.matdes.2016.05.083.

[34] B. Vandenbroucke, J.-P. Kruth, Selective Laser Melting of Biocompatible Metals for
Rapid, Rapid Prototyp. J. 13 (2007) 148–159.

[35] N. Read, W. Wang, K. Essa, M.M. Attallah, Selective laser melting of AlSi10Mg alloy:
Process optimisation and mechanical properties development, Mater. Des. 65 (2015)
417–424. https://doi.org/10.1016/j.matdes.2014.09.044.

[36] X. Lou, D. Gandy, Advanced Manufacturing for Nuclear Energy, Jom. 71 (2019) 2834–
2836. https://doi.org/10.1007/s11837-019-03607-4.

[37] A. Hinojos, J. Mireles, A. Reichardt, P. Frigola, P. Hosemann, L.E. Murr, R.B. Wicker,
Joining of Inconel 718 and 316 Stainless Steel using electron beam melting additive
manufacturing technology, Mater. Des. 94 (2016) 17–27.
https://doi.org/10.1016/j.matdes.2016.01.041.

[38] W. Liu, J.N. DuPont, Fabrication of functionally graded TiC/Ti composites by laser
engineered net shaping, Scr. Mater. 48 (2003) 1337–1342. https://doi.org/10.1016/S1359-
6462(03)00020-4.

[39] B.E. Carroll, R.A. Otis, J.P. Borgonia, J.O. Suh, R.P. Dillon, A.A. Shapiro, D.C.
Hofmann, Z.K. Liu, A.M. Beese, Functionally graded material of 304L stainless steel and
inconel 625 fabricated by directed energy deposition: Characterization and
thermodynamic modeling, Acta Mater. 108 (2016) 46–54.
https://doi.org/10.1016/j.actamat.2016.02.019.

[40] L.D. Bobbio, B. Bocklund, R. Otis, J.P. Borgonia, R.P. Dillon, A.A. Shapiro, B.
McEnerney, Z.K. Liu, A.M. Beese, Characterization of a functionally graded material of
Ti-6Al-4V to 304L stainless steel with an intermediate V section, J. Alloys Compd. 742
(2018) 1031–1036. https://doi.org/10.1016/j.jallcom.2018.01.156.

[41] F.F. Noecker, J.N. DuPont, Functionally graded copper – steel using laser engineered net
shapingTM process, Int. Congr. Appl. Lasers Electro-Optics. 2002 (2002) 185430.
https://doi.org/10.2351/1.5066217.

[42] Y.M. Wang, T. Voisin, J.T. McKeown, J. Ye, N.P. Calta, Z. Li, Z. Zeng, Y. Zhang, W.
Chen, T.T. Roehling, R.T. Ott, M.K. Santala, P.J. Depond, M.J. Matthews, A. V. Hamza,
T. Zhu, Additively manufactured hierarchical stainless steels with high strength and
ductility, Nat. Mater. 17 (2018) 63–70. https://doi.org/10.1038/NMAT5021.

[43] A.J. Markworth, K.S. Ramesh, W.P. Parks, Modelling studies applied to functionally
graded materials, J. Mater. Sci. 30 (1995) 2183–2193.
https://doi.org/10.1007/BF01184560.

113
[44] K. Terrani, J. Kiggans, D. Chandler, C. Bryan, D. Pinkston, N. Sridharan, M. Gussev, M.
Norfolk, J. Burns, S. Suresh Babu, Additive Manufacturing of Research Reactor Control
Elements and Subsequent Neutron Irradiation, Trans. Am. Nucl. Soc. 114 (2016) 1229–
1230. http://inis.iaea.org/search/search.aspx?orig_q=RN:52032391.

[45] J.R. Burns, D. Chandler, B. Petrovic, K.A. Terrani, Depletion and lifetime performance
analysis of advanced manufactured control elements in the high flux isotope reactor
(HFIR), Proc. 2018 Int. Congr. Adv. Nucl. Power Plants, ICAPP 2018. (2018) 738–745.

[46] J. Rosales, I.J. van Rooyen, C.J. Parga, Characterizing surrogates to develop an additive
manufacturing process for U 3 Si 2 nuclear fuel, J. Nucl. Mater. 518 (2019) 117–128.
https://doi.org/10.1016/j.jnucmat.2019.02.026.

[47] A. V. Gelis, P. Kozak, A.T. Breshears, M.A. Brown, C. Launiere, E.L. Campbell, G.B.
Hall, T.G. Levitskaia, V.E. Holfeltz, G.J. Lumetta, Author Correction: Closing the
Nuclear Fuel Cycle with a Simplified Minor Actinide Lanthanide Separation Process
(ALSEP) and Additive Manufacturing (Scientific Reports, (2019), 9, 1, (12842),
10.1038/s41598-019-48619-x), Sci. Rep. 10 (2020) 1–11. https://doi.org/10.1038/s41598-
020-57574-x.

[48] G. Kalinin, V. Barabash, A. Cardella, J. Dietz, K. Ioki, R. Matera, R.T. Santoro, R.


Tivey, Assessment and selection of materials for ITER in-vessel components, J. Nucl.
Mater. 283–287 (2000) 10–19. https://doi.org/10.1016/S0022-3115(00)00305-6.

[49] B.S. Rodchenkov, Y.S. Strebkov, G.M. Kalinin, O.A. Golosov, Effect of ITER blanket
manufacturing process on the properties of the 316L(N)-IG steel, Fusion Eng. Des. 49–50
(2000) 657–660. https://doi.org/10.1016/S0920-3796(00)00359-8.

[50] R.K. Buddu, N. Chauhan, P.M. Raole, Mechanical properties and microstructural
investigations of TIG welded 40 mm and 60 mm thick SS 316L samples for fusion
reactor vacuum vessel applications, Fusion Eng. Des. 89 (2014) 3149–3158.
https://doi.org/10.1016/j.fusengdes.2014.10.006.

[51] C. Tan, G. Wang, L. Ji, Y. Tong, X.M. Duan, Investigation on 316L/W functionally
graded materials fabricated by mechanical alloying and spark plasma sintering, J. Nucl.
Mater. 469 (2016) 32–38. https://doi.org/10.1016/j.jnucmat.2015.11.024.

[52] D. Kong, X. Ni, C. Dong, X. Lei, L. Zhang, C. Man, J. Yao, X. Cheng, X. Li, Bio-
functional and anti-corrosive 3D printing 316L stainless steel fabricated by selective laser
melting, Mater. Des. 152 (2018) 88–101. https://doi.org/10.1016/j.matdes.2018.04.058.

[53] D. Dzhendov, T. Dikova, Application of selective laser melting in manufacturing of fixed


dental prostheses, J. IMAB - Annu. Proceeding (Scientific Pap. 22 (2016) 1414–1417.
https://doi.org/10.5272/jimab.2016224.1414.

[54] H. Zhou, Q. Fan, 3D reconstruction and SLM survey for dental implants, J. Mech. Med.
Biol. 17 (2017) 1–11. https://doi.org/10.1142/S0219519417500841.

114
[55] S.L. Sing, Concepts of Selective Laser Melting for Orthopaedic Implants BT - Selective
Laser Melting of Novel Titanium-Tantalum Alloy as Orthopaedic Biomaterial, in: S.L.
Sing (Ed.), Springer Singapore, Singapore, 2019: pp. 9–36. https://doi.org/10.1007/978-
981-13-2724-7_2.

[56] J. Čapek, M. Machová, M. Fousová, J. Kubásek, D. Vojtěch, J. Fojt, E. Jablonská, J.


Lipov, T. Ruml, Highly porous, low elastic modulus 316L stainless steel scaffold
prepared by selective laser melting, Mater. Sci. Eng. C. 69 (2016) 631–639.
https://doi.org/10.1016/j.msec.2016.07.027.

[57] K. Jha, J. Verma, C. Prakash, Fabrication and In Vitro Corrosion Characterization of


316L Stainless Steel for Medical Application BT - Biomaterials in Orthopaedics and
Bone Regeneration : Design and Synthesis, in: P.S. Bains, S.S. Sidhu, M. Bahraminasab,
C. Prakash (Eds.), Springer Singapore, Singapore, 2019: pp. 215–226.
https://doi.org/10.1007/978-981-13-9977-0_14.

[58] M.J.K. Lodhi, K.M. Deen, W. Haider, Corrosion behavior of additively manufactured
316L stainless steel in acidic media, Materialia. 2 (2018) 111–121.
https://doi.org/10.1016/j.mtla.2018.06.015.

[59] A.B. Kale, B.K. Kim, D.I. Kim, E.G. Castle, M. Reece, S.H. Choi, An investigation of
the corrosion behavior of 316L stainless steel fabricated by SLM and SPS techniques,
Mater. Charact. 163 (2020) 110204. https://doi.org/10.1016/j.matchar.2020.110204.

[60] K. Saeidi, X. Gao, Y. Zhong, Z.J. Shen, Hardened austenite steel with columnar sub-
grain structure formed by laser melting, Mater. Sci. Eng. A. 625 (2015) 221–229.
https://doi.org/10.1016/j.msea.2014.12.018.

[61] H.W. Huang, Z.B. Wang, J. Lu, K. Lu, Fatigue behaviors of AISI 316L stainless steel
with a gradient nanostructured surface layer, Acta Mater. 87 (2015) 150–160.
https://doi.org/10.1016/j.actamat.2014.12.057.

[62] D. Kong, X. Ni, C. Dong, L. Zhang, C. Man, X. Cheng, X. Li, Anisotropy in the
microstructure and mechanical property for the bulk and porous 316L stainless steel
fabricated via selective laser melting, Mater. Lett. 235 (2019) 1–5.
https://doi.org/10.1016/j.matlet.2018.09.152.

[63] X. Ni, D. Kong, Y. Wen, L. Zhang, W. Wu, B. He, L. Lu, D. Zhu, Anisotropy in
mechanical properties and corrosion resistance of 316L stainless steel fabricated by
selective laser melting, Int. J. Miner. Metall. Mater. 26 (2019) 319–328.
https://doi.org/10.1007/s12613-019-1740-x.

[64] J. Andrew, LA-UR-19-32212 Title : Author ( s ): Intended for : Nuclear Reactor


Materials and Anisotropy Issued :, (2019).

[65] Y.D. Im, K.H. Kim, K.H. Jung, Y.K. Lee, K.H. Song, Anisotropic Mechanical Behavior
of Additive Manufactured AISI 316L Steel, Metall. Mater. Trans. A Phys. Metall. Mater.
Sci. 50 (2019) 2014–2021. https://doi.org/10.1007/s11661-019-05139-7.

115
[66] S. Huang, S.L. Sing, G. de Looze, R. Wilson, W.Y. Yeong, Laser powder bed fusion of
titanium-tantalum alloys: Compositions and designs for biomedical applications, J. Mech.
Behav. Biomed. Mater. 108 (2020) 103775.
https://doi.org/10.1016/j.jmbbm.2020.103775.

[67] T. Pinomaa, M. Lindroos, M. Walbrühl, N. Provatas, A. Laukkanen, The significance of


spatial length scales and solute segregation in strengthening rapid solidification
microstructures of 316L stainless steel, Acta Mater. 184 (2020) 1–16.
https://doi.org/10.1016/j.actamat.2019.10.044.

[68] Y. Kok, X.P. Tan, P. Wang, M.L.S. Nai, N.H. Loh, E. Liu, S.B. Tor, Anisotropy and
heterogeneity of microstructure and mechanical properties in metal additive
manufacturing: A critical review, Mater. Des. 139 (2018) 565–586.
https://doi.org/10.1016/j.matdes.2017.11.021.

[69] Z. Sun, X. Tan, S.B. Tor, C.K. Chua, Simultaneously enhanced strength and ductility for
3D-printed stainless steel 316L by selective laser melting, NPG Asia Mater. 10 (2018)
127–136. https://doi.org/10.1038/s41427-018-0018-5.

[70] M.L. Montero-Sistiaga, M. Godino-Martinez, K. Boschmans, J.P. Kruth, J. Van


Humbeeck, K. Vanmeensel, Microstructure evolution of 316L produced by HP-SLM
(high power selective laser melting), Addit. Manuf. 23 (2018) 402–410.
https://doi.org/10.1016/j.addma.2018.08.028.

[71] H.R. Lashgari, C. Kong, E. Adabifiroozjaei, S. Li, Microstructure, post thermal treatment
response, and tribological properties of 3D printed 17-4 PH stainless steel, Wear. 456–
457 (2020). https://doi.org/10.1016/j.wear.2020.203367.

[72] F. Ahmed, U. Ali, D. Sarker, E. Marzbanrad, K. Choi, Y. Mahmoodkhani, E. Toyserkani,


Study of powder recycling and its effect on printed parts during laser powder-bed fusion
of 17-4 PH stainless steel, J. Mater. Process. Technol. 278 (2020) 116522.
https://doi.org/10.1016/j.jmatprotec.2019.116522.

[73] A. Soltani-Tehrani, J. Pegues, N. Shamsaei, Fatigue behavior of additively manufactured


17-4 PH stainless steel: The effects of part location and powder re-use, Addit. Manuf. 36
(2020). https://doi.org/10.1016/j.addma.2020.101398.

[74] M. Alnajjar, F. Christien, V. Barnier, C. Bosch, K. Wolski, A.D. Fortes, M. Telling,


Influence of microstructure and manganese sulfides on corrosion resistance of selective
laser melted 17-4 PH stainless steel in acidic chloride medium, Corros. Sci. 168 (2020)
108585. https://doi.org/10.1016/j.corsci.2020.108585.

[75] S.J. Zinkle, K.A. Terrani, L.L. Snead, Motivation for utilizing new high-performance
advanced materials in nuclear energy systems, Curr. Opin. Solid State Mater. Sci. 20
(2016) 401–410. https://doi.org/10.1016/j.cossms.2016.10.004.

116
[76] N. Bibhanshu, M.N. Gussev, T.M. Rosseel, Complexity of deformation mechanism in
neutron-irradiated 304L austenitic stainless steel at microstructural scale, Mater. Charact.
178 (2021) 111218. https://doi.org/10.1016/j.matchar.2021.111218.

[77] Z. Wang, W.M. Mook, C. Niederberger, R. Ghisleni, L. Philippe, J. Michler,


Compression of nanowires using a flat indenter: Diametrical elasticity measurement,
Nano Lett. 12 (2012) 2289–2293. https://doi.org/10.1021/nl300103z.

[78] J.Y. Kim, J.R. Greer, Tensile and compressive behavior of gold and molybdenum single
crystals at the nano-scale, Acta Mater. 57 (2009) 5245–5253.
https://doi.org/10.1016/j.actamat.2009.07.027.

[79] C.S. Kaira, T.J. Stannard, V. De Andrade, F. De Carlo, N. Chawla, Exploring novel
deformation mechanisms in aluminum–copper alloys using in situ 4D nanomechanical
testing, Acta Mater. 176 (2019) 242–249. https://doi.org/10.1016/j.actamat.2019.07.016.

[80] M. Zamanzade, J.R. Velayarce, O.T. Abad, C. Motz, A. Barnoush, Mechanical behavior
of iron aluminides: A comparison of nanoindentation, compression and bending of
micropillars, Mater. Sci. Eng. A. 652 (2016) 370–376.
https://doi.org/10.1016/j.msea.2015.11.088.

[81] E. Camposilvan, M. Anglada, Size and plasticity effects in zirconia micropillars


compression, Acta Mater. 103 (2016) 882–892.
https://doi.org/10.1016/j.actamat.2015.10.047.

[82] K. Chu, K. Yan, F. Ren, Q. Sun, A dual-pillar method for measurement of stress-strain
response of material at microscale, Scr. Mater. 172 (2019) 138–143.
https://doi.org/10.1016/j.scriptamat.2019.07.021.

[83] J.S. Weaver, M. Kreitman, J.C. Heigel, M.A. Donmez, Mechanical Property
Characterization of Single Scan Laser Tracks of Nickel Superalloy 625 by
Nanoindentation BT - TMS 2019 148th Annual Meeting & Exhibition Supplemental
Proceedings, in: Springer International Publishing, Cham, 2019: pp. 269–278.

[84] P. Hosemann, Small-scale mechanical testing on nuclear materials: Bridging the


experimental length-scale gap, Scr. Mater. 143 (2017).
https://doi.org/10.1016/j.scriptamat.2017.04.026.

[85] J.S. Weaver, V. Livescu, N.A. Mara, A comparison of adiabatic shear bands in wrought
and additively manufactured 316L stainless steel using nanoindentation and electron
backscatter diffraction, J. Mater. Sci. 55 (2020) 1738–1752.
https://doi.org/10.1007/s10853-019-03994-8.

[86] X. Wang, X. Gong, K. Chou, Scanning Speed Effect on Mechanical Properties of Ti-6Al-
4V Alloy Processed by Electron Beam Additive Manufacturing, Procedia Manuf. 1
(2015) 287–295. https://doi.org/10.1016/j.promfg.2015.09.026.

117
[87] X. Sun, F. Chen, H. Huang, J. Lin, X. Tang, Effects of interfaces on the helium bubble
formation and radiation hardening of an austenitic stainless steel achieved by additive
manufacturing, Appl. Surf. Sci. 467–468 (2019) 1134–1139.
https://doi.org/10.1016/j.apsusc.2018.10.268.

[88] A. Reichardt, R.P. Dillon, J.P. Borgonia, A.A. Shapiro, B.W. McEnerney, T. Momose, P.
Hosemann, Development and characterization of Ti-6Al-4V to 304L stainless steel
gradient components fabricated with laser deposition additive manufacturing, Mater. Des.
104 (2016) 404–413. https://doi.org/10.1016/j.matdes.2016.05.016.

[89] N.T. Aboulkhair, I. Maskery, C. Tuck, I. Ashcroft, N.M. Everitt, The microstructure and
mechanical properties of selectively laser melted AlSi10Mg: The effect of a conventional
T6-like heat treatment, Mater. Sci. Eng. A. 667 (2016) 139–146.
https://doi.org/10.1016/j.msea.2016.04.092.

[90] H. Wang, A. Dhiman, H.E. Ostergaard, Y. Zhang, T. Siegmund, J.J. Kruzic, V. Tomar,
Nanoindentation based properties of Inconel 718 at elevated temperatures: A comparison
of conventional versus additively manufactured samples, Int. J. Plast. 120 (2019) 380–
394. https://doi.org/10.1016/j.ijplas.2019.04.018.

[91] B. AlMangour, J.M. Yang, Understanding the deformation behavior of 17-4 precipitate
hardenable stainless steel produced by direct metal laser sintering using micropillar
compression and TEM, Int. J. Adv. Manuf. Technol. 90 (2017) 119–126.
https://doi.org/10.1007/s00170-016-9367-9.

[92] A. Behroozfar, S. Daryadel, S.R. Morsali, S. Moreno, M. Baniasadi, R.A. Bernal, M.


Minary-Jolandan, Microscale 3D Printing of Nanotwinned Copper, Adv. Mater. 30
(2018) 1–6. https://doi.org/10.1002/adma.201705107.

[93] D. Gu, Y.C. Hagedorn, W. Meiners, G. Meng, R.J.S. Batista, K. Wissenbach, R.


Poprawe, Densification behavior, microstructure evolution, and wear performance of
selective laser melting processed commercially pure titanium, Acta Mater. 60 (2012)
3849–3860. https://doi.org/10.1016/j.actamat.2012.04.006.

[94] J.M. C. zasopismo W Żórawski, J Sienicki, S Kowalski, Additive manufacturing of


titanium with application of cold spray process, Tech. Trans. 10 (2018) 201–208.
https://doi.org/10.4467/2353737xct.18.158.9308.

[95] N.M. Everitt, N.T. Aboulkhair, I. Maskery, C.J. Tuck, I. Ashcroft, Nanoindentation
shows uniform local mechanical properties across melt pools and layers produced by
selective laser melting of AlSi 10Mg alloy, Adv. Mater. Lett. 7 (2016) 13–16.
https://doi.org/10.5185/amlett.2016.6171.

[96] M. Munther, T. Palma, F. Tavangarian, A. Beheshti, K. Davami, Nanomechanical


properties of additively and traditionally manufactured nickel-chromium-based
superalloys through instrumented nanoindentation, Manuf. Lett. 23 (2020) 39–43.
https://doi.org/10.1016/j.mfglet.2019.09.003.

118
[97] M.J. Uddin, E. Ramirez-Cedillo, R.A. Mirshams, H.R. Siller, Nanoindentation and
electron backscatter diffraction mapping in laser powder bed fusion of stainless steel
316L, Mater. Charact. 174 (2021) 111047.
https://doi.org/10.1016/j.matchar.2021.111047.

[98] P. Hosemann, C. Shin, D. Kiener, Small scale mechanical testing of irradiated materials,
J. Mater. Res. 30 (2015) 1231–1245. https://doi.org/10.1557/jmr.2015.26.

[99] B. AlMangour, J.M. Yang, Improving the surface quality and mechanical properties by
shot-peening of 17-4 stainless steel fabricated by additive manufacturing, Mater. Des.
110 (2016) 914–924. https://doi.org/10.1016/j.matdes.2016.08.037.

[100] C.-H. Shiau, C. Sun, M. McMurtrey, R. O’Brien, F.A. Garner, L. Shao, Orientation-
selected micro-pillar compression of additively manufactured 316L stainless steels:
comparison of as-manufactured, annealed, and proton-irradiated variants, J. Nucl. Mater.
(2022) 153739. https://doi.org/10.1016/j.jnucmat.2022.153739.

[101] X. Wang, B. Zheng, K. Yu, S. Jiang, E.J. Lavernia, J.M. Schoenung, The role of cell
boundary orientation on mechanical behavior: A site-specific micro-pillar
characterization study, Addit. Manuf. 46 (2021) 102154.
https://doi.org/10.1016/j.addma.2021.102154.

[102] J. Menčík, Uncertainties and Errors in Nanoindentation, in: 2012.

[103] J.F. Ziegler, M.D. Ziegler, J.P. Biersack, SRIM - The stopping and range of ions in
matter (2010), Nucl. Instruments Methods Phys. Res. Sect. B Beam Interact. with Mater.
Atoms. 268 (2010) 1818–1823. https://doi.org/10.1016/j.nimb.2010.02.091.

[104] J.M. Young, S. Singh, T.A. Byers, D.C. Jones, B. Rout, Synthesis of crystalline phases in
space silicate analogues with helium ion irradiation, Nucl. Instruments Methods Phys.
Res. Sect. B Beam Interact. with Mater. Atoms. 443 (2019) 79–83.
https://doi.org/10.1016/j.nimb.2019.01.052.

[105] T.H. Pham, J.J. Kim, S.E. Kim, Estimating constitutive equation of structural steel using
indentation, Int. J. Mech. Sci. 90 (2015) 151–161.
https://doi.org/10.1016/j.ijmecsci.2014.11.007.

[106] D. Caillard, J.-L. Martin, Introduction BT - Thermally Activated Mechanisms in Crystal


Plasticity, Therm. Act. Mech. Cryst. Plast. 8 (2003) 3–9.

[107] V. Maier-Kiener, K. Durst, Advanced Nanoindentation Testing for Studying Strain-Rate


Sensitivity and Activation Volume, Jom. 69 (2017) 2246–2255.
https://doi.org/10.1007/s11837-017-2536-y.

[108] S. Vadalakonda, R. Banerjee, A. Puthcode, R. Mirshams, Comparison of incipient


plasticity in bcc and fcc metals studied using nanoindentation, Mater. Sci. Eng. A. 426
(2006) 208–213. https://doi.org/10.1016/j.msea.2006.04.001.

119
[109] A. Gouldstone, H.J. Koh, K.Y. Zeng, A.E. Giannakopoulos, S. Suresh, Discrete and
continuous deformation during nanoindentation of thin films, Acta Mater. 48 (2000)
2277–2295. https://doi.org/10.1016/S1359-6454(00)00009-4.

[110] S.Y. Park, Y.C. Kim, R.S. Ruoff, J.Y. Kim, Incipient plasticity and fully plastic contact
behavior of copper coated with a graphene layer, APL Mater. 7 (2019) 0–5.
https://doi.org/10.1063/1.5086333.

[111] L. Zhao, M. Alam, J. Zhang, R. Janisch, A. Hartmaier, Amorphization-governed elasto-


plastic deformation under nanoindentation in cubic (3C) silicon carbide, Ceram. Int. 46
(2020) 12470–12479. https://doi.org/10.1016/j.ceramint.2020.02.009.

[112] A. Yadollahi, N. Shamsaei, S.M. Thompson, A. Elwany, L. Bian, Effects of building


orientation and heat treatment on fatigue behavior of selective laser melted 17-4 PH
stainless steel, Int. J. Fatigue. 94 (2017) 218–235.
https://doi.org/10.1016/j.ijfatigue.2016.03.014.

[113] Fischer-Cripps AC , Author, Nicholson DW , Reviewer, Nanoindentation. Mechanical


Engineering Series, Appl. Mech. Rev. 57 (2004) B12–B12.
https://doi.org/10.1115/1.1704625.

[114] G.S. Was, Fundamentals of Radiation Materials Science, 2020.


https://doi.org/10.1016/B978-0-12-803581-8.00668-8.

[115] W.D. Nix, H. Gao, Indentation size effects in crystalline materials: A law for strain
gradient plasticity, J. Mech. Phys. Solids. 46 (1998) 411–425.
https://doi.org/10.1016/S0022-5096(97)00086-0.

[116] M.D. Uchic, D.M. Dimiduk, J.N. Florando, W.D. Nix, Sample dimensions influence
strength and crystal plasticity, Science (80-. ). 305 (2004) 986–989.
https://doi.org/10.1126/science.1098993.

[117] K.E. Knipling, D.J. Rowenhorst, R.W. Fonda, G. Spanos, Effects of focused ion beam
milling on austenite stability in ferrous alloys, Mater. Charact. 61 (2010) 1–6.
https://doi.org/DOI:101016/jmatchar200909013.

[118] J.R. Trelewicz, G.P. Halada, O.K. Donaldson, G. Manogharan, Microstructure and
Corrosion Resistance of Laser Additively Manufactured 316L Stainless Steel, Jom. 68
(2016) 850–859. https://doi.org/10.1007/s11837-016-1822-4.

[119] P.K. Samal, I.I.I. Warzel Roland, S.O. Shah, Powder Metallurgy Stainless Steels
Applications, in: Powder Metall., ASM International, 2015.
https://doi.org/10.31399/asm.hb.v07.a0006098.

[120] A. Leicht, C.H. Yu, V. Luzin, U. Klement, E. Hryha, Effect of scan rotation on the
microstructure development and mechanical properties of 316L parts produced by laser
powder bed fusion, Mater. Charact. 163 (2020) 2–10.
https://doi.org/10.1016/j.matchar.2020.110309.

120
[121] A. Leicht, M. Rashidi, U. Klement, E. Hryha, Effect of process parameters on the
microstructure, tensile strength and productivity of 316L parts produced by laser powder
bed fusion, Mater. Charact. 159 (2020) 110016.
https://doi.org/10.1016/j.matchar.2019.110016.

[122] Y. Song, Q. Sun, K. Guo, X. Wang, J. Liu, J. Sun, Effect of scanning strategies on the
microstructure and mechanical behavior of 316L stainless steel fabricated by selective
laser melting, Mater. Sci. Eng. A. 793 (2020) 139879.
https://doi.org/10.1016/j.msea.2020.139879.

[123] J.J. Marattukalam, D. Karlsson, V. Pacheco, P. Beran, U. Wiklund, U. Jansson, B.


Hjörvarsson, M. Sahlberg, The effect of laser scanning strategies on texture, mechanical
properties, and site-specific grain orientation in selective laser melted 316L SS, Mater.
Des. 193 (2020). https://doi.org/10.1016/j.matdes.2020.108852.

[124] U. Scipioni Bertoli, B.E. MacDonald, J.M. Schoenung, Stability of cellular


microstructure in laser powder bed fusion of 316L stainless steel, Mater. Sci. Eng. A. 739
(2019) 109–117. https://doi.org/10.1016/j.msea.2018.10.051.

[125] P. Krakhmalev, G. Fredriksson, K. Svensson, I. Yadroitsev, I. Yadroitsava, M.


Thuvander, R. Peng, Microstructure, solidification texture, and thermal stability of 316 L
stainless steel manufactured by laser powder bed fusion, Metals (Basel). 8 (2018).
https://doi.org/10.3390/met8080643.

[126] J. Suryawanshi, K.G. Prashanth, U. Ramamurty, Mechanical behavior of selective laser


melted 316L stainless steel, Mater. Sci. Eng. A. 696 (2017) 113–121.
https://doi.org/10.1016/j.msea.2017.04.058.

[127] D. Wang, C. Song, Y. Yang, Y. Bai, Investigation of crystal growth mechanism during
selective laser melting and mechanical property characterization of 316L stainless steel
parts, Mater. Des. 100 (2016) 291–299. https://doi.org/10.1016/j.matdes.2016.03.111.

[128] Y. Zhong, L. Liu, S. Wikman, D. Cui, Z. Shen, Intragranular cellular segregation network
structure strengthening 316L stainless steel prepared by selective laser melting, J. Nucl.
Mater. 470 (2016) 170–178. https://doi.org/10.1016/j.jnucmat.2015.12.034.

[129] A. Röttger, J. Boes, W. Theisen, M. Thiele, C. Esen, A. Edelmann, R. Hellmann,


Microstructure and mechanical properties of 316L austenitic stainless steel processed by
different SLM devices, Int. J. Adv. Manuf. Technol. 108 (2020) 769–783.
https://doi.org/10.1007/s00170-020-05371-1.

[130] A. Aggarwal, S. Patel, A. Kumar, Selective Laser Melting of 316L Stainless Steel:
Physics of Melting Mode Transition and Its Influence on Microstructural and Mechanical
Behavior, Jom. 71 (2019) 1105–1116. https://doi.org/10.1007/s11837-018-3271-8.

[131] P. Lu, Z. Cheng-Lin, W. Liang, L. Tong, L. Xiao-Cheng, Research on mechanical


properties and microstructure by selective laser melting of 316L stainless steel, Mater.
Res. Express. 6 (2019). https://doi.org/10.1088/2053-1591/ab6b67.

121
[132] X. qing Ni, D. cheng Kong, Y. Wen, L. Zhang, W. heng Wu, B. bei He, L. Lu, D. xiang
Zhu, Anisotropy in mechanical properties and corrosion resistance of 316L stainless steel
fabricated by selective laser melting, Int. J. Miner. Metall. Mater. 26 (2019) 319–328.
https://doi.org/10.1007/s12613-019-1740-x.

[133] A. Leicht, U. Klement, E. Hryha, Effect of build geometry on the microstructural


development of 316L parts produced by additive manufacturing, Mater. Charact. 143
(2018) 137–143. https://doi.org/10.1016/j.matchar.2018.04.040.

[134] M.S. Pham, B. Dovgyy, P.A. Hooper, Twinning induced plasticity in austenitic stainless
steel 316L made by additive manufacturing, Mater. Sci. Eng. A. 704 (2017) 102–111.
https://doi.org/10.1016/j.msea.2017.07.082.

[135] W.M. Tucho, V.H. Lysne, H. Austbø, A. Sjolyst-Kverneland, V. Hansen, Investigation of


effects of process parameters on microstructure and hardness of SLM manufactured
SS316L, J. Alloys Compd. 740 (2018) 910–925.
https://doi.org/10.1016/j.jallcom.2018.01.098.

[136] O. Andreau, I. Koutiri, P. Peyre, J.D. Penot, N. Saintier, E. Pessard, T. De Terris, C.


Dupuy, T. Baudin, Texture control of 316L parts by modulation of the melt pool
morphology in selective laser melting, J. Mater. Process. Technol. 264 (2019) 21–31.
https://doi.org/10.1016/j.jmatprotec.2018.08.049.

[137] M. Alnajjar, F. Christien, K. Wolski, C. Bosch, Evidence of austenite by-passing in a


stainless steel obtained from laser melting additive manufacturing, Addit. Manuf. 25
(2019) 187–195. https://doi.org/10.1016/j.addma.2018.11.004.

[138] A. Sathyanath, A. Meena, Microstructural evolution and strain hardening behavior of


heat-treated 17-4 PH stainless steel, Mater. Today Commun. 25 (2020) 101416.
https://doi.org/10.1016/j.mtcomm.2020.101416.

[139] S.D. Meredith, J.S. Zuback, J.S. Keist, T.A. Palmer, Impact of composition on the heat
treatment response of additively manufactured 17–4 PH grade stainless steel, Mater. Sci.
Eng. A. 738 (2018) 44–56. https://doi.org/10.1016/j.msea.2018.09.066.

[140] S. Vunnam, A. Saboo, C. Sudbrack, T.L. Starr, Effect of powder chemical composition
on the as-built microstructure of 17-4 PH stainless steel processed by selective laser
melting, Addit. Manuf. 30 (2019) 100876. https://doi.org/10.1016/j.addma.2019.100876.

[141] U. Ali, R. Esmaeilizadeh, F. Ahmed, D. Sarker, W. Muhammad, A. Keshavarzkermani,


Y. Mahmoodkhani, E. Marzbanrad, E. Toyserkani, Identification and characterization of
spatter particles and their effect on surface roughness, density and mechanical response
of 17-4 PH stainless steel laser powder-bed fusion parts, Mater. Sci. Eng. A. 756 (2019)
98–107. https://doi.org/10.1016/j.msea.2019.04.026.

[142] T.H. Hsu, Y.J. Chang, C.Y. Huang, H.W. Yen, C.P. Chen, K.K. Jen, A.C. Yeh,
Microstructure and property of a selective laser melting process induced oxide dispersion

122
strengthened 17-4 PH stainless steel, J. Alloys Compd. 803 (2019) 30–41.
https://doi.org/10.1016/j.jallcom.2019.06.289.

[143] Y. Sun, R.J. Hebert, M. Aindow, Effect of heat treatments on microstructural evolution of
additively manufactured and wrought 17-4PH stainless steel, Mater. Des. 156 (2018)
429–440. https://doi.org/10.1016/j.matdes.2018.07.015.

[144] X. Xiao, L. Yu, Cross-sectional nano-indentation of ion-irradiated steels: Finite element


simulations based on the strain-gradient crystal plasticity theory, Int. J. Eng. Sci. 143
(2019) 56–72. https://doi.org/10.1016/j.ijengsci.2019.06.015.

[145] F. Röder, C. Heintze, S. Pecko, S. Akhmadaliev, F. Bergner, A. Ulbricht, E. Altstadt,


Nanoindentation of ion-irradiated reactor pressure vessel steels–model-based
interpretation and comparison with neutron irradiation, Philos. Mag. 98 (2018) 911–933.
https://doi.org/10.1080/14786435.2018.1425007.

[146] X.F. Liu, C.K. Tse, Impact of degree mixing pattern on consensus formation in social
networks, Phys. A Stat. Mech. Its Appl. 407 (2014) 1–6.
https://doi.org/10.1016/j.physa.2014.03.086.

[147] E. Ramirez-Cedillo, M.J. Uddin, J.A. Sandoval-Robles, R.A. Mirshams, L. Ruiz-Huerta,


C.A. Rodriguez, H.R. Siller, Process planning of L-PBF of AISI 316L for improving
surface quality and relating part integrity with microstructural characteristics, Surf.
Coatings Technol. 396 (2020) 125956. https://doi.org/10.1016/j.surfcoat.2020.125956.

[148] H. Irrinki, M. Dexter, B. Barmore, R. Enneti, S. Pasebani, S. Badwe, J. Stitzel, R.


Malhotra, S. V. Atre, Effects of Powder Attributes and Laser Powder Bed Fusion (L-
PBF) Process Conditions on the Densification and Mechanical Properties of 17-4 PH
Stainless Steel, Jom. 68 (2016) 860–868. https://doi.org/10.1007/s11837-015-1770-4.

[149] H. Irrinki, S.D. Nath, M. Alhofors, J. Stitzel, O. Gulsoy, S. V. Atre, Microstructures,


properties, and applications of laser sintered 17-4PH stainless steel, J. Am. Ceram. Soc.
102 (2019) 5679–5690. https://doi.org/10.1111/jace.16372.

[150] J. Hay, Strain-Rate Sensitivity of Thin Metal Films by Instrumented Indentation


Application Note, (n.d.) 1–8.

[151] M. Alnajjar, F. Christien, K. Wolski, C. Bosch, Evidence of austenite by-passing in a


stainless steel obtained from laser melting additive manufacturing, Addit. Manuf. 25
(2019) 187–195. https://doi.org/10.1016/j.addma.2018.11.004.

[152] X. Lin, Y. Cao, X. Wu, H. Yang, J. Chen, W. Huang, Microstructure and mechanical
properties of laser forming repaired 17-4PH stainless steel, Mater. Sci. Eng. A. 553
(2012) 80–88. https://doi.org/10.1016/j.msea.2012.05.095.

[153] T.H. Hsu, Y.J. Chang, C.Y. Huang, H.W. Yen, C.P. Chen, K.K. Jen, A.C. Yeh,
Microstructure and property of a selective laser melting process induced oxide dispersion

123
strengthened 17-4 PH stainless steel, J. Alloys Compd. 803 (2019) 30–41.
https://doi.org/10.1016/j.jallcom.2019.06.289.

[154] S. Takaki, K. Fukunaga, J. Syarif, T. Tsuchiyama, Effect of grain refinement on thermal


stability of metastable austenitic steel, Mater. Trans. 45 (2004) 2245–2251.
https://doi.org/10.2320/matertrans.45.2245.

[155] H.K. Rafi, D. Pal, N. Patil, T.L. Starr, B.E. Stucker, Microstructure and Mechanical
Behavior of 17-4 Precipitation Hardenable Steel Processed by Selective Laser Melting, J.
Mater. Eng. Perform. 23 (2014) 4421–4428. https://doi.org/10.1007/s11665-014-1226-y.

[156] J.S. Weaver, J. Whiting, V. Tondare, C. Beauchamp, M. Peltz, J. Tarr, T.Q. Phan, M.A.
Donmez, The effects of particle size distribution on the rheological properties of the
powder and the mechanical properties of additively manufactured 17-4 PH stainless steel,
Addit. Manuf. 39 (2021) 101851. https://doi.org/10.1016/j.addma.2021.101851.

[157] S. Vunnam, S. Dobson, A. Saboo, D. Frankel, C. Sudbrack, T.L. Starr, Effect of Powder
Chemical Composition on Microstructures and, (2019) 463–477.

[158] L. Zai, C. Zhang, Y. Wang, W. Guo, D. Wellmann, X. Tong, Y. Tian, Laser powder bed
fusion of precipitation-hardened martensitic stainless steels: A review, Metals (Basel). 10
(2020). https://doi.org/10.3390/met10020255.

[159] S. Suwas, R.K. Ray, Crystallographic Texture of Materials, Springer London, 2016.
https://books.google.com/books?id=nstbvgAACAAJ.

[160] A. Kudzal, B. McWilliams, C. Hofmeister, F. Kellogg, J. Yu, J. Taggart-Scarff, J. Liang,


Effect of scan pattern on the microstructure and mechanical properties of Powder Bed
Fusion additive manufactured 17-4 stainless steel, Mater. Des. 133 (2017) 205–215.
https://doi.org/10.1016/j.matdes.2017.07.047.

[161] P. Leo, M. Cabibbo, A. Del Prete, S. Giganto, S. Martínez-Pellitero, J. Barreiro, Laser


defocusing effect on the microstructure and defects of 17-4ph parts additively
manufactured by slm at a low energy input, Metals (Basel). 11 (2021).
https://doi.org/10.3390/met11040588.

[162] H.R. Lashgari, Y. Xue, C. Onggowarsito, C. Kong, S. Li, Microstructure, Tribological


Properties and Corrosion Behaviour of Additively Manufactured 17-4PH Stainless Steel:
Effects of Scanning Pattern, Build Orientation, and Single vs. Double scan, Mater. Today
Commun. 25 (2020). https://doi.org/10.1016/j.mtcomm.2020.101535.

[163] H. Irrinki, J.S.D. Jangam, S. Pasebani, S. Badwe, J. Stitzel, K. Kate, O. Gulsoy, S. V.


Atre, Effects of particle characteristics on the microstructure and mechanical properties
of 17-4 PH stainless steel fabricated by laser-powder bed fusion, Powder Technol. 331
(2018) 192–203. https://doi.org/10.1016/j.powtec.2018.03.025.

124
[164] J. Lin, F. Chen, X. Tang, J. Liu, S. Shen, G. Ge, Radiation-induced swelling and
hardening of 316L stainless steel fabricated by selected laser melting, Vacuum. 174
(2020) 109183. https://doi.org/10.1016/j.vacuum.2020.109183.

[165] T.H. Hsu, P.C. Huang, M.Y. Lee, K.C. Chang, C.C. Lee, M.Y. Li, C.P. Chen, K.K. Jen,
A.C. Yeh, Effect of processing parameters on the fractions of martensite in 17-4 PH
stainless steel fabricated by selective laser melting, J. Alloys Compd. 859 (2021) 157758.
https://doi.org/10.1016/j.jallcom.2020.157758.

[166] B.S. Hengfeng Gu*, Haijun Gong*, Deepankar Pal*, Khalid Rafi*, Thomas Starr†,
Influences of Energy Density on Porosity and Microstructure of Selective Laser Melted
17- 4PH Stainless Steel, Mater. Sci. (2013) 117-99 ‫ ;ص‬8 ‫ﺷﻤﺎره‬.

[167] R. Rashid, S.H. Masood, D. Ruan, S. Palanisamy, R.A. Rahman Rashid, M. Brandt,
Effect of scan strategy on density and metallurgical properties of 17-4PH parts printed by
Selective Laser Melting (SLM), J. Mater. Process. Technol. 249 (2017) 502–511.
https://doi.org/10.1016/j.jmatprotec.2017.06.023.

[168] S. Romano, P.D. Nezhadfar, N. Shamsaei, M. Seifi, S. Beretta, High cycle fatigue
behavior and life prediction for additively manufactured 17-4 PH stainless steel: Effect of
sub-surface porosity and surface roughness, Theor. Appl. Fract. Mech. 106 (2020).
https://doi.org/10.1016/j.tafmec.2020.102477.

[169] D. Basu, Z. Wu, J.L.L. Meyer, E. Larson, R. Kuo, A. Rollett, Entrapped Gas and Process
Parameter-Induced Porosity Formation in Additively Manufactured 17-4 PH Stainless
Steel, J. Mater. Eng. Perform. (2021). https://doi.org/10.1007/s11665-021-05695-3.

[170] Z. Wu, D. Basu, J.L.L. Meyer, E. Larson, R. Kuo, J. Beuth, A. Rollett, Study of Powder
Gas Entrapment and Its Effects on Porosity in 17-4 PH Stainless Steel Parts Fabricated in
Laser Powder Bed Fusion, Jom. 73 (2021) 177–188. https://doi.org/10.1007/s11837-020-
04491-z.

[171] H. Gu, H. Gong, D. Pal, K. Rafi, T. Starr, B. Stucker, Influences of Energy Density on
Porosity and Microstructure of Selective Laser Melted 17-4PH Stainless Steel, in: 2013.

[172] Z. Li, T. Voisin, J.T. McKeown, J. Ye, T. Braun, C. Kamath, W.E. King, Y.M. Wang,
Tensile properties, strain rate sensitivity, and activation volume of additively
manufactured 316L stainless steels, Int. J. Plast. 120 (2019) 395–410.
https://doi.org/10.1016/j.ijplas.2019.05.009.

[173] B. Merle, W.H. Higgins, G.M. Pharr, Critical issues in conducting constant strain rate
nanoindentation tests at higher strain rates, J. Mater. Res. (2019).
https://doi.org/10.1557/jmr.2019.292.

[174] M. Zhang, J. Li, B. Tang, H. Kou, J. Fan, Mechanical characterization and strain-rate
sensitivity measurement of Ti-7333 alloy based on nanoindentation and crystal plasticity
modeling, Prog. Nat. Sci. Mater. Int. 28 (2018) 718–723.
https://doi.org/10.1016/j.pnsc.2018.10.003.

125
[175] X. Wang, Y. Liu, T. Shi, Y. Wang, Strain rate dependence of mechanical property in a
selective laser melted 17–4 PH stainless steel with different states, Mater. Sci. Eng. A.
792 (2020) 139776. https://doi.org/10.1016/j.msea.2020.139776.

[176] L. Tian, Z.M. Jiao, G.Z. Yuan, S.G. Ma, Z.H. Wang, H.J. Yang, Y. Zhang, J.W. Qiao,
Effect of Strain Rate on Deformation Behavior of AlCoCrFeNi High-Entropy Alloy by
Nanoindentation, J. Mater. Eng. Perform. 25 (2016) 2255–2260.
https://doi.org/10.1007/s11665-016-2082-8.

[177] T. LEBRUN, K. TANIGAKI, K. HORIKAWA, H. KOBAYASHI, Strain rate sensitivity


and mechanical anisotropy of selective laser melted 17-4 PH stainless steel, Mech. Eng.
J. 1 (2014) SMM0049–SMM0049. https://doi.org/10.1299/mej.2014smm0049.

[178] A. Nawaz, W.G. Mao, C. Lu, Y.G. Shen, Mechanical properties, stress distributions and
nanoscale deformation mechanisms in single crystal 6H-SiC by nanoindentation, J.
Alloys Compd. 708 (2017) 1046–1053. https://doi.org/10.1016/j.jallcom.2017.03.100.

[179] S.R. Jian, J.Y. Juang, Y.S. Lai, Cross-sectional transmission electron microscopy
observations of structural damage in Al0.16Ga0.84N thin film under contact loading, J.
Appl. Phys. 103 (2008). https://doi.org/10.1063/1.2836939.

[180] S. Sinha, R.A. Mirshams, T. Wang, S.S. Nene, M. Frank, K. Liu, R.S. Mishra,
Nanoindentation behavior of high entropy alloys with transformation-induced plasticity,
Sci. Rep. 9 (2019) 1–11. https://doi.org/10.1038/s41598-019-43174-x.

[181] C.K. Dolph, D.J. da Silva, M.J. Swenson, J.P. Wharry, Plastic zone size for
nanoindentation of irradiated Fe—9%Cr ODS, J. Nucl. Mater. 481 (2016) 33–45.
https://doi.org/10.1016/j.jnucmat.2016.08.033.

[182] W.B. Lee, Y.P. Chen, Simulation of micro-indentation hardness of FCC single crystals
by mechanism-based strain gradient crystal plasticity, Int. J. Plast. 26 (2010) 1527–1540.
https://doi.org/10.1016/j.ijplas.2010.01.011.

[183] H. Fei, A. Abraham, N. Chawla, H. Jiang, Evaluation of micro-pillar compression tests


for accurate determination of elastic-plastic constitutive relations, J. Appl. Mech. Trans.
ASME. 79 (2012) 1–9. https://doi.org/10.1115/1.4006767.

[184] H. Li, T. Zhu, N. Takata, M. Kobashi, M. Yoshino, Thermal activation process of plastic
deformation in Fe–18Cr single-crystal micropillars with high-density dislocations, Mater.
Sci. Eng. A. 819 (2021) 141459. https://doi.org/10.1016/j.msea.2021.141459.

[185] R. Niu, X. An, L. Li, Z. Zhang, Y.W. Mai, X. Liao, Mechanical properties and
deformation behaviours of submicron-sized Cu–Al single crystals, Acta Mater. 223
(2022). https://doi.org/10.1016/j.actamat.2021.117460.

[186] C.D. Hardie, S.G. Roberts, Nanoindentation of model Fe-Cr alloys with self-ion
irradiation, J. Nucl. Mater. 433 (2013) 174–179.
https://doi.org/10.1016/j.jnucmat.2012.09.003.

126
[187] Z. Jiao, G.S. Was, Impact of localized deformation on IASCC in austenitic stainless
steels, J. Nucl. Mater. 408 (2011) 246–256.
https://doi.org/10.1016/j.jnucmat.2010.10.087.

[188] A. Reichardt, A. Lupinacci, D. Frazer, N. Bailey, H. Vo, C. Howard, Z. Jiao, A.M.


Minor, P. Chou, P. Hosemann, Nanoindentation and in situ microcompression in different
dose regimes of proton beam irradiated 304 SS, J. Nucl. Mater. 486 (2017) 323–331.
https://doi.org/10.1016/j.jnucmat.2017.01.036.

[189] C.M. Sánchez Camargo, Mechanical multi-scale characterization of metallic materials by


nanoindentation test, 2019. http://www.theses.fr/2019ESAE0010/document.

[190] D. Tumbajoy-Spinel, S. Descartes, J.M. Bergheau, V. Lacaille, G. Guillonneau, J.


Michler, G. Kermouche, Assessment of mechanical property gradients after impact-based
surface treatment: Application to pure α-iron, Mater. Sci. Eng. A. 667 (2016) 189–198.
https://doi.org/10.1016/j.msea.2016.04.059.

[191] L.J. Yu, H.W. Yen, J.Y. Wu, J.J. Yu, C.R. Kao, Micromechanical behavior of single
crystalline Ni3Sn4 in micro joints for chip-stacking applications, Mater. Sci. Eng. A. 685
(2017) 123–130. https://doi.org/10.1016/j.msea.2017.01.004.

[192] I.N. Sneddon, The relation between load and penetration in the axisymmetric boussinesq
problem for a punch of arbitrary profile, Int. J. Eng. Sci. 3 (1965) 47–57.
https://doi.org/https://doi.org/10.1016/0020-7225(65)90019-4.

[193] J.S. Weaver, S. Pathak, A. Reichardt, H.T. Vo, S.A. Maloy, P. Hosemann, N.A. Mara,
Spherical nanoindentation of proton irradiated 304 stainless steel: A comparison of small
scale mechanical test techniques for measuring irradiation hardening, J. Nucl. Mater. 493
(2017) 368–379. https://doi.org/10.1016/j.jnucmat.2017.06.031.

[194] G. Dehm, B.N. Jaya, R. Raghavan, C. Kirchlechner, Overview on micro- and


nanomechanical testing: New insights in interface plasticity and fracture at small length
scales, Acta Mater. 142 (2018) 248–282. https://doi.org/10.1016/j.actamat.2017.06.019.

[195] Q. Ma, D.R. Clarke, Size dependent hardness of silver single crystals, J. Mater. Res. 10
(1995) 853–863. https://doi.org/DOI: 10.1557/JMR.1995.0853.

[196] J.G. Swadener, E.P. George, G.M. Pharr, The correlation of the indentation size effect
measured with indenters of various shapes, J. Mech. Phys. Solids. 50 (2002) 681–694.
https://doi.org/10.1016/S0022-5096(01)00103-X.

[197] H. Zhang, B.E. Schuster, Q. Wei, K.T. Ramesh, The design of accurate micro-
compression experiments, Scr. Mater. 54 (2006) 181–186.
https://doi.org/10.1016/j.scriptamat.2005.06.043.

[198] C. Kirchlechner, J. Keckes, C. Motz, W. Grosinger, M.W. Kapp, J.S. Micha, O. Ulrich,
G. Dehm, Impact of instrumental constraints and imperfections on the dislocation

127
structure in micron-sized Cu compression pillars, Acta Mater. 59 (2011) 5618–5626.
https://doi.org/10.1016/j.actamat.2011.05.037.

128

You might also like