These
These
L’INSA Lyon a mis en place une procédure de contrôle systématique via un outil de
détection de similitudes (logiciel Compilatio). Après le dépôt du manuscrit de thèse,
celui-ci est analysé par l’outil. Pour tout taux de similarité supérieur à 10%, le manuscrit
est vérifié par l’équipe de FEDORA. Il s’agit notamment d’exclure les auto-citations, à
condition qu’elles soient correctement référencées avec citation expresse dans le
manuscrit.
Par ce document, il est attesté que ce manuscrit, dans la forme communiquée par la
personne doctorante à l’INSA Lyon, satisfait aux exigences de l’Etablissement concernant
le taux maximal de similitude admissible.
1
ScSo : Histoire, Géographie, Aménagement, Urbanisme, Archéologie, Science politique, Sociologie, Anthropologie
Cette thèse aura représenté pour moi, au-delà du projet scientifique, trois années importantes
et une formidable source d’évolution sur le plan personnel. A toutes les personnes qui ont rendu
cette expérience unique, riche en apprentissages et un très beau souvenir pour toute la suite, je
tiens à leur dire de tout cœur merci.
Merci d’abord à mes formidables encadrants de thèse qui m’ont accompagné tout du long,
et qui m’ont aidé à faire mes premiers pas dans ce beau domaine qu’est la recherche. Merci de
m’avoir soutenu dans mes démarches et les voies que j’ai voulu emprunter, qui devaient pourtant
vous sembler assez obscures par moment... tout en m’aidant à me replacer parfois pour m’éviter
de m’égarer dans l’infinité des possibles.
Mon choix de faire cette thèse s’est décidé je pense que tu t’en souviens Jean-Yves dans
les combles d’une maison à la campagne en plein confinement Covid-19, pendant un échange
par visio. Quelques dizaines de minutes, cela aura été suffisant pour voir toute la simplicité,
bienveillance et autres qualités qui font de toi un excellent encadrant de thèse, en plus d’une belle
personne qui est source d’inspiration et de bonne humeur pour tout le monde. Cette simplicité
aura eu raison d’un bout de mon trop de sérieux et de rigidité, et ce n’est pas peu dire ! Merci
pour tous les échanges que l’on a pu avoir, au coin d’un tableau blanc ou d’une tasse de chicorée.
Guilhem, je pense que j’ai rarement eu l’occasion de travailler avec une personne aussi fiable,
constante et efficace (pense à prendre du temps pour toi aussi !). Que ce soit lors de notre
mésaventure sur les superalliages base Nickel ou plus récemment, tu as toujours répondu présent
et proposé ton aide quand il y en avait besoin... alors que ton emploi du temps bien rempli aurait
facilement pu t’en dissuader. Merci aussi pour tes retours détaillés et toujours pertinents, qui
ont forgé aussi mes qualités scientifiques durant ces trois années.
Rémy, cela fait un moment maintenant que l’on se connait, et j’ai apprécié de t’avoir autant en
tant que professeur (et responsable du département SIM :)), encadrant de projet qu’encadrant de
thèse. À y réfléchir, mes questionnements qui frôlaient parfois la métaphysique des matériaux
ont peut-être dû te faire poser quelques questions ou avoir des sueurs froides lorsque tu as
appris que tu allais m’encadrer pour trois ans au complet :) En espérant que ma ténacité qui
frôlait parfois l’entêtement n’aura pas dépassé votre limite d’endurance, et qu’au final ces trois
années ont été pour toi une bonne expérience comme elles l’ont été pour moi. Tes jeux de
mot aussi subtils que bien placés ont toujours su créer de beaux moments lors de nos réunions
d’avancement, desquelles je garde de très bons souvenirs. Vous formez une très belle équipe, qui
j’espère aura l’occasion de vivre encore d’autres aventures au cœur des mystères de la fatigue...
Remerciements | v
Plus largement, j’aimerais remercier l’ensemble des personnes que j’ai pu rencontrer au lab-
oratoire MatéIS, l’un des endroits les plus inspirants où il m’a été donné de travailler. Merci à
tous mes cobureaux : Masato et Amin, Sébastien et Maureen, Louis Lesage, Justine Taurines
et Théophile... Merci pour l’ambiance inspirante et la bonne humeur pendant ces trois années,
qui m’aidaient quand j’étais trop pris par les difficultés techniques et que je pouvais en perdre
de vue l’essentiel. Une pensée en particulier à Louis H. avec qui j’ai eu la chance de partager le
même maître/directeur de thèse et la période de rédaction, avec aussi des expériences uniques
au synchrotron... preuves d’ailleurs que c’est tout à fait possible de bien vivre ses expériences
et de passer de très bons moments même quand les résultats ne sont pas au rendez-vous :) En
parlant de synchrotron, merci au passage à James Ball et Thomas Connolley pour l’opportunité
de faire de la 3D DRX et de la DCT à l’ESRF, ainsi que pour m’avoir proposé plusieurs scans
de tomographie X à Diamond.
Pour en revenir au laboratoire, merci à Xavier, Sophie et Sylvain pour vos discussions sur le
projet Aéroprint ; autant de personnes avec qui j’ai eu beaucoup de plaisir à partager pendant
ce projet et dont j’ai aussi apprécié toutes les qualités scientifiques. Merci à tous les doctorants,
post-doctorants, stagiaires, permanents pour les moments de partage et d’entraide, les parties
de babyfoot ou de tennis, les 16h de vendredi, les team buildings, la période aux algécos qui
vi | Remerciements
Ces trois années de thèse n’auraient évidemment pas été aussi riches et vivantes sans la
présence de toutes les personnes qui m’ont entouré au-delà du travail. Merci à ma famille
pour leur soutien, ainsi que pour tout ce qu’ils m’ont apporté et qui m’a permis d’en arriver là
aujourd’hui.
Merci à tous les amis de la Protection Civile : Stéphanie, Timon, Paul, Zachary, Lounès,
Clotilde, Thibault, Dimitri, Marc, Grégoire, Christopher... la liste est encore longue ! Cette
année et quelques avec vous aura été une expérience aussi intense qu’enrichissante, stimulante,
faite de partages et de belles rencontres, de galères et d’entraide... le complément idéal pour
redescendre les deux pieds sur terre dans les périodes d’intense réflexion, de code ou de rédaction
:)
Un profond merci enfin à l’association UCM et toutes les personnes qu’elle m’a permis de ren-
contrer. Je n’aurais probablement pas été capable de commencer cette thèse sans toute l’aide,
les compréhensions, le soutien et l’inspiration que vous m’avez apporté et qui continuent de
m’animer aujourd’hui. Merci à tous les professeurs qui ont su m’accompagner avec bienveillance
et profondeur : Anthony, Thibaut, Christiane, Alexis, Guillaume... Merci aussi pour les mo-
ments de partage avec tous les amis d’UCM en France, qui sont une profonde source de remise
en question, d’inspiration, de joie et de soutien dans les moments de grand questionnement.
Remerciements | vii
Manufacturing defects are known to significantly impact the fatigue properties of additively
manufactured (AM) components. Notably, the large surface roughness, typical of AM processes,
leads to increased stress concentrations that promote crack initiation, thereby reducing fatigue
resistance.
Traditionally, surface roughness and related defects are evaluated using tools like white light
interferometers. However, these instruments offer a limited, single-perspective and only partially
three-dimensional analysis. Those limitations do not enable the thorough characterization of the
complex surfaces and hidden defects typical in AM components.
This study describes a methodology for performing a 3D surface analysis using X-ray Com-
puted Tomography (XCT) data. The method is illustrated on various samples, ranging from
simple cylinders to more intricate architected structures. It turns out to be very efficient at
detecting critical surface defects, such as notches hidden by partially melted powder particles.
The methodology is then applied to examine the effect of surface defects on the fatigue
properties of Ti64 produced by Laser Powder Bed Fusion (L-PBF). This analysis includes both
as-built surfaces and those subjected to post-treatments, specifically investigating the impact
of Plasma electrolytic Polishing (PeP) and surface oxygen contamination (presence of an α-case
layer) resulting from high-temperature heat treatment (860 °C).
Using XCT for 3D characterization, defects responsible for fatigue failure are identified,
the latter being predominantly surface valleys. The method’s ability to predict crack initiation
locations is also evaluated, as well as its potential to estimate the fatigue resistance of a specimen
before testing.
Summary | ix
Les défauts de fabrication sont connus pour avoir un impact significatif sur les propriétés
de fatigue des composants élaborés par Fabrication Additive (FA). En particulier, l’importante
rugosité de surface, typique des procédés de FA, génère des concentrations de contrainte locales
qui favorisent l’amorçage de fissures, réduisant ainsi la résistance à la fatigue.
Traditionnellement, la rugosité de surface et les défauts associés sont caractérisés à l’aide
d’outils tels que les interféromètres à lumière blanche. Toutefois, ces instruments offrent une
analyse qui n’est que partiellement tridimensionnelle car limitée à une unique perspective.
Ces limitations ne permettent pas une caractérisation approfondie des surfaces de composants
élaborés par FA, qui peuvent avoir des géométries 3D complexes et des défauts de surface cachés
derrière un importante rugosité.
Ce manuscrit décrit une méthodologie permettant d’effectuer une analyse de surface en 3D à
partir d’une caractérisation par tomographie aux rayons X. La méthode est illustrée sur divers
échantillons, allant de simples cylindres à des structures architecturées plus complexes. Elle
s’avère efficace pour détecter les défauts de surface critiques, tels que les entailles cachées par
des particules de poudre partiellement fondues.
La méthodologie est ensuite appliquée pour étudier l’impact des défauts de surface sur la
tenue fatigue du Ti64 élaboré par fusion laser sur lit de poudre. Cette analyse porte à la
fois sur les surfaces brutes de fabrication ou après différents post-traitements, en étudiant en
particulier l’impact d’un polissage électrolytique plasma et de la contamination par l’oxygène
de la surface (présence d’une couche d’α-case) résultant d’un traitement thermique à haute
température (860 °C).
L’utilisation de la tomographie aux rayons X pour la caractérisation 3D des défauts a permis
d’identifier ceux responsables de la rupture par fatigue, ces derniers étant principalement des
vallées en surface. La capacité de la méthode à prédire le site d’amorçage des fissures est
également évaluée, ainsi que son potentiel à estimer la résistance à la fatigue d’un spécimen
avant essai.
Résumé | xi
Remerciements v
Summary ix
Résumé xi
Contents xiii
Introduction 1
Contents | xiii
3 Materials characterization 71
3.1 Samples manufacturing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.1.1 Processing conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.1.2 Fatigue and tensile specimens . . . . . . . . . . . . . . . . . . . . . . . . . 73
3.2 Microstructure characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.2.1 Microstructure of the bulk material . . . . . . . . . . . . . . . . . . . . . . 76
3.2.2 Microstructure near the surface . . . . . . . . . . . . . . . . . . . . . . . . 78
3.3 Tensile properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
3.4 Defects characterization using XCT . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.4.1 Acquisition setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.4.2 Porosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
3.4.3 Surface defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
Intermediate summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
xiv | CONTENTS
References 211
CONTENTS | xv
Introduction | 1
Figure 2 | Surfaces inherited from AM processes. (a) Surface of a Ti64 thin strut
fabricated by Electron Powder Bed Fusion (E-PBF), adapted from [PER 19]. The surface is
reconstructed from an X-ray Computed Tomography (XCT) scan. (b) The surface of a Ti64
sample manufactured by Laser Powder Bed Fusion (L-PBF), observed using Scanning Electron
Microscopy (SEM) [NAK 19].
2 | Introduction
Introduction | 3
4 | Introduction
Introduction | 5
Contents
1.1 An introduction to fatigue . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2 Defects inherited from L-PBF: application to Ti64 . . . . . . . . . . . . 11
1.2.1 A classification of defects based on their spatial distribution . . . . . . . . . 12
1.2.2 Surface defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.2.3 Internal defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.3 Impact of different defects on fatigue properties . . . . . . . . . . . . . . 16
1.3.1 Identifying the defects at the origin of crack initiation . . . . . . . . . . . . 16
1.3.2 A quantitative comparison of the impact of surface vs internal defects . . . 19
1.4 Modeling the impact of defects on fatigue properties . . . . . . . . . . . 23
1.4.1 A brief exploration of theory: Kt , Kf , K... which difference? . . . . . . . . 23
1.4.2 Estimation of defects harmfulness in the case of L-PBF Ti64 . . . . . . . . 27
1.5 Towards a 3D local characterization of AM surfaces . . . . . . . . . . . 34
Intermediate summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
This chapter aims to provide the required insights to aid in comprehending the work con-
ducted during this PhD. It primarily focuses on the core research topic − assessing how defects
affect fatigue properties of Ti-6AL-4V (Ti64) alloy manufacturing by Laser Powder Bed Fusion
(L-PBF).
Ample literature exists on this subject, providing valuable insights that guide research and
highlight areas for improvement. Nonetheless, it seems also advisable to use previous studies
with caution. Results obtained on L-PBF Ti64 exhibit variability − e.g. due to the large choice
of machines and process parameters that have been used − and improvements in the L-PBF
technology have occurred over the years. Consequently, conclusions from earlier studies might
not directly apply today.
|7
UTS
LCF
HCF
Run-out
σNf
Infinite life
log N
Nf
Figure 1.1 | Schematic of a typical Wöhler (or S-N) curve. Noteworthy stress levels
− the tensile strength UTS and the fatigue limit − are indicated on the stress axis. The three
main regions − LCF, HCF and infinite life − are delimited.
Each Wöhler curve is measured under given conditions for the fatigue tests. These conditions
can have a strong impact on the results. The main parameters describing a fatigue test and that
have to be taken into account are the following:
σmin
• The stress ratio R “ σmax
• The mean stress σm “ σmin `σ
2
max
stress
1/f
σmax
σa
σm time
σmin
Figure 1.2 | Graphical definition of minimum and maximum and mean stresses during fatigue
cycling σmin , σmax , and σm , of the stress amplitude σa and the frequency f .
An introduction to fatigue | 9
The crack initiation period can be split into several steps. It begins with local cyclic slip,
i.e. local plastic deformation at each loading cycle. The regions where cyclic slip preferentially
occurs depend on the local stress field, which depends itself on the defects that act as stress
concentrators. It is also influenced by the local microstructure since cyclic slip will be facil-
itated in grains with slip systems oriented unfavorably with respect to the loading direction
[TUB 23, BOO 16, WU 23]. Cyclic slip eventually leads to the nucleation of a micro-crack,
which will continue to grow under cyclic loading. At this step, crack growth is however still de-
pendent on the environment of the initiation site − e.g. the local microstructure and stress field.
Grain boundaries, or α{β interfaces in Ti alloys, can for instance act as barriers for micro-crack
propagation. Depending on the local stress gradient, the crack may slow down at these obstacles
or even stop definitely. For these reasons, micro-crack growth is still considered to be part of
the initiation process. Only when crack growth becomes dependent on bulk properties alone, it
can be considered a macro-crack.
Nonetheless, the limit between micro-crack and macro-crack propagation − or in other words,
between crack initiation and growth − is highly dependent on the material studied. For instance,
for a given material without any defect, the majority of the fatigue life may be spent to initiate
the crack − see Fig.1.4. It might however be very different if defects such as inclusions, pores or
surface scratches are present [ZER 19]. In the particular case of Additive manufacturing (AM),
typical defects include pores, lacks of fusion or surface valleys − see next section for more details.
Although they are different in nature, they will all in a similar manner act as stress concentration
raisers and thus promote local plastic deformation. As a consequence, crack initiation may occur
at the very beginning of cycling, bypassing a wide percentage of the fatigue life that was spent to
initiate a crack. As a result, the fatigue strength will be drastically reduced. Thus, defects can
have a strong impact on fatigue strength by facilitating the crack initiation stage.
starting from
log (crack length) defect
micro-cracks
starting from
no defect
nucleation
Figure 1.4 | Different scenarios of fatigue crack growth depending on whether crack
initiation occurs at a defect or not. If a sufficiently large defect is present, almost all fatigue life
corresponds to macro-crack propagation, while it is the inverse in the case where no defect is
present. Adapted from [SCH 09, p.22].
Figure 1.5 | Increase in sub-surface pore density with a decrease in laser scanning
speed in the contour scan. The figure is adapted from [KAR 23]. Samples are fabricated by
L-PBF using pre-alloyed A20X aluminum powder.
Sub-surface defects are also of particular interest regarding their impact on mechanical prop-
erties. Indeed, the same defect will be much more harmful if it is at a sufficiently
close distance from the surface than if it is in the core, as is discussed in Section 1.3.
Finally, sub-surface defects have the advantage of being the only internal defect that can be
detected by some Non-Destructive Testing (NDT) methods such as Eddy Current, or eliminated
by some surface treatments such as machining or chemical polishing. However, regarding this
last point, surface treatments involving material removal may also transform these defects into
surface defects, making them even more harmful [PER 20].
Figure 1.7 | Fatigue failure from two surface spatters in IN625 manufactured by
L-PBF [PES 20, Fig.3.28]. It can be seen that the spatter at the origin of the largest crack
caused the formation of a lack of fusion.
Figure 1.8 | XCT slice showing an example of dross formation in the down-skin region
of a 2 mm cylinder fabricated by Electron Powder Bed Fusion, adapted from [PER 19].
Figure 1.9 | Lacks of fusion in L-PBF Ti64 observed using SEM by [PAL 18]
Figure 1.10 | Keyhole porosity observation in L-PBF Ti64. (a) XCT slice showing
keyhole porosity, from [BAY 19]. Keyhole pores are visible all along the last laser track, at the
top of the image. (b) SEM image of a non-spherical keyhole pore, from [PAL 18].
Figure 1.11 | Fatigue crack initiation at surface notches in L-PBF Ti64 [KAS 15]
.
the direction of applied stress. r is the radius of the pore, whereas d is the distance of the center
of the pore to the surface. Therefore, the pore touches the surface for r{pr ` dq “ 0.5.
Another particularity of surface defects is that they are exposed to the external environment.
Conversely, internal defects are in contact with vacuum or inert gas. It has been shown that
the presence of air can have a deleterious effect on the fatigue life of the Ti64 alloy and in
particular on the threshold stress intensity factor Kth , which is the minimum stress intensity
factor required to induce a significant crack growth [XU 21, JUN 21, IRV 74].
Crack initiation on an internal defect can sometimes be seen, even on samples with an as-built
surface [LIU 14]. This was for example observed in 3 cases out of 13 in the study of [CHA 18]
on the fatigue of L-PBF and E-PBF Ti64.
In such cases, the defect is often located in the sub-surface region [LIU 14, AND 19b]. As an
example, [TAM 17] observed in Ti64 manufactured by E-PBF that for 10 out of 11 cases where
initiation was found to occur from a pore, the pore was located in the sub-surface region. Like
surface defects, sub-surface defects are therefore particularly harmful. The reasons are
also similar. FEM calculations introduced previously in Fig.1.12 show that the stress concentra-
tion induced by an internal defect increases when its distance to the surface decreases. Such a
stress concentration promotes crack initiation and can provoke the rapid failure of the ligament
that separates the pore from the surface. This creates a large defect at the surface, which will
be particularly harmful for the reasons invoked in the previous paragraph.
When a specimen is machined, surface defects − which are the most harmful − are removed.
Consequently, the remaining internal defects become the most critical ones and are often found
to be the cause of fatigue failure [KAS 15, Fig.15c]. This effect is more pronounced for samples
with polished surfaces since in the case of machined surfaces, grooves are still present at the
surface and can promote crack initiation [YU 19].
Among all internal defects, lacks of fusion are generally identified as the most
critical ones [GON 15, PLE 20c, SAN 20, LAR 18]. This is due to their elongated shape and
larger size. For example, [CHA 18] observed in the case of machined L-PBF Ti64 specimens 25
initiations on lack of fusion defects against 12 on smaller internal defects such as spherical pores.
The effect is even more pronounced when specimens are printed with their axis parallel to the
build direction, since lacks of fusion are in this case perpendicular to the loading direction.
Similarly, by comparing the mechanical behavior of L-PBF Ti64 manufactured with different
process parameters (excessive vs insufficient energy input), [GON 15] concluded that keyhole
pores − that are small and spherical − do not significantly impact mechanical properties even
for volume fractions of the order of 1%. On the opposite, lacks of fusion generated by an
insufficient energy input − that are larger and elongated − already lead to a substantial decrease
of mechanical properties for volume fractions of 1%. This effect is visible for both tensile and
fatigue properties.
According to the study of [AND 19b] dedicated to the fatigue properties of 316L stainless
steel fabricated by L-PBF, the first-order factor remains the proximity of the defect to the
surface. The defect responsible for the failure, located in the sub-surface region, had thus in
several cases a size much smaller (up to 3.6 times) than that of the largest defect observed in
the core of the sample.
To summarize, it is finally possible to give a first qualitative classification of the types of
defects by order of increasing harmfulness in fatigue for Ti64:
Lacks of fusion > Spherical pores
Surface notches >
Sub-surface > Core
Since machining or polishing enables the removal of surface defects and HIP allows the
removal of internal ones, the combination of these conditions enables decorrelating the influence
of these two families of defects.
Of course, the conclusions must be nuanced for several reasons:
• Machining/polishing without HIP will bring defects that were previously located in the
sub-surface directly at the surface. Even if surface notches present in as-built surfaces are
removed, new surface defects can therefore be created [CHI 21, CAR 19].
• Machining leaves residual stresses at the surface that can have a strong influence on the
mechanical behavior. Typically, if compressive stresses are generated, they will inhibit
crack initiation at the surface.
• HIP treatments − carried out at very high temperatures („900 °C) − will cause significant
modifications in the microstructure such as broadening of α-laths, which may influence
fatigue properties.
• Not all pores are necessarily closed after HIP [PLE 20b, LI 19] and it has therefore been
observed in some cases that crack initiation occurs on a partially closed pore [CHA 17].
Nonetheless, many studies report the influence of defects on fatigue properties by comparing
the above conditions and most of them show good agreement with each other. Among others,
the work of [MAS 18] is representative of most of the published results. The fatigue performance
of the as-built, HIP, machined, and HIP+machined conditions was compared in rotary bending
fatigue tests. Comparing these different conditions, [MAS 18] arrived at the following ranking
in terms of performance:
AB surface w/o HIP ă AB surface with HIP ! Polishing w/o HIP ! Polishing with HIP
where AB is an abbreviation for as-built.
Figure 1.13 | Wöhler curves obtained with or without polishing and HIP of L-PBF
Ti64 [MAS 18] (60Hz rotational bending tests on vertically built specimens)
Before comparing the fatigue performances of the different conditions, we can already notice
in Fig.1.13 that the differences in fatigue performance between conditions are more pronounced
for longer fatigue lives. It is in this zone that crack initiation represents the most important
fraction of fatigue life (compared to the crack propagation regime); the influence of defects
therefore tends to be exacerbated.
The Wöhler curves also generally show more scattering on machined or polished specimens
than on as-built surface specimens. This is slightly noticeable in Fig.1.13, but the effect can
be even more pronounced − see for example [WYC 13]. This difference in behavior can be
attributed to the fact that in the as-built condition, there is a very large statistical representation
of surface defects in each specimen. Conversely, on machined or polished specimens, the critical
defect density is lower and therefore there may be statistically more variation in fatigue life from
one specimen to another.
We can see that the beneficial effect of HIP is limited on specimens with an as-built
surface: the fatigue strength at 107 cycles σ107 is 155 MPa with an as-built surface without HIP
and 195 to 220 MPa in the as-built + HIP condition − which represents an improvement of
25-42%. Once again, we see the most deleterious effect of the surface defects present for
an as-built surface in comparison with the effect of the internal defects suppressed by HIP. This
effect can be partly attributed to the fact that tests are performed in rotational bending, thus
favoring crack initiation at the surface. However, the same tendency can be found in several
other studies with tests not carried out by rotary bending (see e.g. references given in Tab.1.1):
greater harmfulness of surface defects is observed.
Figure 1.14 | (a) Wöhler curves obtained for varying surface states ("Net" = as-built) and
build directions by [EDW 14] (R “ ´0.2). (b) and (c) Optical microscope characterization of
the corresponding material, showing a high pore density [EDW 14].
Surface Build
Study Heat treatment R σf estimate (MPa)
state direction
as-built Z SR -1 155
Table 1.1 | Fatigue limit σf estimations for L-PBF Ti64 in several studies , for varying
stress ratios R, surfaces states, heat treatments (SR = Stress Relieving) and orientations (Z K
build plate). Most of the time, the fatigue limit is estimated by the fatigue strength at 107 cycles.
Figure 1.15 | Different stages in fatigue life and dedicated physical quantities.
Adapted from [SCH 09, p.15].
σpeak
σnominal
σpeak
Kt “ (1.1)
σnominal
Elevated stress concentration triggers unusually high localized stresses. This, in turn, can
activate cyclic slip and potentially lead to crack initiation. Hence, Kt can be regarded as a useful
parameter for assessing crack initiation. From this perspective, the influence of defects
can be quantified by considering their role as stress and strain enhancers.
σpeak
d
ρ
d
d
σpeak Kt “ 1 ` 2
ρ
(1.2)
def Kf ´ 1
q “ (1.4)
Kt ´ 1
However, such an expression does not explain the physics behind Kf : it somehow hides the
question behind another parameter, q. Moreover, Kf actually depends not only on the material
but also on other parameters such as the stress gradient −i.e. how fast the tensile stress decreases
in the material near the defect −which, unlike Kt , depends on the notch size.
Again, many expressions have been proposed to refine the Kf ´ Kt relation so that it can
account for these effects. One can for example cite the work of [PET 59] which considers the
stress that has to be compared to the fatigue limit (of smooth specimens) is not the one at the
notch root, but the one at a certain distance beneath the surface. Using some approximations,
he arrives at Eq.1.5.
Kt ´ 1
Kf “ 1 ` (1.5)
1 ` α{ρ
where ρ is the radius of curvature at the notch root and α a material constant.
Works in this field gradually led to the definition of a broader theory, the Theory of Critical
Distances (TCD) [TAY 08]. The main idea is simple: to induce crack initiation, a stress
raiser has to increase stress not only at a surface hot spot but also over some critical
distance/volume beneath the surface. A possible justification for this is that for a crack
to initiate, it needs to overcome microstructural barriers such as grain boundaries or interfaces
with other phases during the early stages of crack propagation (i.e. small crack propagation).
If the stress fades away too fast from the stress raiser, the remaining stress may not be high
enough to overcome these obstacles [Val 00].
Several methods are proposed in the TCD to calculate this stress. The simpler is the point
method, which goes back to the idea of Peterson [PET 59]. It consists of taking the stress at
a given distance L{2 from the hot spot generated by the stress raiser. It is also possible to
calculate an average stress by integrating stress along a line, over an area or a volume. The
characteristic distance L is a material constant and can be calculated from the threshold stress
intensity factor Kth and the fatigue strength of a smooth specimen σ0 using Eq.1.6.
ˆ ˙2
1 Kth
L“ (1.6)
π σ0
C ¨ pHV ` 120q
σω “ ? (1.7)
area1{6
?
where σω is in MPa, HV is the Vickers hardness in kgf/mm2 , and area is the highest Murakami
parameter measured for the defects contained in the specimen, in µm. The constant C is equal
to 1.43 for a surface defect and 1.56 for an internal one.
rapidly reaches the same value as the one of the central crack.
The impact of defects on the fatigue properties is very often quantified using the Mu-
rakami parameter, which is the square root of the area of the defect projection along the
?
loading direction. It is usually noted area. [MUR 02] has shown that there is a good corre-
lation between this parameter and the SIF of a crack. In cases where it seems reasonable to
?
assimilate the defect to a crack, the area parameter seems therefore a good option to estimate
its harmfulness.
Thus, the Murakami parameter has been used in several materials to predict the impact of
various defects on the fatigue limit: inclusions in steels [MUR 02], but also artificial surface
defects in conventionally processed Ti64 [MAT 03]. Since then, many studies questioned its
validity on additively manufactured materials [PER 20, MAS 18, ROM 17, BAR 23, BER 17,
TAM 17].
Although the simple definition of the Murakami parameter as the square root of the projected
area is straightforward in the case of isolated core defects, some adaptations are needed for
particular cases. It is for instance the case to take into account the proximity of a defect to the
surface, the proximity of two neighboring defects or even the irregular morphology of defects.
?
Figure 1.19 | Definition of the effective Murakami parameter area eff in different
configurations [MAS 18]. The figures correspond to the projection of the defects in the plane
perpendicular to the stress. (a) Internal non-spherical defect. (b) Surface defect. (c) Sub-surface
defect. (d) Interaction between two adjacent defects. (e) Inclined defect in contact with the
surface
.
It can be noticed that this parameter allows − at least partially − to take into account the
greater harmfulness of internal defects located near the surface. Indeed, it can be seen that
the whole ligament connecting such a defect to the surface is considered mechanically inefficient
and part of the defect. This effect is taken into account when the distance of the defect to the
surface is smaller than its size. This is consistent with Fig.1.12 which showed that the stress
concentration induced by a spherical pore is higher when its distance to the surface is smaller
than its radius.
The Murakami parameter is often measured post-mortem on a fracture surface.
The defect responsible for failure can then be identified and its projected area measured (by
?
SEM for example). Typical area values of (internal and surface) defects responsible for failure
in materials manufactured by L-PBF are given thereafter. They give an order of magnitude of
the size of the defects that can cause failure, even if the studies in question concern machined
or polished specimens:
• [HU 20b] found values between 30 µm and 72 µm on 21 machined Ti64 specimens, with an
average of 61 µm.
• [MAS 18] measured values between 25 µm and 138 µm on 12 polished Ti64 specimens, with
an average of 49 µm.
• [AND 19b] found values between 42 µm and 109 µm on 10 polished 316L specimens, with an
average value of 69 µm (only the specimens manufactured with the best process parameters
are considered).
Figure 1.20 | Fatigue limit σw prediction for L-PBF and E-PBF polished specimens
using the Murakami parameter and Eq.1.7 [MAS 18]. The fatigue limit prediction from
the model corresponds to the black line, whereas experimental data is in red and blue. Note
that in this figure, the ordinate does not correspond directly to the stress amplitude during the
σa test but to the amplitude normalized by the factor (HV+120) found in the σω formula. The
acronym DMLS (Direct Metal Laser Sintering) refers here to L-PBF process and EBM (Electron
Beam Melting) refers to EBM.
In addition to estimating the impact of defects on the fatigue limit, it can be interesting to
evaluate their impact on fatigue life reduction in the finite life domain. Several studies have
for example evaluated the possibility of predicting the killer defect by characterizing the defect
?
population. [PER 20] proposed a prediction using the Murakami parameter area. For this
purpose, E-PBF Ti64 chemically polished specimens were scanned using XCT before testing and
?
the defects with the largest area were identified. Fatigue tests in uni-axial tension (R=0.1)
were then performed and the defects responsible for failure were identified.
Figure 1.21 | Attempts to predict the killer defect in E-PBF chemically polished
?
Ti64 fatigue specimens (R “ 0.1) using the area parameter [PER 20]. For each spec-
?
imen, the main defects detected using XCT are ranked according to their size − i.e. the area
parameter. The killer defects identified after fatigue testing are highlighted in red. Ten speci-
mens have undergone only chemical etching (in blue) while three did also undergo HIP treatment
(in green).
Another important limit of this approach is that it is limited to specimens whose defects
can be clearly delineated as shown in Fig.1.19a-e, as is the case for internal or surface pores.
In the case of L-PBF Ti64, it means that it can only be used for machined or polished specimens.
Indeed, as discussed in Section 1.3.1, these are the only cases where most of the cracks will initiate
on pores. Conversely, for specimens with an as-built surface, the defects at the origin of failure
are most often notch-like defects. These have more complex shapes and are difficult or even
impossible to delineate since they tend to be interconnected with each other (see Fig.1.22a-b for
an example of interconnected valleys). Other methods may therefore have to be used to
characterize the mechanical severity of valley-type defects.
?
The study of [NAK 19] also suggests that the area parameter as defined in Fig.1.19a-e
is not suited to characterize specimens with as-built surfaces. In this work, the fatigue limits
of L-PBF and E-PBF Ti64 specimens with as-built surfaces have been estimated. They were
found to be of 155 MPa and 140 MPa respectively (for non-HIPed specimens). According to the
observed fractured specimens in [NAK 19], failure occurs at the surface for as-built surfaces.
Furthermore, no obvious and clearly delineated defects could be found on the SEM fracture
?
surface. Eq.1.7 has then been used to estimate the area parameter of the hypothetical defect
leading to the experimentally measured fatigue limit. The obtained values are 16 505 µm and
9635 µm respectively, which was very far from reality.
To overcome this issue, the authors used semi-analytical formulas to estimate an equivalent
? ?
area (denoted areaR ) parameter based on roughness parameters measurement − see Eq.1.8
and Eq.1.9. These formulas were proposed by [MUR 02] for surfaces with periodical notches of
given depth and separated by a given distance. To use these formulas in the case of as-built sur-
faces, [NAK 19] proposed to replace the notch depth and the notch-to-notch distance by Ra and
?
RSm roughness parameters respectively. By doing so, they could find more consistent areaR
values (264 µm and 203 µm for E-PBF and L-PBF specimens respectively). Such approaches,
based on roughness parameter measurements, are detailed in the next section.
? ˆ ˙ ˆ ˙2 ˆ ˙3
areaR Sz Sz Sz Sz
“ 2.97 ´ 3.51 ´ 9.51 , for ď 0.195 (1.8)
RSm RSm RSm RSm RSm
?
areaR Sz
« 0.38, for ě 0.195 (1.9)
RSm RSm
where Sz is the maximum surface height [Int 21] and RSm the mean width of profile elements.
?
To conclude, the area parameter (with its initial definition) yields interesting results −
even if still perfectible in several cases − to account for the severity of internal or surface pores
which can be clearly delineated. However, samples with as-built surfaces require other solutions.
Since surface defects are of primary importance in the case of as-built L-PBF Ti64 specimens,
many studies have been published on the relation between surface state and fatigue properties.
Most often, the surface state is characterized using standard roughness parameters, whose def-
inition can be found in standards such as [Int 21]. In many cases, the only parameter used is
the arithmetic average roughness (Ra in 2D and Sa for areal measurements).
For instance, [MIA 21] tried to correlate the fatigue limit of E-PBF Ti64 specimens from
several studies with Ra , using a simple curve fitting. Even though a reasonable fit is found,
such relations tend to be valid only in particular cases and do not provide tools that can be
used in a broader context (different materials, manufacturing processes, build directions, etc.).
Furthermore, only a few studies are compared in [MIA 21], which makes the fit questionable.
In general, the Ra parameter is therefore not considered to be a good candidate
to correlate with fatigue performance [ZAB 18, GOC 19], especially for as-built surfaces
of AM materials which often show much more complex geometries than more usual machined
surfaces. This is understandable since Ra only reflects the average surface state, while fatigue
properties are rather determined by the few most critical defects present in the material.
Several studies − concerning both AM and conventional machined materials − conclude
therefore that a parameter such as the maximum valley depth (Rv /Sv ) is better suited
[TAY 91, GOC 19]. However, even in the case of [GOC 19] who found a correlation between Sv
and fatigue life, it can be seen that a lot of scatter remains − see Fig.1.23a-d.
Figure 1.23 | Correlation between roughness parameters Sa /Sv and fatigue life of
L-PBF IN718 specimens, taken from [GOC 19]. Fatigue tests were carried out at R=0.1
and σmax “ 900 MPa with a frequency of 5 Hz. (a) Sa measured by Structured Light scanning
(SL) (b) Sa measured by XCT (c) Sv measured using by SL) (d) Sv measured by XCT.
ˆ ˙ˆ ˙
Ra Ry
Kt “ 1 ` n (1.10)
ρ Rz
where Ra , Ry and Rz are the arithmetic average roughness, peak-to-valley height and 10-point
roughness respectively. ρ is the average radius determined from the dominant profile valleys.
n “ 2 for uniform tension and n “ 1 for shear loads.
[POM 20] used similarly an estimation of Kf using roughness and curvature measurements,
but in a more local approach. The objective was to try predicting, on top of the fatigue limit, the
area of crack initiation. To achieve this result, the surface of the specimen is first segmented into
1 mm squares. For each of them, Kf is calculated from the maximum depth of the valley in this
zone and the radius of curvature at the bottom of the notch, determined using Focus Variation
microscopy (FV). Fig.1.24 shows that with this methodology, the area of crack initiation −
marked by a red cross) − could be effectively identified.
Figure 1.24 | Kf ,s map computed using focus variation microscopy on a cast alu-
minum bending fatigue specimen [POM 20]. The crack initiation site is highlighted by
the red cross and is found to occur in the area of maximum Kf,s .
Notches hidden
Perpendicular by a spatter or
notch Oblique notch powder particles
2.5D
3D
by XCT
Figure 1.25 | Schematic showing the interest of characterizing AM as-built surfaces
in 3D using XCT. Several examples are shown, corresponding to different surface defects. The
orange lines show surfaces as seen by a 2.5D characterization tool which probes the surface from
a single point of view and without the ability to look through matter. The blue lines correspond
to the true surfaces, i.e. the one which would ideally be obtained through XCT using a high
resolution, with no noise nor artifacts. In two of the three examples shown, the 2.5D surface
characterization fails to account for notches because they are hidden by other surface features.
1
Note that the term surface feature is used here because it is very common in the field of surface characterization
and has more general meaning than surface defects. Surface features can thus refer to specific elements of the
surface that are to be characterized because they may be beneficial for a given application, whereas surface defects
refer only to features considered as undesirable.
Tab.1.2 summarizes the different factors that can influence the capability of the various in-
struments to properly characterize surfaces with intricate 3D macroscopic shapes and/or hidden
microscopic surface features.
Table 1.2 Factors influencing the ability of different instruments to properly characterize surface
with complex macroscopic shapes and hidden surface micro-features (e.g. notches hidden by
unmelted powder particles).
Contents
2.1 Materials and XCT data acquisition . . . . . . . . . . . . . . . . . . . . . 40
2.2 3D surface characterization methodology . . . . . . . . . . . . . . . . . . 41
2.2.1 Surface segmentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.2.2 3D roughness calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
2.2.3 3D curvature calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
2.2.4 Quantification of the harmfulness of surface notches . . . . . . . . . . . . . 54
2.3 Application to parts with complex geometries . . . . . . . . . . . . . . . 56
2.3.1 Gyroid structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
2.3.2 Octet-truss lattice structure . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
2.3.3 Impeller . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
2.4 Application to the prediction of crack initiation sites in fatigue . . . . 62
Intermediate summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
This chapter introduces a workflow for 3D surface characterization using XCT, mea-
suring both roughness and curvature. By combining these two measurements, we propose a
parameter representative of the severity of surface features with respect to mechani-
cal properties. This parameter is subsequently applied to the prediction of crack initiation sites
on chemically polished Ti64 samples produced by Electron beam Powder Bed Fusion (E-PBF)
and subsequently polished. Note that such a methodology could be applied to a large variety
of other applications not related to mechanics, e.g. to study the impact of surface roughness on
the osseointegration of biomedical implants [DIA 22, ALL 11].
| 39
Figure 2.1 | Ti64 architected structures fabricated by E-PBF. (a) Gyroid structure.
(b) Octet-truss lattice structure.
Two Ti64 architected structures were also characterized: a gyroid structure [SCH 70,
YAN 19, SOR 21] and an octet-truss lattice structure, see Fig.2.1a-b. Surfaces were kept
as-built in both cases.
1
The data were retrieved from the work of [PER 18], who performed this downscaling to reduce data size.
Table 2.1 | Acquisition parameters for XCT scans performed using the RX Solutions laboratory
tomograph.
Finally, a Ti64 impeller manufactured by L-PBF was scanned at the BM18 beamline at ESRF
synchrotron2 , using a voxel size of 18.5 µm.
All XCT scans were converted to 8-bit after reconstruction to reduce data size. After con-
version, volumes size were 158 Mo, 4.6 Go, 5.1 Go and 8.9 Go for the cylinder, the gyroid (10 µm
scan), the octet-truss lattice and the impeller respectively.
Dark Bright
peak peak
Maximum Maximum
abscissa abscissa
Black White
1
Figure 2.2 | Definition of the TTBH, illustrated based on the E-PBF cylinder normalized
grayscale histogram.
However, it may not always be the best choice if one is looking for some specific surface
features such as surface defects that will be prone to initiate failure during mechanical loading.
In this specific case, partially melted particles are unlikely to be of much interest. They can even
hide more severe surface defects such as notches. Contrariwise, deep and sharp notches are very
often the most critical defects and are therefore of great interest. However, the sharpest ones are
often difficult to segment from XCT data, because the corresponding voxels show intermediate
grayscale values. This leads Otsu’s method to consider many of them as foreground voxels,
erasing notches from the surface although they are the most interesting defects when questions
related to crack initiation and failure must be tackled.
This effect can be seen in Fig.2.3a and Fig.2.3b. Fig.2.3a shows an XCT radial slice of a
2 mm as-built E-PBF cylinder, where two deep notches can be seen. Fig.2.3b shows in orange
the pixels considered as foreground using Otsu’s threshold for segmentation. It can be seen that
both notches are not properly accounted for when using Otsu’s threshold, although
they are clearly visible on the grayscale image in Fig.2.3a.
A threshold corresponding to a higher grayscale value than Otsu’s one will be certainly more
adapted to capture such severe surface defects. Indeed, both partially melted particles and sharp
notches typically show intermediate grayscale values because they are at the interface between
matter and air. Thus, selecting a threshold that stands at the end of the plateau − just at
the left border of the histogram’s bright peak − will enable to discard some parts of unmelted
powder particles while better capturing sharp notches.
Notches
Figure 2.3 | (a) XCT 8-bit radial slice showing two sharp notches and a powder particle (voxel
= 5 µm). (b) Otsu’s threshold application (orange area) (c) TTBH application (purple area).
The thresholding method proposed is inspired by the triangle threshold introduced by [ZAC 77].
It is thereafter referred to as the Triangle Threshold for Bimodal Histograms (TTBH). Its princi-
ple is described schematically in Fig.2.2, and ImageJ/Python implementations have been made
available on an online repository [STE 23].
First, if the histogram is too noisy, it may be useful to smooth it. In the present work, a
moving average of size 10 is applied. Second, the histogram is normalized so that both the bright
peak maximum and the distance between the two histogram peaks are equal to 1. Finally, the
desired threshold is simply the gray value which maximizes the distance d as defined in Fig.2.2.
As required, the obtained threshold is located just at the left edge of the bright peak. Fig.2.3b-c
show that the two sharp notches are better captured using the TTBH than by Otsu’s
method. The powder grain is cropped, which can be both an advantage and a drawback,
depending on which surface features one aims to characterize.
It may be worth mentioning that this thresholding method is particularly sensitive to noise
and artifacts (e.g. beam hardening). This can be at least partially compensated by the use of
the proper noise-reducing filter beforehand. It may also be relevant in some cases to make a
compromise between the TTBH and Otsu’s threshold. A simple and convenient way to do that
can be to calculate both and take an intermediate value.
Following segmentation, the volume undergoes a cleaning process to remove all internal pores
or tiny objects caused by measurement noise. This cleaned binary volume is subsequently
referred to as the sample mask, illustrated in Fig.2.4b. In our case, the sample surface is defined
as the surface mask depicted in Fig.2.4c. It is the binary mask whose foreground is composed
of all surface voxels. These surface voxels are the foreground voxels of the sample mask which
have at least one background voxel as a first neighbor. A connectivity of 1 is used to determine
neighbors.
Figure 2.4 | (a) Schematic example of a grayscale XCT slice. (b) Corresponding sample mask.
(c) Corresponding surface mask.
Sample
voxels
Sample
Background mask
Erosion
1 + Logical XOR
Convert to float/int
2 + 3D gaussian filter
Surface
Voxel voxels
intensity
Max
Smooth
Background mask
4 SEDT Binary
Signed volumes
distance
Max
Smooth
mask SEDT
Min
Float/int volumes
5 Apply surface mask to SEDT
Height
Max
Min
Roughness
+
Background
Micro-roughness
Height
Max
6 S-filter
7 Remove borders if presence of edge effects Min
Roughness
Figure 2.5 | Workflow proposed for the 3D roughness computation. The sample mask
and surface mask are defined in Fig.2.4b-c. The images shown correspond to transverse cross-
sections of a 2 mm E-PBF Ti64 cylinder at different steps of the calculation.
The most conventional roughness parameters can be computed easily from the obtained 3D
roughness measurement. Formulas that have been used to compute roughness parameters in
this chapter and subsequent ones are given in Eq.2.2.
1 ÿ
Sa “ |h| (2.2a)
n surface voxels
d
1 ÿ
Sq “ h2 (2.2b)
n surface voxels
´ ¯ ´ ¯
Sz “ max h ´ min h (2.2c)
surface voxels surface voxels
Sv “ ´ min h (2.2d)
surface voxels
m3
Ssk “ 3{2
(2.2e)
m2
m4
Sku “ ´3 (2.2f)
m22
where
1 ÿ
mi “ ph ´ h̄qi (2.2g)
n surface voxels
is the biased ith central moment, n is the number of surface voxels, h is the height of a given
surface voxel and h̄ is the mean height over all surface voxels.
b 3D
a 2D
Osculating
circle
Figure 2.7 | (a) Definition of the curvature κ and the radius of curvature ρ for a curve in 2D.
(b) Schematic representation of a saddle-like surface. The normal ⃗n and the principal directions
are indicated at the saddle point. In this particular case, κmin ă 0 and κmax ą 0.
Fig.2.7b shows an example of a saddle-type surface, where d⃗min and d⃗max are respectively
displayed in purple and orange.
Knowing principal directions and curvatures, the directional curvature in any direction can
be derived using Eq.2.3 [TAU 95].
¨ ˛J ¨ ˛ ¨ ˛
vn 0 0 0 vn
κ p⃗v q “ ˝ vmin ‚ ¨ ˝0 κmin 0 ‚¨ ˝ vmin ‚
(2.3)
vmax 0 0 κmax vmax
2 2
“ vmin ¨ κmin ` vmax ¨ κmax
where κ p⃗v q is the directional curvature in the direction ⃗v , ⃗v “ vn ¨ ⃗n ` vmin ¨ d⃗min ` vmax ¨ d⃗max
is an arbitrary vector and ⃗n “ d⃗min ˆ d⃗max is the surface normal.
Finally, two other common curvatures are often used: the mean curvature κmean and the
Gaussian curvature κgauss (see definitions in Eq.2.4 and Eq.2.5):
κmin ` κmax
κmean “ (2.4)
2
Point where
Convolution curvature
kernel (rcurv=2) is measured
Foreground
pixel
Background
pixel
For example, κmean is directly related to the volume of the intersection between the convo-
lution kernel and the object. Hence, an estimation of κmean at a given boundary point can be
computed using Eq.2.6 [COE 14].
˜ ¸
16 1 Vinter prcurv q
κmean prcurv q “ ¨ ´ 4 3
(2.6)
3 ¨ rcurv 2 3 ¨ π ¨ rcurv
where rcurv is the radius of the spherical convolution kernel and Vinter prcurv q is the portion
of the convolution kernel’s volume that resides inside the surface’s boundary. By computing the
covariance matrix of the intersection instead of its volume, it is possible to design estimators
for the complete curvature tensor and thus estimate likewise principal curvatures and directions
[LAC 17].
Since computations are made on volumes, this offers the opportunity to perform compu-
tations using software such as ImageJ. Using Eq.2.6, κmean can for instance be computed using
a linear convolution. Since an optimized implementation of this method is already available in
the open-source C++ library DGtal, it has been used in the present work [DGt 22].
´Three
¯ curvatures are computed on this artificial object, namely κmin , κmean and κσ . κσ “
κ dσ is the directional curvature along the direction d⃗σ (vertical in this case). The comparison
⃗
of the three curvatures in Fig.2.9c-e makes it possible to identify which one highlights best the
different notches.
The first one is the minimum curvature κmin , which successfully captures all notches.
Furthermore, all have the same κmin value. The second curvature is κmean . Cups (type A) have
the lowest κmean value whereas linear notches (C) show intermediate values. This difference is
not necessarily desirable, since cups are not expected to reduce the mechanical properties more
than linear notches. Even worse, κmean tends to 0 for the saddle-like notch (B). Although this
notch was chosen as an example for the sake of clarity and seems far from a real case, there are
many regions of the surface where κmin ă 0 and κmax ă 0. Thus, κmean seems less relevant
than κmin to detect notches in general because it is somehow biased by the κmax contribution.
Finally, the third curvature is the directional one, κσ . As illustrated in Fig.2.9e, all notches
are well identified except the one parallel to d⃗σ . High κσ values are also more concentrated at
notches roots in comparison with κmin . κσ can thus be considered the most appropriate
choice when one aims at characterizing surface features with a specific orientation
with respect to a loading direction.
Based on these considerations, both κmin and κσ were thus computed on the same cylindrical
sample used in Section 2.2.2, see Fig.2.10a-b. Since the sample is a fatigue specimen meant to be
loaded along its axis, κσ was computed along this direction. A radius of 30 µm (= 6 voxel size)
was chosen here for the convolution kernel. An additional filter was used similarly to what has
been done for roughness, using λS “ 0.025 mm “ 5 voxel size. These choices were made to
provide a sufficiently detailed curvature measurement without being too much affected by noise.
Fig.2.10a shows that κmin underlines the presence of notches, which correspond to the lowest
values. κσ greatly attenuates the vertical notches, which can be seen when comparing enlarged
views in Fig.2.10a and Fig.2.10b.
In the previous sections, d is estimated by the local height, which corresponds to roughness,
while ρ was defined by the inverse of the curvature κσ . Some approximations still have to be
made to apply Eq.2.7 using those parameters.
First, it is important to emphasize that both the roughness and curvature, as calculated
in Section 2.2.2 and 2.2.3, are estimates. Therefore, their values may depend on the method
σpeak
d
ρ
σpeak
σ
Figure 2.11 | Elliptical notch in a semi-infinite panel, submitted to a tensile stress
σ in the direction d̃σ perpendicular to the ellipse major axis. d and ρ are respectively
the depth of the notch and the radius of curvature at its root. This notch generates a stress
concentration given by Eq.2.7, which means that the local stress at its root σpeak is higher than
the nominal stress in the section.
and parameters used for their measurement. The curvature values, for instance, are notably
influenced by changes in the convolution radius rcurv and in the voxel size. While there are
theoretical proofs, such as the one presented by [COE 14], showing that integral invariant-based
calculations approach exact curvature values as voxel size reduces, the resolution required to
observe this might be very high. Keeping this in mind, one should treat these estimated values
as semi-quantitative ones, that can for instance be used for ranking the severity of notches.
Another approximation to consider is that Eq.2.7 is valid at the notch tip, which might be
hard to detect automatically. As a result, the formula has been generalized and applied to every
point on the surface where both κσ ă 0 and d ă 0. This means it is used across the entire
notches, not just at their root. While Eq.2.7 might not be applicable for numerous points, the
highest values will still be found at the notch roots where the depth is maximal and the curvature
is minimal. Therefore, the values derived are still relevant, especially when identifying areas with
the highest stress concentration. In other words, it provides a semi-quantitative parameter
that reflects the mechanical severity of surface notches. Finally, an approximate value
of Kt , called Kt˚ , can be computed at the sample surface using Eq.2.8.
a
Kt˚ “ 1 ` 2 height ¨ κσ where height ă 0, κσ ă 0 (2.8)
where Kt˚ is the estimated local stress concentration, height is the local surface height
obtained from roughness measurements and κσ is the directional curvature along the loading
direction.
Fig.2.12a-b show an example of a Kt˚ map computed from the same cylindrical sample used
in the previous sections. Roughness and curvature (κsigma ) values used for computations are
the ones presented in Section 2.2.2 and 2.2.3.
Different views of the surface can be provided. The first one in Fig.2.12a is the usual external
view, which corresponds to what can be seen using a conventional 2.5D surface characterization.
The area shown is the same as in Fig.2.6 and Fig.2.10a-b. The two other views shown in
Fig.2.12b, are internal views which can be generated after extracting the surface from the XCT
scan. They offer a unique way to identify deep and sharp notches that would very often be
hidden using conventional characterization methods, see the comparison between Fig.2.12a and
Fig.2.12b. Once again, this illustrates the interest in characterizing the surface as a 3D free-form
one obtained by XCT.
Fig.2.13a shows a picture of the gyroid. Two local tomography scans [STO 08] were acquired
at the center of the structure. Large artifacts were observed on the XCT scans, which made the
use of the presented threshold for bimodal histograms inappropriate. Otsu’s threshold was found
to be more efficient in this case. To examine the influence of voxel size on roughness, curvature,
Table 2.2 Roughness parameters measured from the gyroid XCT scans with two voxel sizes
using λc “ 0.8 mm and λS “ 0.05 mm. To ensure consistency between the values measured for
the two voxel sizes, the parameters were evaluated using only the roughness values present on
the surface available in both scans.
Fig.2.13c shows the 3D κmin map on the same surface as Fig.2.13b. Fig.2.13d illustrates
how curvature evolves with respect to the convolution radius rcurv . The mean curvature is used
here since it is known that it should equal zero for an ideal gyroid, i.e. with no roughness.
The average value across the entire surface is represented by the average of absolute values
mean|κmean |, analogous to the role Sa “ mean|height| plays in roughness metrics. As observed
in Fig.2.13d, mean|κmean | tends to zero with increasing rcurv values. This is because a larger
rcurv measures curvature at a larger scale, corresponding more to the gyroid shape (with zero
mean curvature) than microscopic surface features, which have pronounced curvatures.
Finally, Fig.2.13e-f show the Kt˚ measurements derived from roughness and curvature. Note
that since the gyroid could be mechanically loaded in any direction, the minimum curvature is
used instead of κσ to compute Kt˚ . Fig.2.13e shows the 3D Kt˚ map obtained from the scan
performed with a 10 µm voxel size. Meanwhile, the cumulative distribution functions of the Kt˚
parameter for both voxel sizes are given in Fig.2.13f. The obtained Kt˚ values are slightly higher
for the 5 µm voxel size scan. This can, once again, be attributed to the better ability at high
resolution to properly capture sharp and deep notches.
2.3.3 Impeller
Finally, the methodology has been applied to an impeller manufactured by L-PBF to show
a real-case example. To keep a reasonably low voxel size (18.5 µm), only about one-half of the
impeller has been scanned. Fig.2.15a-b and Fig.2.16a-b show the 3D characterization of the
component from two points of view. Roughness was computed using λc “ 0.25 mm and λS “
0.09 mm. The relatively low λc value allows for minimizing the occurrence of biased roughness
values at sharp features such as corners. The curvature is measured using rcurv “ 100 µm, this
value being larger than in the previous cases because of the larger voxel size.
Fig.2.15a-b show well the complex internal structure of the component. Magnifying windows
make it possible to see that even with a 18.5 µm voxel size, the as-built surface roughness can
still be distinguished.
Fig.2.16a-b show the outside of the impeller. First, it is worth mentioning that although
most of the surface seems to appear in red in Fig.2.16a, this is only a matter of point of view.
Indeed, the small square window showing an internal view shows that valleys (in blue) are for
the most part hidden in the external view.
The magnifying windows show some specific features such as regularly spaced marks, charac-
teristic of surfaces that were machined after printing. A piece of scrap is also visible on the left
part of the window and may have been generated during the machining step. Because this scrap
is very thin, it leads to very high roughness values („450 µm). This shows the difficulty that
can arise when characterizing large real-case components: some unexpected features may
always lead to abnormal values, requiring a cleaning step to be conducted after calculations.
?
b
KImax « 0.65 ¨ σ ¨ π area (2.9)
where KImax is the maximum stress intensity factor along the front of the crack that would
emanate from the surface defect considered and σ is the uniform tensile load applied, in our
case, along the vertical d⃗σ direction).
Tab.2.3 and Tab.2.4 showcase respectively ranks and percentages scores from ten samples,
corresponding to samples 1-10 in [PER 20]. Note that in the second column giving the rank
?
obtained from the area parameter, the second number corresponds to the number of defects
that were pre-sectioned by [PER 20] for the manual measurement. Colors give information
about the quality of the prediction: a green cell indicates a good prediction and the color tends
to red as the quality of the prediction decreases.
From what has been said up to now, the most natural parameter in our study would be
Kt˚ (λc “ 0.8 mm | rcurv “ 30 µm). The choice of λc is dictated by the estimated size of
the relevant surface features (see Section 2.2.2) and rcurv is chosen as small as enabled by the
scan resolution. Although for most samples the killer defect has a rank inferior to 10, none is
Percentages
?
area κσ Kt˚
λc “ 0.8 mm
[PER 20] rcurv “ 100 µm
Sample rcurv “ 100 µm
1 48 84 71
2 86 97 89
3 74 100 82
4 78 94 92
5 61 98 89
6 53 89 91
7 85 94 89
8 97 100 97
9 100 100 100
10 61 66 83
Median 76 96 89
?
Table 2.4 | Killer defect predictions percentage scores for area, κσ , and Kt˚ pa-
rameters. Scores are computed for samples 1 to 10 from the study of [PER 20]. The median
percentages over all samples are also computed for each parameter.
4 10
2 5
50 100 150 200
rcurv (µm)
Figure 2.17 | Quality of initiation sites prediction by κσ , depending on the rcurv used
for curvature computation. The quality of predictions is quantified through the median and
the standard deviation of ranks obtained over all samples. The green vertical line corresponds
to the mean of maximum depths Sv measured for all samples with λ “ 0.8 mm.
Figure 2.18 | Some illustrations of the samples used for the killer defect prediction
study. Samples and XCT volumes used are the same as in [PER 20] (E-PBF Ti64 chemically
etched cylinders). (a) Fracture surface of sample 8, observed by SEM-SE. (b) Fracture surface
of sample 8, extracted from XCT volume. (c) Surface of sample 8 before fatigue testing, where
the killer defect is highlighted with a red arrow. (d-f) Curvatures and Kt˚ maps computed for
sample 8. (g-j) Curvature, Kt˚ and roughness maps computed for sample 1.
One can also notice that the percentage score for rcurv “ 100 µm is very high for most
samples, with a median percentage of 96%. In other words, it is observed that in most cases,
crack initiation occurs on a surface feature whose curvature κσ (rcurv “ 100 µm) is close to
the minimum (maximum in absolute value) reached on the whole surface. This highlights the
relevance of κσ (rcurv “ 100 µm) parameter, but as discussed above, it could also suggest that
most severe notches root tend to be smoothed, leading to similar value for many of them.
a b Shallow notch
Sv Sv
κdeepest = κshallow
Since rcurv “ 100 µm brings better results than rcurv “ 30 µm − when curvature is considered
alone, it seems interesting to see if the predictions obtained with Kt˚ are also better when
computing curvature at this scale. Tab.2.3 shows that it is not really the case, the results
obtained being similar to what was obtained with rcurv “ 30 µm. It can still be noted that
for two of the three "worst-case scenarios" (Samples 1, 4 and 10), the prediction is largely
improved by the the addition of roughness. This may come from the fact that in this case,
roughness still better takes into account the topology at a larger scale than curvature, thus
providing relevant complementary information in some cases. Furthermore, it is interesting to
Materials characterization
Contents
3.1 Samples manufacturing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.1.1 Processing conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.1.2 Fatigue and tensile specimens . . . . . . . . . . . . . . . . . . . . . . . . . . 73
3.2 Microstructure characterization . . . . . . . . . . . . . . . . . . . . . . . . 76
3.2.1 Microstructure of the bulk material . . . . . . . . . . . . . . . . . . . . . . . 76
3.2.2 Microstructure near the surface . . . . . . . . . . . . . . . . . . . . . . . . . 78
3.3 Tensile properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
3.4 Defects characterization using XCT . . . . . . . . . . . . . . . . . . . . . 84
3.4.1 Acquisition setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.4.2 Porosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
3.4.3 Surface defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
Intermediate summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
In Chapter 2, various materials were discussed through multiple case studies. However, the
next chapters of this manuscript will focus on a single material, namely Ti64 manufactured by
L-PBF. The upcoming sections will concentrate on some of its main characteristics, namely its
manufacturing process, microstructure, tensile properties, and defects.
It is worth noting that the manufacturing conditions for the samples match those used in
Quentin Gaillard’s work [GAI 23a]. Gaillard’s research mainly investigated microstructural and
static properties after different heat treatments. To gain a comprehensive understanding of these
aspects, readers are therefore encouraged to refer to Gaillard’s work, which serves as a valuable
resource for this chapter.
| 71
The chemical composition of the powder can be found in Tab.3.2. It is worth highlighting
that only fresh powder was employed. This decision was made to minimize variations in oxygen
contamination, defect formation and mechanical properties, as these can arise from the recycling
of powder [MAH 22].
Ti Al V Fe O C N H
Base 6.22 3.87 0.18 0.14 0.009 0.02 0.004
The powder bed was deposited using a roll. Although it enables obtaining a compact powder
bed, it requires some caution for the manufacturing of thin elements, as freshly printed thin layers
might be swept away by the roll. Thus, the printing of vertical specimens (presented thereafter)
required the use of "sarcophagus-like" support structures that consolidated the powder bed and
prevented specimens from bending towards the rolling direction. See Fig.3.4 and Fig.3.5 for
examples of such support structures.
Within the build chamber, a controlled atmosphere was maintained, with a laminar flow of
argon (Ar). The build plate did not undergo any pre-heating during the printing process.
Other processing parameters such as the laser power and scanning speed cannot be given for
confidentiality reasons. However, the scanning strategy is illustrated schematically in Fig.3.1,
as it is of primary importance to understand subsequent descriptions of microstructural features
and defects.
Each layer underwent scanning in two stages. First, the bulk material was melted using a
cross-hatching approach with a 90° angle between each layer. Subsequently, two contour scans
were implemented. The first contour scan is positioned on the outermost part of the sample.
Following the manufacturing process, the build plates were subjected to cleaning within an
ultrasonic bath containing water. This step is required to eliminate loose powder particles. The
build plates (plate + samples) were not sandblasted, primarily to prevent changes in surface
roughness and to avoid obscuring surface defects like valleys.
A stress-relief heat treatment of 2 h at 720 °C was then applied. XRD measurements
done by Quentin Gaillard [GAI 23a] showed that for the material considered, this heat treatment
achieves a complete relaxation of residual stresses. Here, the heat treatment was applied on the
entire build plate to prevent potential deformations that might arise if samples were detached
before undergoing stress relief.
90° rotation
Layer n
Caption
Figure 3.1 | Scanning strategy adopted for sample manufacturing. Bulk material is
first melted using a cross-hatching strategy, followed by two contour scans. A 90° rotation is
done between each layer.
Subsequently, specimens were detached from the build plate, and the support struc-
tures were manually removed. A final ultrasonic bath was used to clean the samples, removing
residual powder particles trapped between the specimens and support structures. At this stage,
the samples were prepared for subsequent characterization.
Figure 3.2 | Drawing of the flat tensile specimens. The dimensions are given in mm. The
specimen width equals 3.00 ˘ 0.02mm.
Samples manufacturing | 73
9
R2
43
9
R20
70
Figure 3.3 | Geometry of the cylindrical fatigue specimens. Dimensions are given in mm.
Tensile specimens were printed in three orientations: vertical ((along the Z-axis), at a
45° angle with respect to the build direction, and horizontal (XZ). Pictures of the printed
samples with their support structures can be seen in Fig.3.4. The vertical specimens employed
"sarcophagus-like" support structures, while more standard supports were used for the 45° spec-
imens. Although not visible in Fig.3.4, the only points of contact between the supports and
specimens are located on the specimen heads. The middle portion of the support structures ap-
proaches the specimen without direct contact. Additionally, small pillars were used as supports
within the gauge length of the XZ specimens. However, after being manually removed, these
pillars left noticeable surface irregularities that are expected to adversely impact the mechanical
properties, please refer to Section 3.3 for further details.
For the fatigue specimens, two build orientations were investigated: vertical (Z) and
45°. Images of both types of specimens are shown in Fig.3.5. Similar support structures as
those used for the tensile specimens were employed. The primary distinction is that the 45°
specimens are not fully supported within the gauge length. Similar to the tensile specimens, the
only points of contact between these supports and the specimens are located on the specimen
heads.
Figure 3.5 | Pictures of printed fatigue specimens and their support structures.
Specimens were printed in the (a) vertical (Z) and (b) 45° directions. Pictures courtesy of
Quentin Gaillard.
Samples manufacturing | 75
Microstructure characterization | 77
Fig.3.6a and Fig.3.7a provide insight into the morphology and crystallographic orientation
of α laths. Due to crystallographic constraints dictated by the Burger orientation relationships,
these laths can have only 12 orientations relative to a given prior β0 grain orientation. This
leads to a less pronounced texture compared to the prior β0 phase. EBSD data indicate
an average thickness of α-laths of 0.5 µm, correlating with a relatively high hardness of
394 ˘ 4 HV1 kg .
Given that the microstructure of the bulk material has been previously discussed, the sub-
sequent focus will be solely on the distinctive features of the border and contour zones. In this
context, EBSD maps were generated near the surface within vertical (Z) and horizontal (XY)
planes − refer to Fig.3.9a-j and Fig.3.10a-j respectively. Similar to the bulk microstructure,
these maps characterize both α and prior β0 phases.
The border region (spanning from 150 µm to 600 to 800 µm beneath the surface) exhibits
a microstructure similar to that of the bulk material. Nevertheless, the prior β0 grains in
this zone appear to be twice the size of those in the bulk material − see Fig.3.8b-c and
Fig.3.10f. An explanation has been proposed by [GAI 23a] to rationalize this phenomenon. It
hypothesizes that at the beginning of a laser track, the preceding adjacent laser track might not
have had sufficient time to fully solidify near the surface. Consequently, in such instances, β0
grains might solidify from a melt pool twice the usual size, accounting for the observation of
larger β0 grains.
The contour area shows a notably distinct microstructure. In the immediate sub-
surface region (approximately within 0 to 50 µm), small prior β0 grains are apparent, elongated
in a direction close to Z. The slight deviation from the perfect alignment along the vertical axis
can be attributed to local variations in thermal gradients near the surface. Beyond 50 µm, prior
β0 grains exhibit elongation both towards Z and perpendicular to the surface.
Across all these regions and regardless of the prior β0 grain morphology and texture, the mi-
crostructure following printing and heat treatment consists of fine α laths, with minor texturing.
Microstructure characterization | 79
Microstructure characterization | 81
1200
Engineering stress (MPa)
1000
YS (MPa) UTS (MPa) εf (%)
800 Z 1098 1147 7.8
400
Vertical (Z)
45°
200
Horizontal (XZ)
0
0 2 4 6 8 10
Engineering deformation (%)
Figure 3.11 | Tensile curves measured for the three build directions. Values for YS,
UTS and ϵf are presented in the form (mean ˘ std). Standard deviation is not measured for Z
specimens, since two values are not enough to get a meaningful estimation.
The measured YS and UTS values exhibit strong reproducibility and consistency with re-
sults from the literature [NGU 22, GRE 17, KAS 15, BEN 18, VAN 12, CAI 15]. The slightly
lower strength of the horizontal (XZ) specimens can likely be attributed to the significant pro-
tuberances that remain after support structure removal. These defects substantially increase
the XCT-measured section of the specimen, though they are not expected to contribute to the
strength of the material.
In contrast to tensile strength, elongation to failure demonstrates substantial variation, de-
pending on the build orientation. Specifically, XZ specimens display very little defor-
mation before failure (ϵ¯f “ 0.8%). Once again, this could partially stem from support
structures within the gauge length, which leave pronounced surface defects. However, findings
from [GAI 22] suggest that an additional factor contributes to this lack of ductility. This study
reveals that even for tensile specimens with the bottom surface (affected by support structures)
Figure 3.12 | SEM-SE observations of tensile fracture surfaces for (a) an XZ and (b)
a Z specimen. On the XZ specimen, delamination can be seen at the interface of prior β0 grain
boundaries, leaving linear patterns on the fracture surface (see orange arrows).
Tensile properties | 83
Table 3.3 | Number of specimens scanned by XCT, voxel size and characterized length for
tensile and fatigue specimens.
Following reconstruction, the scans were converted into 8-bit unsigned volumes and merged
using an automated Python script developed in-house. To reduce noise, a 2 ˆ 2 ˆ 2 median filter
was subsequently applied to the resulting volume. This volume was then directly employed for
defect characterization, as detailed in the next sections.
As discussed in Chapter 1, defects can be classified into two primary groups: internal defects
(primarily pores) and surface defects. Due to their distinct shapes, origins, and methods to
characterize them, the two categories are examined separately in dedicated sub-sections. The
focus of this characterization, in both cases, is confined to the L-PBF Ti64 fatigue specimens.
The analysis was made using 29 Z specimens and 27 45° specimens.
Sphericity
0.52 0.61 0.7 0.8 0.9
Figure 3.13 | Scale showing the evolution of pore morphology with increasing sphericity. Pores
surfaces were extracted using the marching cubes algorithm.
a b
Z (29)
Total number of pores 35
Pore density (#pore / µm)
(#pore / µm / specimen)
in specimen: 100 45 (27)
30
Pore density
≈ Number of pores in the material
25
whose equivalent diameter Deq is
within a window of 1 µm around 20 µm
20
0
0 10 20 30 40 50 60 70
0 20 Deq (µm) Deq (µm)
c d
2000 Z (29) Z (29)
(#pore / µm / specimen)
1500
Pore density
Pore density
3
1250
1000
2
750
500
1
250
0 0
0.4 0.5 0.6 0.7 0.8 0.9 1.0 0 250 500 750 1000 1250 1500 1750
Sphericity Distance to surface (µm)
Figure 3.14 | Distributions of pores equivalent diameter Deq , sphericity and distance
to surface. (a) Schematic example of pore distribution. Distributions are here computed so
that their integral equals the number of pores in the specimen. (b) Deq , (c) Sphericity and
(d) Distance to surface distributions. For each graph, lines correspond to mean distributions
obtained for all samples while shaded areas represent an interval of one standard deviation
from mean lines. Mean distributions are computed separately for vertically built (Z) and 45°
specimens. The number of characterized specimens for each build orientation is specified in the
graph legend.
To accomplish this, normalized probability distributions were first computed individually for
each specimen through kernel density estimation (KDE) using a Gaussian kernel. The bandwidth
was determined following Silverman’s rule of thumb [SIL 98, p.45]. To mitigate edge effects, the
obtained distributions were cropped based on the minimum and maximum values found in the
data. Following cropping, they were normalized to ensure an integral of 1.
Fig.3.14b demonstrates that for both build orientations, all pores exhibit equivalent
diameters smaller than 70 µm. The majority are notably smaller − 81% and 88% being less
than 30 µm for the Z and 45° specimens, respectively. This graph also indicates that more pores
are measured in 45° specimens, with an average increase of 29%.
In addition to their relatively compact size, pores tend to have high sphericity values,
as illustrated in Fig.3.14c. This tendency is more pronounced for Z specimens. Combining size
and sphericity data suggests that most pores are of the keyhole type, as opposed to lack of
fusion defects. Lack-of-fusion defects typically possess larger dimensions and irregular shapes,
resulting in lower sphericity values.
Fig.3.14d offers insightful information about pore distribution within specimens. The ma-
jority is concentrated in a subsurface layer, with 62% and 71% being situated within
200 µm from the surface for Z and 45° specimens, respectively. The main peak in both Z and
45° distributions indicates that pores tend to form at a specific depth from the surface, ap-
proximately 140 µm. This aligns with the overlap region between contour laser tracks and bulk
material. One hypothesis to explain this could be that inappropriate laser contour scanning
parameters or overlap adjustment might be responsible for these pores. Excessive overlap could
for instance lead to locally high energy density, encouraging keyhole instabilities and pore for-
mation. It is worth noting that even considering this over-concentration of sub-surface porosity,
the specimens still show a high density (ą 99.9%).
An additional peak, close to 0 µm, is observed in the 45° distribution in Fig.3.14d. This
corresponds to pores detected in the down-skin region, just beneath the surface. XCT scan
resolution does not allow precise identification of whether these are actual internal pores. Most
are likely open pores or, more specifically, roots of thin surface valleys not accurately captured
during volume segmentation. Manual inspection of XCT scans supports this assumption for
the majority of these pores, although resolution is insufficient for some of them. Some may be
internal pores resulting from the closure of surface valleys. Notably, these pores tend to exhibit
an elongated morphology, which partly explains the higher number of pores with intermediate
sphericity values in the 45° specimens (see Fig.3.14c).
To summarize, the vast majority of pores found in 45° specimens are similar to those observed
in Z specimens in terms of size, shape, and location: most are small (ă 30 µm) and roughly
spherical pores located at the interface between contour and bulk laser scanning strategies
(« 140 µm deep).
Figure 3.15 | Examples of surface defects on a Z specimen. Enlarged views show SEM-
SE images of different areas along the specimen.
Figure 3.17 | Surface of a Z specimen before and after several cleaning steps. The
same region is shown in three images, and roughness measurements were made using the surface
extracted from the whole XCT scan. Images are curvature 3D representations obtained from
XCT scans. The surface is segmented using Otsu’s threshold to better visualize powder particles.
Curvature and roughness were computed using the following parameters: λc “ 0.8 mm, rcurv “
20 µm and no S-filter.
The results indicate that US cleaning enables the removal of a large part of powder particles.
However, a significant quantity of powder remains and can partially be removed by cleaning
with a plastic brush. This means that (for the experimental conditions investigated) even after
US cleaning, a substantial number of powder particles remain, although they are not firmly
attached to the surface. This can be a concern, as mentioned earlier.
From a research standpoint, this highlights the great importance of thorough surface clean-
ing. As depicted in Fig.3.18a-c, the valleys remain often concealed and hidden by particles
before cleaning. Given that the primary importance of comprehensive surface cleaning (includ-
ing brush-based methods) was realized only during the mid-phase of this PhD research, some of
the characterized specimens underwent US cleaning alone, while others also underwent manual
cleaning using a plastic brush. This can potentially contribute to dispersion in roughness mea-
surements, which show some variations depending on the cleaning step − see Fig.3.17a-c and
Fig.3.18a-c.
Spatters
Large particles can occasionally be observed (see e.g. Fig.3.16). Here, "large" refers to sizes
that significantly exceed the expected maximum powder particle size of 25 µm. Although no
in-depth study has been carried out concerning their origin, they are considered to be most
likely spatters. For the sake of simplicity, they will therefore be referred to as such hereafter.
However, some may also originate from the initial powder itself.
Large protrusions
Large protrusions are frequently observed, particularly on vertical walls. In these areas, they
exhibit a distinctive morphology that is not commonly reported in the literature. Fig.3.19 illus-
trates some of these protrusions on a 3D representation of a Z specimen obtained via XCT. The
representation combines roughness and curvature data to make visualization easier. Notably,
protrusions are identified by their height, depicted in red in Fig.3.19.
These protrusions are generally quite large (around 300 to 400 µm wide) and have a curved
water-drop shape. Furthermore, their spatial distribution appears non-random, although it can
vary significantly between specimens. For instance, Fig.3.19 shows that several protrusions are
vertically aligned. They also tend to be denser on one side of the specimen, while the other
side presents a more conventional plate-pile appearance. The most plausible explanation is that
these protrusions correspond to the initiation or termination points of the laser tracks, where
variations in energy density or melt-pool dynamics could contribute to the formation of such
defects.
Dross
Dross is widely visible in down-skin regions − refer to Section 1.2.2 for dross definition.
This leads to the presence of significant protrusions and valleys in between, see Fig.3.20. The
protrusions vary in shape and size, but are frequently observed to be a few hundred µm wide.
Valleys
Valleys can be seen everywhere on the surface but with different morphologies depending on
their location.
a b c
Notch root
smoothed due to
limited resolution
BD
Down Up
skin skin
45°
Specimen
axis
Figure 3.23 | Definition of up-skin and down-skin regions used for roughness parameters
calculation on 45° specimens.
Results were summarized in the form of violin plots, which are enhanced versions of box
plots. The way those plots can be read is detailed in Fig.3.24.
25
Probability density
Unlikely values
distribution
Q3
20
Adjacent values
Sa (µm)
Q1
Values within 1.5 IQR
of the nearest quartile
15 Defines limits above which points
are considered to be outliers Inter-quartile range (IQR)
Space between 1st and 3rd quartiles
5
Vertical (Z) 45° / Down-skin 45° / Up-skin
Figure 3.24 | Presentation of violin plots for roughness parameters.
20 20
Sa (µm)
Sa (µm)
15 15
10 10
5 5
0 0
Vertical (Z) 45° / Down-skin 45° / Up-skin Vertical (Z) 45° / Down-skin 45° / Up-skin
Figure 3.25 | Violin plots for Sa parameter computed for (a) λc “ 0.25 mm and (b)
0.8 mm. In both cases, an S-filter was applied using λS “ 0.03 mm “ 10 voxel size.
Although tendencies between build orientations and down-skin/up-skin areas are the same
for both cut-off wavelengths, absolute values are significantly different. For instance, the median
Sa value in down-skin area of 45° specimens is around 12.2 µm for λc “ 0.25 mm whereas it is
only about 18.8 µm for λc “ 0.8 mm. One should therefore always be careful when comparing
roughness parameter values computed with different workflows, in particular, different cut-
off wavelengths. In the present case, all characterizations done hereafter use the same cut-
off wavelength to overcome this issue. In all cases, an S-filter is also applied after roughness
computation, using a cut-off λS “ 0.03 mm.
Violin plots for Sa , Sq , Sz , Sv , Ssk and Sku parameters are shown in Fig.3.26a-f. Mean values
for these parameters are summarized in Tab.3.4.
Table 3.4 | Roughness parameters, computed for Z specimens (29 sp.) as well as
down-skin and up-skin regions of 45° specimens (27 sp.). Roughness was computed
from XCT data using the 3D characterization workflow introduced in Chapter 2. The cut-off
values for filters were set to λc “ 0.8 mm and λS “ 0.03 mm. Values presented in the format
(mean ˘ std).
30
20
25
Sq (µm)
Sa (µm)
15 20
15
10
10
5
Vertical (Z) 45° / Down-skin 45° / Up-skin Vertical (Z) 45° / Down-skin 45° / Up-skin
400
c Sz 180 d
Sv
350 160
300 140
120
Sv (µm)
Sz (µm)
250
100
200
80
150 60
100 40
Vertical (Z) 45° / Down-skin 45° / Up-skin Vertical (Z) 45° / Down-skin 45° / Up-skin
2.0
e Ssk 16 f
Sku
1.5 14
12
1.0
10
Sku
Ssk
0.5 8
6
0.0
4
0.5 2
0
Vertical (Z) 45° / Down-skin 45° / Up-skin Vertical (Z) 45° / Down-skin 45° / Up-skin
Figure 3.26 | Violin plots for roughness parameters (a) Sa , (b) Sq , (c) Sz , (d) Sv (e) Ssk
and (f) Sku . Roughness is computed using λc “ 0.8 mm and λS “ 0.03 mm. Violin plots were
computed using 29 Z and 27 45° specimens.
The material exhibited a microstructure and tensile properties consistent with the existing
literature. The microstructure consisted of fine α laths (averaging around 0.5 µm thickness)
along with a minor fraction of β phase (2.2 ˘ 0.3%). Prior β0 grains were reconstructed
from EBSD data and showed a typical columnar morphology with the growth direction
parallel to the build direction. They showed a significantly different morphology near the
surface, which can probably be explained by differences in local energy density and thermal
gradients during printing. Everywhere, though, the final microstructure is primarily
composed of α laths and exhibits only little texturing. For more detailed information
regarding the material microstructure, readers are encouraged to refer to [GAI 23a] and
[GAI 23b].
Compared to conventional wrought Ti64 [NGU 22], the tensile strength was relatively
high (Y¯S “ 1098 MPa for Z specimens), but the elongation to failure was rather limited
(ϵ¯f “ 7.8% for Z specimens). This can be attributed partly to the aforementioned fine
microstructure. XZ specimens displayed notably low ductility (ϵ¯f “ 0.8%), possibly due to
the presence of significant defects left after support removal. Additionally, prior β0 grain
boundaries, aligned perpendicular to the loading direction for these specimens, seemed to
serve as weak points contributing to premature failure.
XCT-based characterization revealed very low porosity (< 0.002%) with predominantly
near-spherical and small-sized pores (over 80% smaller than 30 µm). However, the majority
of these pores were located beneath the surface (« 140 µm deep), likely originating from
contour scans.
Surface roughness was also evaluated via XCT, utilizing the 3D characterization approach
introduced in Chapter 2. The surface roughness was found to be relatively low in com-
parison to literature values, particularly for up-skin and vertical regions. For instance, the
mean Sa value across all Z specimens was 7.3 µm. In contrast, down-skin regions exhibited
higher roughness with a mean Sa of 19.3 µm. Among the observed surface defects with the
potential to influence fatigue properties, valleys were prominent. These valleys originated
from the plate-pile effect in up-skin and vertical surfaces, and from dross formation in down-
skin regions. These valleys could exhibit intricate morphologies with re-entrant features,
frequently concealed by unmelted powder particles. Conventional 2.5D surface characteri-
zation tools may encounter limitations in properly assessing such surfaces, making XCT a
suitable solution for this purpose.
Contents
4.1 Fatigue properties and crack initiation mechanisms for as-built specimens104
4.1.1 Protocol for fatigue tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.1.2 Results of fatigue tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.1.3 Killer defects identification: the need for a 3D characterization . . . . . . . 111
4.2 Alpha-case formation and surface embrittlement induced by heat treat-
ments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
4.2.1 Alpha-case formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
4.2.2 Impact on tensile properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
4.2.3 Impact on fatigue properties . . . . . . . . . . . . . . . . . . . . . . . . . . 123
4.3 Improvement of fatigue properties by Plasma electrolytic Polishing (PeP)131
4.3.1 Principle of PeP surface treatment . . . . . . . . . . . . . . . . . . . . . . . 131
4.3.2 Conditions of PeP treatments applied to fatigue specimens . . . . . . . . . 133
4.3.3 Roughness of specimens polished by PeP . . . . . . . . . . . . . . . . . . . 138
4.3.4 Fatigue resistance improvement and change in crack initiation mechanism . 146
Intermediate summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
| 103
Specimens were first tested with their as-built surfaces. A stress-relief heat treatment of
2 h at 720 °C was applied to these specimens, ensuring they were free of residual stresses. In
this condition, 22 Z and 18 45° specimens underwent testing. Among these, 10 Z and 10 45°
specimens were XCT scanned both before and after failure for in-depth analysis.
2
The stress levels were calculated based on a theoretical section of S “ π p3 mmq
4 « 7.07 mm2 .
Yet, minor geometrical deviations from the CAD definition meant that the effective specimen
gauge diameter was slightly reduced. Thus, the mean section of the gauge area for all XCT
scanned specimens was determined, and the average and standard deviation were calculated
for each build orientation. The obtained values are 6.91 ˘ 0.07 mm2 for Z specimens and
6.99 ˘ 0.32 mm2 for 45° specimens. They were used for correcting stress levels in Wöhler
curves. However, for the sake of simplicity, the same rounded theoretical stress levels − i.e.
those obtained using the theoretical section − will be used in the discussions for both Z and 45°
specimens.
In addition to as-built specimens, four machined and manually polished (M&P) vertical
specimens were tested as a reference. These specimens were initially machined from rectangular
bars through turning and then manually polished with SiC P1200 paper. Thus, they were free of
both the as-built surface roughness and the high concentration of subsurface pores highlighted
in Section 3.4.2. Following their preparation, a supplementary stress-relief heat treatment of
2 h at 660 °C was performed. This ensured the elimination of any residual stresses potentially
arising from machining and polishing while keeping the microstructure almost unchanged.
104 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
1100
R = 0.1
1000
900
800
σmax (MPa)
Vertical | as-built
700 45° | as-built
600 Vertical | M&P
Number of
2 4 3 superimposed points
500
400 32 4
300 2
2
200
104 105 106 107
Nf
Figure 4.1 | Wöhler curves of as-built specimens. M&P specimens are also represented
as a reference. They were submitted to a stress-relief treatment (2 h at 660 °C) after machining
and polishing. Fatigue tests are performed at R “ 0.1.
Fig.4.2a-b show fracture surfaces for as-built Z and 45° specimens. The relatively flat region
in the upper parts of Fig.4.2a-b show unambiguously that in both cases, cracks initiated from
the surface. More generally, it has been observed that crack initiation always occurred at
the surface for as-built specimens. It is worth noting that no single defect can be identified
in Fig.4.2a-b, and that crack initiation seems to have occurred in vast zones. This does not
necessarily mean that multiple cracks initiated at different locations since this usually leads to
the formation of more ridges than what can be seen in Fig.4.2a-b. The mechanism behind this
particular fracture surface morphology is discussed in Section 4.1.3.2.
Fig.4.1 shows that M&P specimens exhibit significantly higher fatigue strength
than as-built specimens. Despite the low number of tested specimens, fatigue strength at
106 cycles of M&P specimens can be estimated at about 850 MPa. By looking at M&P fracture
surfaces in Fig.4.3a-b, this huge difference can be explained by a change in the crack initiation
mechanism. Indeed, all crack initiations occurred on an internal pore in the case of M&P
specimens. The pore was deep inside the material (ă 290 µm below the surface) in 3 cases
over 4, whereas it was directly beneath the surface in one case. Fig.4.2a-b show two surface
crack initiations on as-built specimens and Fig.4.3a-b show two crack initiations on bulk pores
of different sizes in polished ones.
Fatigue properties and crack initiation mechanisms for as-built specimens | 105
These results confirm that in our case and as it was reported several times in the literature,
surface defects of as-built samples are much more critical than internal pores. Only
when surface defects are removed by machining/polishing, do pores become potential crack
initiation sites. Note that for as-built specimens, the high density of sub-surface pores described
in Section 3.4.2 seem to have negligible impact on the fatigue resistance compared to surface
defects.
Concerning the M&P specimens, it is also interesting to note that their fatigue strength
seems to be in the upper range of what can be found in the literature for L-PBF Ti64 (see again
Tab.1.1 in the state of the art for details). This can be primarily attributed to the good bulk
material integrity, which, as shown in Section 3.4.2, contains only a few pores of limited size
and rather high sphericity. This can be confirmed by the fracture surfaces in Fig.4.3a-b of both
polished specimens that were tested at 950 MPa. In those two cases, crack initiation was caused
by small rounded pores. This is especially true in Fig.4.3b, where the killer defect is a perfectly
spherical pore of only 11 µm.
106 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Figure 4.3 | SEM observations of two polished specimens fracture surfaces showing
crack initiation at small rounded pores. Large views are obtained in BSE topographical
mode, whereas higher resolution images are obtained in SE mode.
The good properties of polished specimens still need to be qualified because of the reduced
number of specimens and the large scatter observed. The two specimens tested at 950 MPa reach
for instance very different fatigue lives: one breaks at 21 441 cycles while the other lasts almost
100 times longer (1 714 472 cycles). This can be attributed to significant differences between
the defect populations of both specimens. Thus, the equivalent diameter can be estimated to
be 52 µm in Fig.4.3a, much larger than the 11 µm found in Fig.4.3b. Furthermore, the pore in
Fig.4.3a is much less spherical than the one in Fig.4.3b. A Hot Isostatic Pressing treatment
(HIP) could in this case be relevant to remove porosity and therefore achieve more reproducible
fatigue properties.
However, a HIP treatment would significantly affect the microstructure. This could have ad-
verse effects, since the fine microstructure obtained after low-temperature heat treatment (2 h at
720 °C) could also contribute to the good observed properties. However, it seems difficult to draw
a conclusion on this question since contradictory studies exist to explain which microstructure
should be the most efficient regarding fatigue properties [POL 14]. The lamellar microstructure
obtained with L-PBF could be rather advantageous regarding crack propagation, as cracks could
tend to deviate from their trajectory during propagation to follow α-laths interfaces [GAL 17].
All these bifurcations imply that more energy is needed to propagate cracks. This hypothesis
is coherent with the high density of micro-cracks found in the propagation zone of the fracture
surface and which most likely follow α-laths interfaces, see e.g. Fig.4.4. Some studies even state
that finer α-laths will result in better fatigue strength [LIU 19, LIU 17]. As a reminder, the
α-laths width in the present study is particularly small (« 0.5 µm, see [GAI 23b]). However,
other studies suggest conversely that better fatigue strengths can be reached with coarser laths
[GAL 17, WAN 23] and further work would thus be needed to clarify the matter.
Fatigue properties and crack initiation mechanisms for as-built specimens | 107
Regarding 45° specimens, it is interesting to note that all crack initiations occurred
in the down-skin region, as has already been found in the literature [PEG 18]. Again,
this is coherent with roughness measurements in Section 3.4.3.2 showing that the down-skin
region is much rougher than the up-skin one. This probably explains why, as can be noticed
in Fig.4.1, the fatigue life of the 45° specimens is on average slightly lower than that of the Z
specimens, especially at lower stresses. For instance, at 350 MPa, median cycles to failure for Z
and 45° specimens are 255 152 cycles and 97 800 cycles, respectively. Nevertheless, the distinction
between Z and 45° orientations is minimal compared to the larger difference between as-built
and M&P specimens.
108 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Looking in more detail into the position of initiation sites in 45° specimens, it can be noticed
that failure occurred most frequently in the lower section, closer to the build plate during
fabrication (see Fig.4.5a). This observation is not reflected in Z specimens. Fig.4.5b highlights
this trend, plotting the initiation sites position along the specimen axis in 10 Z and 10 45°
specimens. A position of 0 means that the initiation site is located in the middle of the gauge
length.
-4 -2 0 2 4 Position along
0 specimen axis (mm)
1
Build plate
b
Z (10)
0.4 45° (10)
Probability density
0.3
Gauge
length
0.2
0.1
0.0
4 3 2 1 0 1 2 3 4
Build plate
Position along specimen axis (mm)
Figure 4.5 | Initiation sites positions along specimen axis for Z and 45° specimens.
Each point corresponds to a specimen. 10 Z and 10 45° were used for this study. For each
specimen, the initiation site has been identified on the corresponding XCT scan following the
methodology described in Section 4.1.3. The position of the initiation site was then estimated.
A position of 0 means that the initiation site is located in the middle of the gauge length.
Probability density curves were computed by kernel density estimation.
Fatigue properties and crack initiation mechanisms for as-built specimens | 109
-4 -2 0 2 4 Position along
a 0 specimen axis (mm)
b c
Sa (µm) Sv (µm) Sa (µm) Sv (µm)
140 140
30 Z Sa 30 45° | downskin
Sv 120 Sa 120
Initiation site position
25 25 Sv
Initiation site position
100 100
20 20
80 80
15 15
60 60
10 10
40 40
4 2 0 2 4 4 2 0 2 4
Position along specimen axis (mm) Position along specimen axis (mm)
Figure 4.6 | Roughness profiles along specimens’ axis. (a) Scheme showing the area used
to measure roughness parameters at each position along a specimen axis. (b) Sa and Sv profiles
for Z specimens. (b) Sa and Sv profiles for down-skin regions of 45° specimens. In (b) and (c),
gray points indicate the position of initiation sites. Roughness parameters were measured using
only specimens where the initiation site has been identified (10 Z and 10 45° specimens). Curves
represent the average roughness profile (among specimens of the same build orientation) whereas
shaded areas correspond to a distance of one standard deviation from the average curves.
Refer to Fig.4.6b for roughness parameters profile visuals for Z specimens, and Fig.4.6c for
those of 45° specimens. Due to the consistent failure in the down-skin region of 45° specimens,
only points from this region were used in roughness computations. For a clearer definition of
this down-skin region, see Fig.3.23.
As can be seen in Fig.4.6b, Sa and Sv roughness parameters show only little variations along
the specimen axis of Z specimens. This is consistent with the uniform distribution of initiation
sites marked by gray spots.
110 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Conversely, Fig.4.6c reveals an upward trend in both Sa and Sv values along the axis in the
down-skin region of 45° specimens. This is probably due to the fact that because of the presence
of the fillet radius, the upper part of 45° specimens is nearly horizontal during manufacturing.
Conversely, lower parts are near vertical, which leads to smoother surfaces.
Surprisingly, initiation sites predominantly cluster in lower parts of 45° specimens,
i.e. where Sa and Sv values are lowest. This tends to indicate that surface roughness might
not be the sole determinant of initiation site position. Another influencing factor might for
instance be the presence of support structures under the upper sections of 45° specimens during
manufacturing (see Fig.3.5 for an image of support structures). Lower parts of the specimens
where failure occurs did not require the use of support structures and were therefore built
directly on the powder bed. Such factors might lead to a slightly different microstructure (due
to differences in thermal history) or minor geometric shifts (due to residual stresses), potentially
influencing fatigue crack initiation.
In any case, these results indicate that roughness parameters (or at least the sim-
ple Sa and Sv parameters) are not necessarily reliable indicators for the fatigue
performance of L-PBF as-built specimens.
Fatigue properties and crack initiation mechanisms for as-built specimens | 111
To go further in the initiation site identification, XCT was used to get a broader 3D point of
view. To this end, 10 Z and 10 45° fatigue specimens were scanned before and after failure. The
workflow for initiation site identification is described in Fig.4.8a-d. The specimen taken as an
example is the one that is shown in Fig.4.7b. This allows to check whether river lines emerging
from the surface away from the spatter could indicate crack initiation from another defect.
112 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Figure 4.8 | Workflow for crack initiation site identification using both SEM and
XCT scans (before and after failure). The specimen used as an example was built at 45°
and failed after 125 534 cycles at 300 MPa. (a) SEM-SE fracture surface observation. (b) Surface
mesh of the fracture surface obtained from XCT scan. (c) Crack initiation site observed on the
fracture surface from the side of the specimen. Gray levels correspond to κmin , estimated using
rcurv “ 20 µm. (d) Same area as in (c), trace backed on initial XCT scan before failure.
Fatigue properties and crack initiation mechanisms for as-built specimens | 113
Once the initiation site is identified on the SEM image, the same fracture surface is observed
in 3D using XCT. For this purpose, the surface is extracted from the XCT scan containing the
fracture surface. The mesh extracted from the XCT scan using the conventional marching cubes
algorithm is for instance shown in Fig.4.8b. By comparing Fig.4.8a and Fig.4.8b, the initiation
site is identified on the XCT fracture surface.
Finally, the initial XCT scan before the fatigue test is compared to the XCT fracture sur-
face to identify what surface feature caused crack initiation. This could be done using mesh
representations of surfaces before and after failure. But while mesh representations give a good
perception of volume, they might not always effectively render the most relevant surface fea-
tures. On the opposite, it has been shown in Section 2.2.3 that the minimum curvature greatly
highlights surface valleys, which, as will be shown, are the surface features of primary interest.
Therefore, for the killer defect identification surfaces were rather visualized by using a point
cloud of the surface colored using a minimum curvature measurement, see Fig.4.8c-d.
The spatter that was previously seen on the SEM observation can be seen in Fig.4.8b-d
obtained from XCT data as well. The crack path seems to follow a surface valley over a
long distance, despite the fact that this valley is not aligned on a horizontal plane.
This is highlighted by the magnifying windows of Fig.4.8c-d. It is precisely the same area that
was identified previously as a potential initiation site. Although such a valley can be identified
in the 3D representation (especially when they are underlined by the 3D rendering of curvature),
it cannot be identified using SEM only, see Fig.4.8a. This confirms the relevance of such a
3D characterization method for crack initiation site identification on L-PBF as-built
surfaces.
The precise alignment of the crack with valleys indicates their contribution to fatigue failure.
Parallel river lines in SEM images, rather than emerging solely from the spatter, suggest crack
propagation from the entire area, including the spatter and valleys. Two scenarios could be
considered:
1. Valleys and the spatter both served as crack initiation points, leading to simultaneous crack
initiation over a vast perimeter. Although the valleys may appear to be less pronounced
defects than the spatter, their extent could allow many cracks to initiate along the surface
before coalescing, causing rapid damage and short fatigue life, as reported by [PEG 18] in
L-PBF Ti64.
2. The crack primarily began at the spatter, but valleys accelerated subsequent crack prop-
agation due to the generated stress concentration.
114 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
To confirm these theories, advanced techniques, like in-situ crack propagation in synchrotron
XCT, would have been necessary. As of now, no definitive answer can be provided.
Nonetheless, the 3D initiation site characterization in Fig.4.8a-d offers valuable insights.
Fig.4.9a magnifies an area where a small detail emerges: a profound, narrow pit-like defect
likely present before fatigue cycling. This defect seemingly led to a crack that merged with the
primary one, implying it might have contributed to crack initiation.
Conventional pre-failure surface characterization (e.g. using a profilometer or an interfer-
ometer) could easily overlook such defects. The XCT characterization, however, accounts for
them. As can be seen in Fig.4.9e, the valley located behind the spherical spatter is indeed clearly
visible. Furthermore, Fig.4.9d-e show the pit-like defect is visible as well. To see it, one has to
look to the surface from the inside, i.e. as if one would stand in the core of the material. This
point of view is made possible by the XCT characterization and surface extraction. This show-
cases the capability of 3D XCT in delving deeper into crack initiation mechanisms, especially in
potentially uncovering obscured features commonly found in as-built surfaces.
Using the presented workflow, initiation sites were identified for 10 Z and 10 45° specimens.
Apart from the two crack initiations caused by spatters − i.e. in most cases, failure was
caused by surface valleys.
While both build orientations exhibited valleys, their morphologies varied. The morphology
of valleys in the 45° specimens is showcased in Fig.4.8d. The focus here, and in subsequent
discussions, is on valleys in the down-skin area, as failures in all specimens occurred in this
region. Compared to Z specimens, these valleys appear deeper and more irregular.
The Z specimens’ valleys, illustrated earlier in Fig.3.19, are generally shallower, horizontal,
and regular unless interrupted by prominent protrusions. Fig.4.10a-d depicts the identification
of a plate-pile valley as the crack initiation site on a Z specimen. It can be noted that for
the specimen chosen as an example, two initiation sites can be identified. This is a frequently
encountered case for both Z and 45° specimens, especially at high stresses.
Fatigue properties and crack initiation mechanisms for as-built specimens | 115
116 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Figure 4.10 | Workflow for crack initiation site identification using both SEM and
XCT scans (before and after failure). The specimen used as an example is a vertically built
one which failed after 129 339 cycles at 350 MPa. (a) SEM-BSE (topographical mode) fracture
surface observation. (b) Fracture surface obtained from XCT scan. Colors correspond to κmin
measurements performed using rcurv “ 20 µm and λS “ 10 voxel size. (c) Crack initiation site
observed on the fracture surface from the side of the specimen. (d) Same area as in sub-figure
c, identified on the initial XCT scan before fatigue failure. A valley is identified to be the cause
of crack initiation by comparison with (c).
Fatigue properties and crack initiation mechanisms for as-built specimens | 117
118 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Figure 4.11 | Identification of a secondary crack − i.e. a crack that didn’t propagate
until failure − in a lower section of the fracture specimen. The fatigue specimen in
question is the same as in Fig.4.10. Grey levels in all sub-figures correspond to a minimum
curvature measurement. (a) Fractured specimen observed from the inside to reveal the partially
propagated crack in black. This internal view is possible thanks to the fact that the surface was
extracted from XCT data. (b) Fractured specimen observed from the outside, in the same area
as in sub-figure a. The crack path is highlighted by white arrows. (c) Same area as in sub-figure
b, but before fatigue testing. It can be seen that the secondary crack highlighted in (a) and (b)
follows valleys that were initially present.
Fatigue properties and crack initiation mechanisms for as-built specimens | 119
120 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Figure 4.12 | Optical microscope micrographs revealing α-case for different heat
treatment conditions , from [GAI 23b, Fig.19]. Micrographs were taken at the edge of cross-
sections in the horizontal XY plane of (a) cubes before and after various heat treatments in
the industrial furnace, and (b) half tensile specimens before and after heat treatments in the
laboratory furnace. The etched material is obtained with a Weck reagent.
When performed in the industrial furnace, the heat treatment leads to comparable results
for temperatures lower or equal to 800 °C. Yet, at 860 °C and beyond, there is a reduction
in elongation to failure, contrasting with the previous increase. For instance, at 980 °C, the
elongation to failure after the industrial treatment is measured at 4.0 ˘ 0.8%, while a value of
18.7 ˘ 0.7% is obtained after the laboratory treatment. This significant drop in elongation
to failure stems from the growth of a pronounced α-case layer, resulting in surface
brittleness.
Evidence of this surface fragility is the profusion of cracks observable on the surfaces of tensile
specimens post-failure, as illustrated in Fig.4.14a.
These cracks remain confined to the brittle α-case layer. They likely serve to release the
tensile stresses, given the α-case layer’s inability to accommodate deformation due to its limited
ductility. The propensity of α-case to initiate cracks prompts questions about its
potential influence on fatigue damage. Consequently, this led to further research into the
effects of α-case on fatigue properties.
Figure 4.14 | Surface cracks formed in α-case layer during tensile testing. (a) Visual-
ization of cracks on the surface of a broken flat tensile specimen subjected to the industrial heat
treatment at 920 °C for 2 h. Visualization of cracks on a cross-section of the same specimen (b)
before and (c) after a chemical attack with Weck reagent. (d) Visualization of blunt cracks at
the α-case boundary. Drawn from [GAI 23b, Fig.13]
122 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
4.2.3 Impact on fatigue properties
4.2.3.1 Heat treatment conditions and resulting α-case and microstructure
Only Z specimens were used for this study, with the same geometry (3 mm diameter cylin-
drical gauge area) and L-PBF manufacturing conditions as the ones detailed in Section 3.1.
Three heat treatment conditions are compared in this study. They consist of a stress relief
possibly followed by a higher-temperature heat treatment:
• The stress relieved (SR) condition, which consists of a 720 °C|2 h treatment performed
in the industrial furnace used in Quentin Gaillard’s study. Note that it corresponds to
the state called as-built elsewhere in this manuscript (same thermal cycle and furnace).
It is a convenient denomination in the rest of the manuscript which focuses on the study
of surface topography, since it refers to the fact that the surface topography is kept in its
as-built state. However, this denomination would not be appropriate for this study which
focuses on heat treatments and their impact on α-case formation. Furthermore, the choice
has been made to call this condition SR in the present section to prevent confusion with
the work of Quentin Gaillard, where the as-built state refers to the condition without any
heat treatment.
• The HT condition refers to samples that received a 860 °C|4 h heat treatment in the
laboratory furnace after the stress-relief treatment in the industrial furnace.
• The HT|α-case condition pertains to samples treated at 860 °C for 4 h in the industrial
furnace following stress-relief.
The α-case layers observed for each condition are presented in Fig.4.15a-c.
Figure 4.15 | α-case characterization in SR, HT and HT|α-case conditions. The optical
micrographs were obtained from longitudinal cross-sections of fatigue specimens subjected to a
chemical attack with Weck reagent. The Weck reagent reveals the α-case layer which appears
in white. (Courtesy of Quentin Gaillard)
Section 4.2.2 showed that the thin α-case layer formed in the SR condition did not have any
significant impact on tensile properties, including elongation to failure. As a first approach, the
influence of α-case in the SR condition is therefore considered to be negligible, even for fatigue
properties. The only condition considered to induce a significant α-case layer is the HT|α-case
condition. This enables the separation of the influence of microstructure and α-case formation.
Figure 4.16 | SEM-BSE images showing the microstructure of the bulk material
(a) of a SR sample and (b) of an HT|α-case sample. The β phase appears brighter than the α
phase. (Courtesy of Quentin Gaillard)
To quantify the α-laths thickening, two EBSD maps were obtained on two fatigue samples
in the SR and HT|α-case states (HT|α-case and HT conditions are equivalent in terms of bulk
microstructure). The respective EBSD maps were 600 ˆ 445 µm2 and 600 ˆ 500 µm2 in size,
captured with a 0.25 µm step size. The determined α-laths thickness shifted from 0.61 µm in the
SR state to 1.06 µm for the 860 °C|4 h heat treatment.
124 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
4.2.3.2 Influence of α-case on fatigue properties
Fatigue tests on HT and HT|α-case samples were performed under the same conditions as
those used for the SR specimens (R “ 0.1 and f “ 10 Hz). However, due to time constraints,
tests were stopped here before 107 cycles when no failure occurred.
Fig.4.17 shows the Wöhler curves for the three conditions. Results for M&P specimens
presented in Section 4.1.1 are also shown as a reference.
1200
1000
σmax (MPa)
800
SR
HT
HT|α-case
600
M&P
2 4 Number of
3 superimposed points
400 3 4
200
104 105 106 107
Nf
Figure 4.17 | Wöhler curves for SR, HT and HT|α-case conditions. M&P specimens
are also represented as a reference. They did undergo a stress-relief treatment after polishing (2 h
at 660 °C), ensuring that no residual stress remains from machining. Fatigue tests are performed
at R “ 0.1. All specimens were manufactured vertically.
Figure 4.18 | SEM-SE observation of the final failure area of fatigue samples, reveal-
ing the brittle behavior of α-case layer (a) for an SR specimen that failed after 262 790 cycles
at 350 MPa, (b) an HT specimen that failed after 22 477 cycles at 550 MPa and (c) an HT|α-case
specimen that failed after 25 852 cycles at 450 MPa.
The observations indicate that the α-case layer can make the surface more brittle. This effect
is especially pronounced for the HT|α-case condition, which has the thickest α-case layer. Thin
brittle layers in the SR and HT conditions might also play a role. While it was previously stated
in Section 4.2.2 that thin α-case layers do not affect tensile properties, their influence on fatigue
could differ. They might, for example, facilitate the early stage of the crack initiation process.
It is interesting here to draw a parallel with the study of [VI 23]. The latter discussed the
influence of a 10 µm brittle oxide layer on the fatigue behavior (R “ 0.1) of 7075-T6 aluminum
alloy. By comparing specimens with and without this oxide layer, it has been shown that
the fatigue lifetime was reduced by a factor of 3 in the presence of the oxide layer. SEM
fractography and acoustic emission analysis during fatigue cycling were used to examine further
126 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
the mechanisms behind this observation. The findings suggested two potential mechanisms. The
first involves the oxide layer cracking, promoting crack initiation in the underlying material.
The second proposes that continuous crack formation in the oxide layer aids in the main crack’s
propagation, a phenomenon referred to as the continuous initiation process.
A parallel can be made here, with the oxide layer corresponding in the present case to the
brittle α-case oxygen-enriched metallic layer. In particular, the hypothesis of a continuous
initiation process seems coherent when looking at our fracture surfaces, which show comparable
appearances as the ones obtained by [VI 23, Fig.13] for specimens with the oxide layer. [VI 23]
noticed that when no oxide layer was present, cracks initiated from point defects. However,
when the brittle oxide layer was present, cracks seemed to initiate from a large circular segment,
possibly due to the occurrence of a continuous initiation process. Such large crack initiations
are typically observed in the present work, as can be seen in Fig.4.19a-c.
Figure 4.19 | SEM-BSE topological mode observation of fracture surfaces for (a)
a SR specimen that failed after 73 547 cycles at 300 MPa, (b) an HT specimen that failed af-
ter 40 901 cycles at 450 MPa and (c) an HT|α-case specimen that failed after 198 645 cycles at
300 MPa. In every case, crack initiation seems to occur from a large portion of the surface.
Yet, the observed fracture surface patterns are consistent across all conditions (SR, HT,
and HT|α-case) and are not specific to the HT|α-case. One possibility is that α-case always
induces a continuous initiation since some α-case layer exists in all treatments, even though it
seems discontinuous in the HT condition. A second explanation could be that the particular
fracture surface morphology observed comes from the presence of elongated surface valleys rather
than α-case. Indeed, it has been shown in 4.1.3.2 that cracks tend to follow surface valleys
inherited from the L-PBF process. This suggests that surface valleys could either promote
crack initiation or facilitate crack propagation by a continuous initiation-like mechanism. More
advanced characterizations would be required to better understand these mechanisms. Acoustic
emission analysis, as performed by [VI 23], would for instance have brought valuable information.
To verify if the presence of α-case could promote crack initiation, the crack density within
specimens in SR, HT and HT|α-case conditions was studied. For this purpose, the aim was to
count the number of cracks that did initiate, even though they did not propagate until failure.
In other words, cracks were searched below (or above) the fracture surface.
Figure 4.20 | Observation of final fracture cracks in HT|α-case specimens. (a) SEM-
SE image (b) Optical microscope image of a longitudinal cross-section after a chemical attack
with Weck reagent to reveal α-case in white.
Fig.4.20b reveals these cracks to be notably wide but shallow. Their propagation seems to
stop at the end of the α-case layer. They are very similar to cracks observed by Quentin Gaillard
[GAI 23b, Fig.13] on tensile specimens after failure, as shown in Fig.4.14z=a-d. Interestingly,
these cracks reside near the final fracture zone, where the material can be expected to bear very
high loads at the end of crack propagation. Towards the end of fatigue life, in the ligament
ahead of the propagating main crack, the material is submitted to high levels of stress This
likely explains the similarity between the fracture surfaces observed in the final fracture region
and the ones found on tensile specimens.
A closer look reveals the densest clustering of these cracks at the fringe of the final failure
region, as illustrated in Fig.4.20 and schematically represented in Fig.4.21.
Such cracks underscore the potential of the α-case layer to diminish fracture toughness,
accelerating the terminal failure. This behavior mirrors Quentin Gaillard’s findings during
tensile tests, illustrated in Fig.4.13b. Nevertheless, the presence of these cracks is only related
to the final failure phase, excluding the crack initiation or growth phases.
Therefore, in what follows, the focus has been made on a second type of crack. These initiated
and propagated during fatigue cycling, even though they did not necessarily propagate through
the whole specimen section. They will be referred to as fatigue cracks in what follows. Such
cracks are much slimmer and straighter, perpendicular to the loading direction, in contrast to
the final failure cracks. Moreover, their propagation often extends beyond the α-case layer, as
can be seen on the longitudinal cross-section in Fig.4.22.
128 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Crack initiation
Highest density
of final
failure cracks
Crack propagation
area
Final rupture
area
b 5
a
Fatigue crack linear density (mm-1)
SR
HT
4 HT| -case
Run-out
h 3
2
e
1.1⤫e 1
0
250 300 350 400 450 500 550 600 650 700
σmax (MPa)
Figure 4.23 | Fatigue crack linear density measurements on SR, HT and HT|α-case
specimens at different stress levels. (a) Scheme specifying the height h over which cracks
are counted for each specimen. (b) Cracks densities measured for all specimens, as a function
of the maximum stress during fatigue test.
The results shown in Fig.4.23b indicate that the crack density increases with the value of
maximum cyclic stress for all SR, HT, and HT|α-case specimens. Such a trend is anticipated,
aligning with findings that a greater number of nucleated cracks appear at elevated stress levels
[SCH 09, Fig.2.20].
Further analysis of Fig.4.23b reveals that, excluding the readings at 300 MPa, the HT|α-case
specimens exhibit considerably higher fatigue crack densities. This supports the theory that the
70 µm α-case layer in HT|α-case specimens promotes crack initiation during cycling,
potentially reducing fatigue lifespan. Densities for SR and HT specimens are similar, with a
marginally elevated count for SR. This could be due to the wider and more regular α-case layer
present in SR specimens, although it is only 10 µm wide.
A deeper exploration could shed light on how the α-case affects fatigue properties. For
instance, testing specimens with thicker α-case layers or M&P specimens with an equivalent
α-case layer would provide complementary information. Alternative characterization techniques
could have also proven useful. For instance, using acoustic emission, as explored by [VI 23], could
provide insights into the timing of crack initiation during cyclic loading and its relationship with
the α-case presence.
130 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
This research establishes that the 70 µm α-case layer developed during heat treatment slightly
diminishes the fatigue resistance of L-PBF Ti64 specimens. Combined with the surface rough-
ness considerations discussed in Section 4.1, it underscores the significance of L-PBF Ti64’s
surface integrity on fatigue properties. M&P specimens’ superior properties hint that strategic
surface polishing might significantly enhance fatigue attributes by both smoothing the surface
and eliminating the α-case layer. Although the impact of surface roughness seems much more
pronounced than the one of the α-case layer, the use of surface polishing could still be interesting
regarding the α-case issue. Indeed, one of the advantages of using material-removal polishing is
that it would enable heat treatments to be carried out with fewer constraints on the vacuum to
be achieved in the furnace, as the α-case layer (if of a reasonable depth) would be eliminated by
polishing.
However, polishing additively manufactured components poses challenges. Such components
typically have rougher surfaces than their machined counterparts. Moreover, additive manufac-
turing techniques like L-PBF are prized for making parts with intricate designs, making most
polishing methods unsuitable due to inaccessibility. Despite these issues, various technologies
aim to address these challenges [MAL 21, KHA 21, NUT 19, BOB 21]. In this PhD work, a
specific technique, Plasma Electrolytic Polishing (PeP), was explored to enhance fatigue prop-
erties.
Figure 4.25 | Principle scheme of PeP process drawn from [HUA 21].
132 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Because polishing mechanisms are different from the ones active in EP, PeP does not re-
quire the use of a cathode whose geometry is similar to the one of the workpiece.
Simple geometries such as rings or plates can therefore be used [ZEI 22], which is a huge advan-
tage for the treatment of parts of complex geometry. Some constraints still exist, like the need
for the cathode to be of significantly larger size than the workpiece [ZEI 22, NAV 22, HUA 21]
which could be limiting for large industrial parts. Furthermore, material removal rate di-
rectly depends on the immersion depth [ALE 05, BEL 20]. This issue comes from the
effect of hydrostatic pressure on the gas layer. Some inhomogeneity can therefore be expected,
especially for large parts. However, some solutions could exist in those cases, such as the use of
electrolyte jet PeP [HUA 21] where the electrolyte is used in the form of a jet propelled towards
the workpiece rather than a bath.
One of the characteristics of PeP is the high applied voltage and current density which
are needed for the gas layer to form and the polishing to occur. Typical voltage values range
from 200 to 350 V and current densities from 0.2 to 0.5 A/cm2 [HUA 21]. This makes PeP much
more demanding than conventional EP and can require the use of costly and energy-consuming
generators for large parts.
On a more positive side, electrolytes used for PeP (e.g. made of low-concentration salt
solutions) are generally considered to be less harmful to the environment than the ones
used in conventional EP [BAS 22, ZEI 22, NAV 22, HUA 21]. For titanium alloys, the most
commonly used electrolyte is a (low concentrated) fluorine salt-based aqueous solution [HUA 21].
On the whole, PeP presents interesting perspectives for the surface treatment of metallic
additive manufactured parts of complex geometries. However, as it is more recent than other
methods such as (electro-)chemical polishing or shot-peening, more studies are needed to esti-
mate its practical relevance.
134 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Figure 4.28 | PeP conditions for Z and 45° fatigue specimens. "Position" refers to the
specimen orientation of the specimen in the electrolytic bath. Red crosses indicate the number
of specimens for which treatment failed, see Fig.4.27 for an example. Note that the hole used
to hold the specimens in the bath is visible on the 45° specimen support structures. The hole
was mistakenly made for the 45° specimen in the vertical condition, but another hole was made
on the opposite head to allow the treatment to be performed in the vertical position.
Table 4.1 | Matter removals measured for the different PeP treatment conditions.
It corresponds to the difference between the minimum radius of the as-built specimens and the
minimum radius of the polished specimen, both measured from XCT data. The values shown
are the average and standard deviation for all specimens treated with the same condition.
The thickness of removed material is approximately 180 µm for specimens treated for 40 min
(i.e. Z|Vertical Vertical 45° Vertical condition, material removal
40min , 45°|40min , and 45°|40min conditions). For the Z|1h20
is nearly double, at approximately 340 µm.
Polishing at 45° resulted in a slightly higher matter removal than polishing in the vertical
position for 45° specimens, although the difference is small and could be considered within the
measurement uncertainty. Although no clear explanation was found, it could be attributed to
the specimen’s position in the bath. The position might alter the gas layer behavior around the
samples during PeP, influencing the material removal rate.
It should be noted that the polishing is not homogeneous across the specimen surfaces.
Fig.4.29 demonstrates that material removal thickness evolves along the specimen axis. Note
that each curve in Fig.4.29 represents a single specimen, not an average as in Fig.4.28.
450
400
Matter removal thickness (µm)
350
0
4 2 0 2 4
Position along specimen axis (mm)
Figure 4.29 | Matter removal thicknesses measured along the specimen axis of 4
specimens which underwent PeP with different treatment conditions. By convention,
the position along the specimen axis decreases in the direction of the built plate (for instance in
the direction of the residual support structures for 45° specimens.)
136 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Material removal appears to have a linear trend along the specimen axis. Maybe the most
striking point here is that the slope of the linear regression is negative for the Z|Vertical Vertical
40min , Z|1h20
45° Vertical
and 45°|40min conditions, whereas it is positive for the 45°|40min one. This can be correlated
to the position of the specimens during the PeP treatment. For the Z|Vertical Vertical and
40min , Z|1h20
45°
45°|40min conditions, the specimens are pointing downwards. Hence, the matter removal rate is
at its maximum for the smallest immersion depth. For the 45°|Vertical40min condition, specimens are
pointing upwards and therefore the positive slope shows that here again, the matter removal
rate is at its maximum for the smallest immersion depth.
However, these results seem difficult to explain regarding the existing literature. Indeed, the
effects of immersion depth on the material removal rate were found to be explained by variations
in the hydrostatic pressure. Hydrostatic pressure is an important parameter since it influences
the shape and width of the gas layer, which itself conditions the treatment efficiency. Such an
effect makes the material removal rate higher for higher immersion depths [ALE 05, BEL 20],
which is the exact opposite of what is observed in the present work.
Furthermore, the effect of hydrostatic pressure seem much less pronounced in the study of
[ALE 05] on a 12Cr-18Ni-10Ti steel than what is seen here. For instance, assuming a treatment of
40 min, the evolution of the thickness of removed material with immersion depth was according
to their measurements of 0.15 µm/mm, which is about 50 times less than what is observed
in the present study. One can also add that the material removal thickness evolution seems
very similar between the Z|Vertical 45°
40min and the 45°|40min conditions, although for this specimen the
immersion depth variation is smaller than for the 45°|45° 40min condition where the specimen is
oriented at 45°. Inversely, the slope is smaller of the 45°|Vertical
40min condition. All these observations
suggest that the material removal rate variations along the specimens’ axis are likely induced by
another factor than hydrostatic pressure variation. Since no other explanation could be found,
the observed material removal inhomogeneity will be simply taken into account as
they are in the subsequent analysis.
Chapter 3 highlighted that a unique microstructure and pore distribution are observed near
the surface. It is thus relevant to analyze the relationship between material removal thickness and
these distributions under different polishing conditions. Key findings are compiled in Fig.4.30.
Only the average matter removal presented in Tab.4.1 is shown here, but it is worth keeping in
mind that matter removal can vary among specimens and within a specimen itself.
Fig.4.30 superimposes:
• the distributions of the pores distance to the surface for both Z and 45° specimens
• a crystallographic orientation map (EBSD) of the reconstructed β0 grains close to the
surface
• vertical lines showing the mean value of the thickness of the removed material for each
PeP treatment condition
Comparing average material removals with pore distance distributions, the Z|Vertical
1h20 treat-
ment seems to entirely eliminate sub-surface porosity. For other conditions, the removal depth
aligns with the tail-end of sub-surface porosity. This could imply that residual sub-surface
pores could get exposed during polishing, forming new surface flaws that would negatively
impact fatigue resistance.
138 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Figure 4.31 | Surface of a Z specimen before and after a 40 min PeP treatment
(Z|Vertical
40min condition). The color used to represent the surface combines the roughness (using
λc “ 0.8 mm) as well as the minimum curvature (using rcurv “ 20 µm) to exacerbate valleys
and details in general. Although PeP completely smoothed the original surface roughness, a few
small pits can be identified on the surface. They are attributed to pores that were brought to
the surface by polishing.
Figure 4.32 | Surface of a Z specimen before and after a 1 h 20 min PeP treatment
(Z|Vertical
1h20 condition). The color used to represent the surface combines the roughness (using
λc “ 0.8 mm) as well as the minimum curvature (using rcurv “ 20 µm) to exacerbate valleys and
details in general. This 1 h 20 min treatment enables to erase all the initial roughness, and no
surface holes caused by pores can be seen as in Fig.4.31 for Z|Vertical
40min condition.
140 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
The influence of PeP on the 45°|Vertical
40min condition is demonstrated in Fig.4.33. Polishing
yields a uniformly smooth surface, but upon closer inspection, the down-skin region reveals
some pits. They bear resemblance in depth and form to those seen in the Z|Vertical
40min condition.
Vertical
Since the matter removal thickness for the 45°|40min condition (171 ˘ 7µm) is much larger than
the maximum valley depth of the as-built specimen (88 µm), it is considered here that these
defects originate as well from sub-surface pores that were brought to the surface by PeP.
Looking at the two magnifying windows, it is possible to notice that most of the surface pits
are found in the bottom region of the sample. This observation is consistent with the results
shown in 4.3.2.1. Indeed, Fig.4.30 showed that the polishing was done until the near-end of the
sub-surface pores. Fig.4.29 showed on its part that the matter removal thickness was lower at
the bottom of the specimen. Thus, the higher density of defect at the bottom in Fig.4.33 may
be explained by the lower matter removal in this region, which did not enable the complete
removal of the surface layer where most pores are located.
Figure 4.33 | Surface of a 45° specimen before and after a 40 min PeP treatment
performed vertically (45°|Vertical
40min condition). The color used to represent the surface com-
bines the roughness (using λc “ 0.8 mm) as well as the minimum curvature (using rcurv “ 20 µm)
to exacerbate valleys and details in general. A few small pits can still be identified on the surface
after PeP, especially at the bottom of the specimen.
Figure 4.34 | Surface of a 45° specimen before and after a 40 min PeP treatment at
45° (45°|45°
40min condition). The color used to represent the surface combines the roughness
(using λc “ 0.8 mm) as well as the minimum curvature (using rcurv “ 20 µm) to exacerbate
valleys and details in general. Large and deep valleys can still be identified at the top of
the specimen. They are remains of the down-skin surface roughness. Most of the valleys in
question are not taken into account by roughness measurements since they are out of the zone
characterized by XCT. An optical microscopy image shows that defects grow even larger outside
this zone.
142 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
However, since XCT scans only enable the characterization of a 9 mm high portion of the
specimens, these large defects are mostly not accounted for. The optical microscope observation
in Fig.4.34 shows that even larger defects can be seen in higher parts of the specimens that
are not characterized by XCT. Yet, as they are far away from the specimens’ gauge area, their
impact on fatigue tests might be minimal.
Finally, 45°|45°
40min specimens are the only ones for which large defects remain after PeP. These
defects are located in the upper part of the down-skin region. As evidenced in Fig.4.29, the
45°|45°
40min condition has minimal material removal at the top of the down-skin area. Additionally,
this region is closest to horizontal during creation, resulting in maximal roughness (see Fig.4.6c).
The combination of this initial roughness and minimal removal can explain the residual valleys
after PeP and their location in the upper part of the down-skin region.
It is important to notice here that the XCT characterization is limited to a height of 9 mm,
which is barely enough to reach the beginning of those defects. The optical microscope image
in Fig.4.34 shows indeed that many defects can be seen further in the radius fillet. Although
these defects are very large, the fact they are not situated in the specimens’ gauge length may
lead in practice to a limited impact on fatigue tests.
To understand PeP’s impact on roughness evolution more broadly, the evolution of Sa and
Sv roughness parameters were analyzed for each specimen. Vertical specimens’ roughness evo-
lution (Z|Vertical Vertical conditions) is showcased in Fig.4.35. Each specimen’s roughness
40min and Z|1h20
parameters, before and after PeP, are illustrated. The arithmetical mean height Sa on the x-axis
reflects the overall average surface condition, while the maximum valley depth Sv on the y-axis
sheds light on the deepest remaining surface defects.
60
Sv (µm)
40
20
Z
0
1 2 3 4 5 6 7 8
Sa (µm)
Figure 4.35 | Evolution of Sa and Sv roughness parameters induced by the PeP
treatment of Z specimens. Roughness was computed using λc “ 0.8 mm and λS “ 0.03 mm.
Dashed lines are represented only to enable the identification of points that were obtained before
and after the PeP treatment of a given specimen. They do not mean that a linear relationship
exists between Sa and Sv parameters.
125
Sv (µm)
100
75
50
25
45° down-skin
5 10 15 20 25
Sa (µm)
Figure 4.36 | Evolution of Sa and Sv roughness parameters induced by the PeP
treatment of 45° specimens in the down-skin region. Roughness was computed using
λc “ 0.8 mm and λS “ 0.03 mm. Dashed lines are represented only to enable the identification
of points that were obtained before and after the PeP treatment of a given specimen. They do
not mean that a linear relationship exists between Sa and Sv parameters.
Results are also more scattered, especially for Sv . For example, the 45°|45°40min specimen
illustrated in Fig.4.34 shows an abnormally high Sv value of 97 µm. Fig.4.34 showed the existence
of large remaining valleys in the upper part of the down-skin region, responsible for this high
measured Sv . Furthermore, the optical microscope observation has shown that even larger and
deeper defects are present further away from the gauge length. The same is true for other
45°|45°
40min specimens, which means that their relatively low value is due to the fact that the Z
span of the XCT scan is not sufficient to capture these defects. Sv values for the 45°|45° 40min
condition can thus be considered to be underestimated. However, one may also argue
that the large defects in question should not be taken into account, since they are too far from
the gauge length. This makes crack initiation unlikely in this area.
144 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
On the contrary, 45°|45°
40min specimens are smoother in the up-skin region than 45°|40min
Vertical
specimens, see Fig.4.37. They even show results comparable to Z|1h20 Vertical specimens with an
average S¯v of 5 µm. These results may indicate that surfaces oriented downwards during PeP
are better polished than surfaces oriented upwards, since for the 45°|45°
40min condition, the up-skin
region is oriented downwards.
50
40
Sv (µm)
30
20
10
45° up-skin
1 2 3 4 5 6
Sa (µm)
Figure 4.37 | Evolution of Sa and Sv roughness parameters induced by the PeP
treatment of 45° specimens in the up-skin region. Roughness was computed using λc “
0.8 mm and λS “ 0.03 mm. Dashed lines are represented only to enable the identification of
points that were obtained before and after the PeP treatment of a given specimen. They do not
mean that a linear relationship exists between Sa and Sv parameters.
Table 4.2 presents a variety of roughness parameters calculated after PeP across the four
tested conditions. For a more detailed description of roughness parameter computations, refer
to Chapter 2 and Section 3.4.3.2.
While many key insights from these results are highlighted in previous figures, a notable
observation is the predominance of negative skewness Ssk in most cases. This contrasts with
the mainly positive skewness seen in as-built specimens. This difference arises because, post-
PeP, the surface mostly consists of smooth areas interspersed with occasional notches that have
negative local heights. Such a pattern is characteristic of surfaces dominated by valleys, which
result in negative skewness.
Regarding kurtosis, elevated values in some conditions can be correlated to the existence
of occasional, yet notably deep, defects. The observed variability suggests that the occurrence
and depth of these defects can differ significantly between specimens. As a result, variations
in fatigue behavior might also be anticipated. Conversely, the low kurtosis value for
conditions like Z|Vertical
1h20 signifies the lack of exceptionally deep surface irregularities. It indicates
that PeP effectively produces a consistently smooth and even surface. Such a surface is
anticipated to exhibit superior fatigue resistance.
Table 4.2 | Roughness parameters measured for specimens polished by PeP. Rough-
ness was computed from XCT data using the 3D characterization workflow introduced in Chapter
2. The cut-off values for filters were set to λc “ 0.8 mm and λS “ 0.03 mm. Roughness param-
eter values are presented in the format (mean ˘ std) for conditions where at least 3 specimens
were available. Otherwise, only the mean value is given.
Having characterized the surface state post-PeP, the next step is to assess its impact on
fatigue properties. Uniaxial fatigue tests were conducted under conditions identical to those
of as-built specimens, with details available in Section 4.1.1. Unlike as-built specimens, the
PeP specimens’ fatigue tests were stopped at 2 ˆ 106 cycles when no failure occurred. Non-
failed specimens underwent further testing at elevated stresses for 2 ˆ 106 cycles, increasing the
maximum stress value until failure and starting again the cycle count from zero at each stress
level. On Wöhler curves, 2 ˆ 106 cycles run-outs get marked by an arrow, and points from the
same specimen tested under varying stresses are connected with a dashed line.
Due to the substantial material removal by PeP, XCT was used to measure the specimen’s
section to set stress levels correctly. The average section across the gauge length was used.
Results from M&P and as-built specimens are shown to allow for comparison and better gauge
the enhancements resulting from PeP.
The S-N diagram in Fig.4.38 displays results for 45° specimens. It can be seen from Fig.4.38
that PeP significantly enhances fatigue life. A comparable enhancement was observed for both
45°|Vertical 45°
40min and 45°|40min conditions, though this would require further specimen testing for vali-
dation.
Considering available data, the fatigue strength for 45° specimens post-PeP polishing is esti-
mated between 450 MPa and 550 MPa. This translates to an 80%-120% improvement over
as-built specimens. Although significant, this enhancement is still far from the one provided
by machining and manual polishing. The limited improvement can be primarily attributed to
pores either exposed on the surface or slightly beneath it by PeP. Fig.4.39a-b illustrate
two examples of crack initiation sites on 45°|Vertical
40min specimen fractures where such defects can
be identified.
146 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
1200
45° R = 0.1
1000 As-built
PeP
σmax (MPa)
PeP
800 M&P
2 Number of
3 superimposed
points
600 2
2
400 32
2
200
104 105 106 107
Nf
Figure 4.38 | Wöhler for 45° specimens before and after PeP in both 45°|Vertical
40min and
45°
45°|40min conditions. Results for 45° as-built specimens and M&P specimens are also shown
for comparison.
Fig.4.39a corresponds to the failure on a pore brought to the surface by PeP. It is the most
frequent initiation site observed. It can be seen that the inner surface of the defect is somehow
degraded, probably under the effects of PeP treatment (see Fig.4.39b for comparison with a
subsurface pore with a smooth wall/inner surface).
Fig.4.39b shows another case, where the failure occurred from a sub-surface pore that was
brought just beneath the surface by PeP. These two examples confirm the detrimental effect of
the few remaining defects on the fatigue properties of specimens polished after PeP, especially
if the polishing depth falls close to the sub-surface layer where most pores are located.
For one 45°|45°
40min specimen, the main crack initiated from a residual surface valley which was
not entirely suppressed after PeP. The specimen is the one shown in Fig.4.34, which presented
the largest defect among all 45°|45°
40min specimens. The initiation site can be seen in the SEM-SE
observation in Fig.4.40a.
One crack initiation site found on another 45°|45°
40min specimen, shown in Fig.4.40b, could also
correspond to a residual surface valley, although its origin could not be precisely identified. It
may as well be a pore brought to the surface as the ones in Fig.4.39a-b. Nonetheless, this shows
that insufficient polishing, especially in down-skin areas, residual valleys can facilitate fatigue
failure. However, similar fatigue lives for both 45°|Vertical 45°
40min and 45°|40min suggests this is not a
major influence on fatigue resistance in our case. This can be explained by the fact that residual
valleys are mostly located in the upper part of the specimen, far away from the gauge area.
Figure 4.40 | Crack initiation sites of 45°|45° 40min specimens, observed by SEM-SE.
(a) Residual surface valley in the upper part of the down-skin region. This observation has
been observed on the same specimen that was used for Fig.4.34. The specimen failed after
2 ˆ 106 cycles at 450 MPa followed by 91 990 cycles at 550 MPa. (b) Crack initiation site whose
origin could not be found with certainty. It corresponds most probably to a residual surface
valley or a pore brought to the surface by PeP. The specimen failed after 2 ˆ 106 cycles at
550 MPa followed by 55 086 cycles at 650 MPa.
148 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
1200
Z R = 0.1
1000 2
As-built
2
PeP
σmax (MPa)
800 PeP
M&P
Number of
600 4 3 superimposed
points
400 4
4
3
200
104 105 106 107
Nf
Figure 4.41 | Wöhler for Z specimens before and after PeP in both Z|Vertical 40min and
Vertical
Z|1h20 conditions. Results for Z as-built specimens and M&P specimens are also shown for
comparison.
Fig.4.41 presents the S-N diagram for Z specimens. The data indicates that the Z|Vertical
40min
polishing yields lesser improvements than observed for 45°|Vertical
40min specimens. The enhancement
in terms of fatigue strength at 2 ˆ 106 cycles can be estimated to be around 150 MPa. In terms
of initiation site, all failures of Z|Vertical
40min specimens occurred from pores brought to the surface
by PeP, similarly to what has been shown in Fig.4.39a.
However, the Z|Vertical
1h20 condition displayed impressive fatigue resistance, as evi-
denced by one specimen enduring successive cycles at increasing stresses and finally failing at
1100 MPa. As a reminder, tensile tests in Section 3.3 enabled to estimate the yield strength of
Z specimens to be about 1098 MPa. This exceptional resistance might result from the gauge
section being particularly small in the Z|Vertical
1h20 condition, reducing the probability of having a
large defect.
Thus, Fig.4.35 showed for instance that no significant surface defects could be found on
Z|Vertical
1h20 using the XCT scan with a 3 µm voxel size. As a result, the crack leading to failure
initiated from internal defects in this case.
In terms of porosity, only a few pores could be identified on XCT scans in both Z|Vertical
1h20
specimens. For instance, only 5 pores could be identified within the specimen that showed the
highest fatigue resistance. Apart from one having an equivalent diameter of 32 µm, all have
equivalent diameters inferior to 13 µm. The sub-surface pore that caused crack initiation could
not even be identified on the XCT scan due to resolution limitation. It is shown on the SEM-SE
fractography in Fig.4.42a.
150 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Figure 4.43 | Evolution of fatigue resistance and crack initiation mechanisms with
increases PeP matter removal. Contour pores refer to the pores located at about 140 µm
from the surface, between the contour scans and the bulk material (see Section 3.4.2).
Z|Vertical
40min Z|Vertical
1h20 45°|Vertical
40min 45°|45°
40min
Pore brought to the surface 5 2 2
Pore brought near the surface 1 1
Residual valley 1
Unidentified nature 1 1
Table 4.3 | Number of crack initiation on surface defects (mostly pores brought to
the surface by PeP) and sub-surface pores for each PeP condition. The number of
defects that could not be identified on the surface extracted from the XCT scan is also specified.
This issue comes either from a lack of resolution or simply from the fact that the sub-surface
defect had no connection on the surface and was therefore not present in the extracted surface.
For as-built specimens, all failures originated from the surface. Detailed insights from 3D
XCT characterization before and after failure reveal that killer defects are primarily surface
valleys. These drastically reduced the fatigue resistance of the material, evident when
comparing against the fatigue resistance of vertically built, machined and manually polished
(M&P) specimens. Fatigue strength at 2 ˆ 106 cycles for as-built specimens, regardless of
build orientation, is found close to 250 MPa. In contrast, M&P specimens reached a fatigue
strength of roughly 850 MPa. There is a small yet notable difference in the fatigue life
between vertically built and 45° specimens. The latter display shorter fatigue lives, likely
due to the increased roughness in their down-skin area, which consistently was the site of
failure.
Another part of the study concerned specimens subjected to a more intense heat treatment
(4 h at 860 °C). This focused on the effects of the α-case, a brittle oxygen-contaminated
surface layer forming during heat treatments. In our case, the treatment yielded a 70 µm
layer. This layer further compromised the fatigue resistance of as-built specimens, though
possibly not as seriously as surface roughness. The diminished fatigue life could be due to
the α-case’s tendency to promote crack initiation.
Plasma Electrolytic Polishing (PeP), a recent surface polishing method, was tested as a
solution to mitigate both roughness and α-case in as-built surfaces. While some specimens
experienced uneven material removal or unexplained polishing failures, the process effec-
tively minimized initial surface roughness. Because material removal during a 40 min PeP
treatment was insufficient (about 180 µm), sub-surface pores were brought to the surface,
leading to the formation of surface pits. Consequently, cracks emerged from these exposed
pores, limiting the improvement in fatigue resistance. The fatigue resistance at 2 ˆ 106 cycles
after such treatment improved from 250 MPa to approximately 400 to 550 MPa. A more
prolonged PeP treatment (1 h 20 min) polished beyond this porosity overconcentration layer,
significantly enhancing fatigue resistance. Thus, the fatigue strength at 2 ˆ 106 cycles for
these specimens was larger than 1000 MPa, slightly surpassing M&P samples.
While PeP offers enhanced fatigue properties, it may not be feasible for many practical
scenarios. Factors such as complex part shapes or prohibitive post-processing costs and time
considerations might hinder its adoption. Hence, understanding, modeling, and predicting
the influence of as-built surface defects on fatigue properties remain crucial. Traditional
?
models, like the Murakami parameter area, fall short when addressing the complexities
of surface valleys leading to failure. Thus, the search for alternative methods is explored in
Chapter 5.
152 | Chapter 4 – Fatigue properties and crack initiation mechanisms of as-built and
post-treated specimens
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Chapter 5
Contents
5.1 Surface defects segmentation . . . . . . . . . . . . . . . . . . . . . . . . . 154
5.2.2 Correlation between the selected parameters and the fatigue lifetime . . . . 163
Chapter 4 primarily focused on the description of fatigue properties and defects at the origin
of failure. This chapter aims at rationalizing the impact of surface defects on fatigue, based
on the analysis of XCT data. Specimens in their as-built state and those polished by PeP will
be analyzed. Specimens subjected to a heat treatment at 860 °C have been excluded from our
analysis to alleviate the effect of α-case formation or changes in microstructure.
The first section introduces the methodology employed to segment surface defects, i.e. to
identify them. In other words, this step consists of going from a description of the surface as a
scalar field (= values such as Kt˚ at each surface voxel) to a defect-based description (= a single
Kt˚ value associated with each segmented defect). This defect segmentation will be used in the
subsequent sections.
| 153
The segmentation method was chosen to best overcome these difficulties. Its different steps
are detailed hereafter for an as-built (45°) specimen. Examples are then shown for the segmen-
tation of potential killer defects (i.e. valleys or pit-like defects) for all surface states, and finally,
other examples are given for the segmentation of protrusions.
As shown in Section 4.1.3.2, valleys are most easily identified based on their low κmin or κσ
values (the sign being taken into account). Thus, defect points are first separated from the rest
of the surface by thresholding, based on the local curvature.
While κmin successfully reveals valleys, it does not seem suited for their segmentation since
it does not allow to differentiate those that are parallel to the specimens’ axis from those that
are perpendicular to it. Therefore, using κmin would make it very difficult to segment individual
valleys, as many of the valleys perpendicular to the specimen axis would be connected by valleys
that are parallel to the specimen axis. This would lead to the segmentation of valley networks,
especially in down-skin regions of 45° specimens, see Fig.5.2c.
To avoid this problem, the curvature oriented in the direction of the specimens’ axis, κσ , is
used instead of κmin . A drawback of this method is that when computed at a fine scale, κσ
is more noisy than κmin . For this reason, the curvature is computed here at a relatively large
scale, i.e. using rcurv “ 50 µm.
The value for the threshold is determined automatically based on the histogram of the κσ
values. The method is similar to that of the triangle threshold for bimodal histograms
(TTBH) introduced in Section 2.2.1. The definition of the threshold is schematically illustrated
in Fig.5.1.
As for the TTBH, the value of the threshold is determined by maximizing the distance d
between the hypotenuse of the triangle (orange dashed lines in Fig.5.1) and the normalized
histogram. This enables the separation of the main histogram peak and the left tail, which
contains points with the lowest κσ , i.e. defect points. However, the obtained threshold value is
not always the most suitable. In some cases, slightly increasing or decreasing the threshold value
allows to achieve better results. To allow for such flexibility, an offset parameter is included and
can be set between -1 and 1. This modifies the final threshold value as schematized in Fig.5.1.
Fig.5.2a-f show the surface before and after thresholding of a 45° specimens (offset “ `0.4), in
both down-skin and up-skin regions. The surface is represented in grayscale using the minimum
curvature computed at a rather low scale (rcurv “ 20 µm) to provide a detailed visualization
of the surface. Fig.5.2c shows the complex network of interconnected valleys that would be
identified if κmin was used instead of κσ . Contrariwise, the use of κσ breaks this network
and results in separated defects, see Fig.5.2d,f.
A clustering algorithm called DBSCAN is then used to gather defect points together, and
thus obtain a set of isolated and well-identified defects. Details on how this algorithm works
can be found in [EST 96]. Two parameters are required in this model: Eps and MinPts. The
meaning of the MinPts will not be discussed extensively here, since its value has always been set
to 1 in our case. With MinPts“ 1, the DBSCAN algorithm simply clusters together all points
that are separated by a distance shorter than Eps.
dmax
+offset
-1 0 0.4 1
Normalized
Threshold curvature
-1 Decreasing κ 0 Increasing κ 1
Figure 5.1 | Definition of the method for the automatic determination of the thresh-
old for valleys thresholding, based on κσ values. The threshold is computed using the
normalized histogram computed for κσ prcurv “ 50 µmq by KDE.
A small adjustment is introduced here to further improve the ability of the model to cluster
elongated valleys that are perpendicular to the specimen axis, without connecting two parallel
transverse valleys that are close to one another. To do this, all points are dilated along the
specimen axis by a factor aspect_ratio, e.g. equal to 5 for as-built 45° specimens. In other
words, the component of each point along the specimen axis (i.e. the Z component for vertical
specimens) is multiplied by aspect_ratio, which artificially increases the distance between points
along the specimen axis. This prevents to some extent the algorithm from merging two
parallel valleys that are very close to each other at some point.
After clustering, many of the identified clusters are small in size. This leads to high numbers
of clusters that are not necessarily significant. For instance, 2949 clusters were found for the
specimen shown in Fig.5.2. To reduce this number and keep only the most significant clusters,
only the ones having a size greater than min_cluster_size were kept. The minimum cluster size
was set at 500 for as-built 45° specimens, which resulted in a final number of clusters of 692.
Table 5.1 | Parameters used for the main valleys segmentation in each surface state condition:
as-built Z or 45°, as well as specimens polished by PeP (all conditions included).
The result of the clustering for the as-built 45° specimen shown in Fig.5.2 is presented in
Fig.5.3a-b. Likewise, the results of the clustering for an as-built Z specimen and for a Z specimen
polished by PeP are shown in Fig.5.4a-b and Fig.5.5a-b respectively.
Figure 5.3 | Final valleys segmentation in (a) down-skin and (b) up-skin regions of
an as-built 45° specimens. A random color is attributed to each cluster.
Overall, the quality of the segmentation seems reasonable given the resolution of the 3D
images. However, we can see in Fig.5.3a-b and Fig.5.4a-b that a significant proportion of
the valleys are not or only partially captured, a problem that could be circumvented by
using more advanced segmentation techniques and/or higher resolution XCT scans.
Since protrusions have much larger width and height, and exhibit a more rounded morphology,
their segmentation is easier than that of valleys. The κσ parameter is not necessarily the most
suited in such a case. A more simple roughness measurement is contrariwise very well adapted.
Thus, the methodology presented for valley segmentation can simply be applied, using the
roughness measurement instead of κσ . Clusters will thus be formed from surface points that
have the highest height. To do that, the thresholding method illustrated in Fig.5.1 is adapted
to isolate the right tail of the height distribution (highest values) instead of the left tail.
Eps and aspect_ratio parameters were taken equal to 2 and 1 respectively, which roughly
means that only points in direct contact will be clustered together, without any preferential
direction. offset and min_cluster_size parameters are taken equal to -0.25 and 1000 respectively,
for both as-built Z and 45° specimens. The segmentation of protrusions has not been performed
on specimens polished by PeP, since none is left after polishing.
As an example, the result of protrusions segmentation in the same area of the Z specimen
shown in Fig.5.4 is presented in Fig.5.6a-b. All the main protrusions are identified.
The aim of this section is to test the ability of different parameters (e.g. the roughness
parameter Sa ), calculated for a given specimen, to inform about the impact of surface defects
on the fatigue resistance.
Sz˚
ˆ ˙ˆ ˙
Sa
K̄tAR “1`2 ˚ (5.1)
ρ̄ S10z
Sa is the arithmetic mean roughness. Sz˚ and S10z ˚ definitions are close to the ones of standard
ISO parameters Sz and S10z [Int 21]. Still, it is slightly different because the method used in
our case for valley segmentation (presented in the previous section) is different from the one
advised in the ISO standard [Int 21]. Thus, in the present case, Sz˚ is the sum of the highest
peak height and the deepest valley depth, whereas S10z ˚ is the sum of the average depth of the
five deepest valleys and the average height of the five highest peaks. Finally, ρ̄ is the average
radius of curvature measured on the five deepest valleys. The radius of curvature is calculated
here as the inverse of the curvature along the loading direction. Valleys and peaks are segmented
using the method presented in the previous section. When less than 5 valleys or 5 peaks are
detected, the average is computed from the few detected ones.
The last parameter that is evaluated is the maximum value of the estimated stress
concentration factor Kt˚ measured on a whole specimen, denoted Kt,max ˚ . A slight
˚
modification was made compared to the definition of Kt given in Section 2.2.4. This modification
aims to take into account the variation of the section along its axis. Since the specimen
section is larger within a fillet radius, the stress experienced by the material is reduced in this
area. Kt˚ is thus corrected at each point of the surface by the ratio of the minimum section
along the whole specimen (i.e. within the gauge length) Smin and the section S at this point.
The resulting formula for Kt˚ is given in Eq.5.2:
Smin ´ a ¯
Kt˚ “ ¨ 1 ` 2 height ¨ κσ (5.2)
S
Smin
Fig.5.7a-d show the Kt˚ values before and after correction, as well as the S factor, measured
on a 45° specimen.
Relationships between fatigue resistance and parameters derived from the surface
characterization using XCT | 161
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
Figure 5.7 | Correction of the Kt˚ parameter to take into account the variation of
the section along the specimens’ axis. (a) Kt˚ map before correction, (b) Smin S correction
factor, (c) Kt˚ map after correction, and (d) Internal view of the Kt˚ map after correction.
5.2.2 Correlation between the selected parameters and the fatigue lifetime
To evaluate the correlation between the four parameters defined previously (Sa , Sv , K̄tAR
˚
and Kt,max ) and the fatigue resistance, S-N diagrams are created for each parameter, with data
points color-coded according to the parameter value. Figure 5.8a provides an ideal example
of such a diagram, assuming a strong correlation between the investigated parameter and the
fatigue resistance.
b
a
+
Nf ≈ cste
+ σmax
Parameter (harmfulness)
Parameter +
c
σmax
σmax = cste
+
Nf +
Nf
Parameter +
Figure 5.8 | Schematic representation of an ideal correlation between fatigue re-
sistance and a given parameter representative of surface defects harmfulness. (a)
Hypothetical S-N diagram that would be obtained if the fatigue lifetime of specimens was di-
rectly correlated to the harmfulness parameter. (b) Plot of σmax as a function of the harmfulness
parameter, for specimens that failed after a given number of cycles. (c) Plot of Nf as a function
of the harmfulness parameter, at a given maximum stress level.
In Fig.5.8a, specimens with defects that are not severe (low harmfulness, reflected by the
selected parameter) tend to have longer lifespans or endure higher maximum stresses. This
unified graph representation allows for a comprehensive overview.
Relationships between fatigue resistance and parameters derived from the surface
characterization using XCT | 163
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
To assess possible correlations, the detailed plots in Fig.5.8b-c are helpful. These graphs
plot the maximum stress σmax or the number of cycles to failure Nf against the harmfulness
parameter. A strong correlation in these plots indicates the parameter’s reliability in predicting
fatigue resistance. However, caution is needed when generalizing these results, as they require
plotting data for a single σmax or Nf value only.
Plots are first shown for Sa in Fig.5.9a-c. All surface states (e.g. as-built and PeP) are
represented on the same plot. This enables to evaluate whether Sa is relevant to reflect defects
harmfulness for very different surfaces.
700 b
a
Sa (µm) 600 Nf ∈ [5e4, 7e4]
σmax (MPa)
1200
500
10
1000 400
8 300
σmax (MPa)
800 2 6 10
Sa (µm)
6 3·105 c
600
σmax = 350 MPa
2·105
4
Nf
400
2 105
200
104 105 Nf 106 107 6·104
7 9 11
Sa (µm)
Figure 5.9 | Correlation between Sa and the fatigue resistance of specimens with
different surface states. (a) S-N curve with points coloring according to the specimens’
Sa . (b) σmax vs. Sa plot for point with Nf P r5 ˆ 105 , 7 ˆ 104 s. (c) Nf vs. Sa plot for
σmax “ 350 MPa.
The S-N diagram in Fig.5.9a shows that some correlation exists between Sa and the fatigue
resistance. Specifically, specimens with a lower Sa , such as those polished by PeP, tend to have
longer fatigue lives and withstand higher stresses.
Additionally, the plot of fatigue life (Nf ) against Sa in Fig.5.9c shows a fairly good correlation
at 350 MPa. Sa effectively captures the fact that, on average, as-built Z specimens
reach higher fatigue lives than as-built 45° specimens which exhibit higher Sa values.
However, there is a noticeable scatter: for example, some Z specimens have similar fatigue
lives in comparison with 45° specimens, despite having significantly different Sa values. The
quality of the correlation also depends on the stress levels. For instance, at 300 MPa, an as-built
700 b
a Sv (µm)
1200 600 Nf ∈ [5e4, 7e4]
σmax (MPa)
160
500
1000
400
120
σmax (MPa)
800 300
20 40 60 80
Sv (µm)
80
600 3·105 c
2·105
σmax = 350 MPa
400 40
Nf
105
200 5 7
104 10 Nf 10 6
10
6·104
60 100 140
Sv (µm)
Figure 5.10 | Correlation between Sv and the fatigue resistance of specimens with
different surface states. (a) S-N curve with points coloring according to the specimens’ Sv . (b)
σmax vs. Sv plot for point with Nf P r5 ˆ 105 , 7 ˆ 104 s. (c) Nf vs. Sv plot for σmax “ 350 MPa.
In Fig.5.10b, the use of Sv avoids the saturation effect seen with Sa for PeP-polished spec-
imens. However, the correlation of Sv with fatigue life (Nf ) is not as good as that
of Sa . Notably, for as-built 45° specimens at 350 MPa, the highest fatigue lives paradoxically
correspond to the higher Sv values. Thus, while Sv generally shows that as-built Z specimens
have higher fatigue lives and lower Sv values on average than as-built 45° specimens, it fails to
adequately explain the scatter in fatigue lives among specimens with similar surface states.
Relationships between fatigue resistance and parameters derived from the surface
characterization using XCT | 165
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
700 b
a 600
Nf ∈ [5e4, 7e4]
σmax (MPa)
1200
3.00 500
1000 2.75
400
2.50
300
σmax (MPa)
2.00
600
3·105 c
1.75 σmax = 350 MPa
2·105
1.50
400
Nf
1.25 105
200
104 105 Nf 106 107
6·104
1.8 2.4 3.0
Figure 5.11 | Correlation between K̄tAR and the fatigue resistance of specimens with
different surface states. (a) S-N curve with points coloring according to the specimens’
K̄tAR . (b) σmax vs. K̄tAR plot for point with Nf P r5 ˆ 105 , 7 ˆ 104 s. (c) Nf vs. K̄tAR plot for
σmax “ 350 MPa.
Fig.5.11a-c presents results for K̄tAR which lie between those of Sa and Sv , leaning more
towards Sa . This was somewhat expected, since K̄tAR is derived from Sa and Sz˚ (Sz˚ is the
average of the deepest valley depth and the highest peak height and is thus close to Sv ). The
contribution of curvature (at least with a convolution radius of 20 µm) appears to
be limited in this case.
˚
Fig.5.12a-c displays plots for Kt,max , showing trends similar to those for Sv . This similarity
˚
is likely due to the fact that Kt,max , like Sv , is a parameter that mainly characterizes the most
˚
critical defect. Thus, the correlation of Kt,max with Nf at 350 MPa (Fig.5.12c) is not very good,
despite the clear distinction between the average behaviors of as-built Z and 45° specimens.
In contrast, the correlation between σmax and Kt,max ˚ in Fig.5.12b is much more
pronounced and effectively differentiates between varying surface states, i.e. for both
˚
as-built and PeP-polished specimens. It could be expected that a parameter such as Kt,max
˚
yields a better correlation with σmax than Nf , since Kt,max quantifies a stress concentration
˚
effect. Fig.5.12a also shows that Kt,max allows to better account for the transition between
as-built and PeP specimens, without the distinct separation between the two as was noticed
with Sa .
σmax (MPa)
1200
7 500
1000 6 400
300
σmax (MPa)
800 5 3 4 5 6
4 3·105 c
600
2·105
σmax = 350 MPa
3
400
Nf
2 105
200
104 105 Nf 106 107 6·104
4.5 5.5 6.5
Figure 5.12 | Correlation between Kt,max ˚ and the fatigue resistance of specimens
with different surface states. (a) S-N curve with points coloring according to the specimens’
˚
Kt,max ˚
. (b) σmax vs. Kt,max plot for point with Nf P r5 ˆ 105 , 7 ˆ 104 s. (c) Nf vs. Kt,max
˚ plot
for σmax “ 350 MPa.
A possible solution to improve these results, especially the correlation between parameters
and Nf , would be to consider the shape of surface defects, particularly for as-built specimens. For
˚
example, Kt,max being a local measurement, it does not account for the lengths of surface valleys.
Yet, as discussed in Section 4.1.3.2, the elongated nature of valleys might be a critical factor in
fatigue failure. To estimate whether taking this factor into account could improve predictions,
the length of the crack initiation zone (ℓCIZ ) was manually measured on the fracture surfaces of
as-built specimens using SEM. ℓCIZ represents the length of the surface perimeter from which
river lines appear to originate. Assessing whether including valley length improves the predictive
˚
accuracy of Kt,max could be an interesting pathway in the future.
Fig.5.13a demonstrates that at higher maximum stress levels, the values of ℓCIZ (length
of the crack initiation zone) tend to be higher. This trend is likely because increased stress
levels promote the initiation of multiple cracks [SCH 09, Fig.2.20], leading to larger ℓCIZ values.
Additionally, there is a noticeable trend because specimens with smaller crack initiation zones
achieve longer fatigue lives at a given stress level, as shown in Fig.5.13b. This observation
suggests that considering the length of valleys might be essential for a more accurate
prediction of their impact on fatigue resistance.
Relationships between fatigue resistance and parameters derived from the surface
characterization using XCT | 167
Cette thèse est accessible à l'adresse : https://theses.insa-lyon.fr/publication/2024ISAL0014/these.pdf
© [F. Steinhilber], [2024], INSA Lyon, tous droits réservés
a (mm)
600 >4
Z|as-built b
550 3.5 3·10 5
45°|as-built
500 3 2·105 σmax = 350 MPa
σmax (MPa)
Nf
450 2.5
105
400 2
6·104
350 1.5
0.5 1.5 2.5
300 1 (mm)
Z|as-built 45°|as-built
250 < 0.5
5 7
104 10 Nf 10 6
10
Figure 5.13 | Correlation between ℓCIZ and the fatigue resistance of as-built spec-
imens. (a) S-N curve with points coloring according to ℓCIZ . (b) Nf vs. ℓCIZ plot for
σmax “ 350 MPa.
6 6
Sv values of
5 PeP specimens 5
Median rank - κσ
4 4
3 3
2 2
1 1
0 0
0 25 50 75 100 125 150
rcurv (µm)
Figure 5.14 | Median ranks obtained for κσ and Kt˚ for the different values tested
for rcurv . The range of Sv values obtained for PeP specimens (except the Z|Vertical
1h20 condition) is
given in green.
Probability density
(surface points Kt*) PeP
3 PDF (≈ histogram)
of Kt* values of surface points
0 0
Probability density
PDF of max Kt values*
0.5 (defects max Kt*)
of segmented defects
1 2 3 4
Figure 5.15 | Distributions of surface points and defects Kt˚ values for 45°|Vertical
40min -2
˚
specimen. The blue line shows the PDF for Kt values of surface points, considering only
points with negative heights and κσ . The orange points and curve represent the maximum Kt˚
value for each segmented defect and its PDF. The killer defect is marked in purple. PDFs are
calculated using the KDE method with Silverman’s rule of thumb for bandwidth selection.
Table 5.2 | Killer defect predictions scores (rank & percentage) for specimens pol-
ished by PeP using roughness, κσ and Kt˚ parameters. The Specimen column refers to
the name of individual specimens and the ndef ects column gives the total number of defects that
were automatically segmented. Colors are given for guidance, the color tending to green for a
successful killer defect prediction while it tends to red for a less accurate prediction.
First, we can notice that for 5 specimens out of 14, the killer defect could not
be captured by XCT due to a lack of resolution. Therefore, the analysis could not
be conducted for those specimens. For the remaining specimens, the ranks obtained using
roughness, κσ or Kt˚ are all very similar. This is likely due to the small and point-like shape of
the defects in PeP specimens. Notably, their point-like shape reduces the benefit of using the
curvature oriented in the direction of the specimen axis.
Among the 8 specimens that could be analyzed, the killer defect could be suc-
cessfully predicted in 4 cases using the Kt˚ parameter. These results are comparable to or
even slightly better than the ones obtained on chemically polished E-PBF samples in Section
2.4. This may be because the surface obtained for L-PBF polished by PeP is smoother and the
few remaining defects show a more regular/globular morphology.
5.3.1.2 The impact of XCT scan resolution on the prediction of killer defect
The objective here is to compare the killer defect as characterized by XCT with its charac-
terization with SEM. Since SEM provides a better resolution than XCT, it is considered as a
ground truth. This is possible for specimens polished by PeP since defects can be unambiguously
delineated on SEM observations, which is not possible for valleys in as-built specimens.
Four cases will be successively discussed:
• a case where the killer defect is effectively captured by XCT, leading to a successful pre-
diction (example 1);
• a case showing that the method can also be used to account for near-surface internal pores
(example 2);
• a case where the limited XCT scan resolution enables to only partially capture the killer
defect, leading to a much less accurate prediction (example 3);
• a case where the killer defect could not even be found in the surface extracted from XCT
scans (example 4).
a b c
Killer defect Otsu (124/255) TTBH (195/255)
30 µm Specimen axis
Figure 5.17 | Comparison of the abilities of Otsu’s threshold and TTBH to capture
the killer defect shown in Fig.5.16. (a) Transverse XCT slice (voxel size = 3 µm). (b)
Application of Otsu’s threshold, which does not capture the defect. (c) Successful segmentation
of the killer defect using the TTBH.
30 µm 30 µm
SEM-SE Specimen axis
c d
Otsu (126/255) TTBH (198/255)
Defect
not connected
to surface
Manual removal
of the ligament
Figure 5.18 | Comparison of the abilities of Otsu’s threshold and TTBH to capture
the killer defect on the 45°|Vertical
40min -2 specimen. The specimen failed after 1 173 759 cycles
at 550 MPa, and the killer defect is a sub-surface pore that was brought just beneath the surface
by the PeP treatment. (a) SEM-SE observation of the initiation site on the fracture surface. (b)
Transverse XCT slice (voxel size = 3 µm). (c) Application of Otsu’s threshold, which enables
capturing some part of the defect only. Since the defect is not connected to the surface, it would
not be considered a surface defect. (d) Successful segmentation of the defect using the TTBH.
The small ligament that separated the pore from the surface was removed by the TTBH. (e)
An alternative solution to take into account sub-surface pores in the present workflow dedicated
to surface defects analysis. The ligament that separates the pore from the surface is artificially
removed, revealing the underlying defect even when using Otsu’s threshold.
a Initiation site b
XCT
10 µm 30 µm
Specimen axis
c d
Otsu (124/255) TTBH (198/255)
a b
Initiation site XCT
SEM-SE
Specimen axis
10 µm 30 µm
c d
Otsu (124/255) TTBH (199/255)
Figure 5.20 | Unsuccessful segmentation of the killer defect on the 45°|45° 40min -3 spec-
6
imen due to a lack of resolution. The specimen failed after 2 ˆ 10 cycles at 550 MPa
followed by 52 369 cycles at 650 MPa. The killer defect is a subsurface pore that was brought to
the surface by PeP, although its connection to the surface is very small. (a) SEM-SE observa-
tion of the initiation site on the fracture surface. (b) Transverse XCT slice (voxel size = 3 µm).
(c) Application of Otsu’s threshold, which entirely misses the killer defect. (d) Application of
the TTBH which enables capturing some part of the defect only, and fails at establishing its
connection with the surface.
The SEM observation in Fig.5.20a shows that the killer defect is a pore that is barely con-
nected to the surface. Moreover, the portion of the defect close to the surface seems very thin,
making its characterization by XCT very difficult. As a result, Fig.5.20b shows that only a
limited portion of the defect can be detected in the XCT scan. It corresponds to the bottom of
the defect, which is larger. This portion of the defect is successfully segmented using the TTBH
threshold, see Fig.5.20d. However, most of the defect is not captured, and in particular its con-
nection to the surface is not accounted for. Thus, the methodology for killer defect prediction
could not be applied.
Probability density
(surface points Kt*) Z|as-built
1.0
0.5
0 0
Probability density
(defects max Kt*)
0.5
1 2 3 4
Figure 5.21 | Distributions of surface points and defects Kt˚ values for as-built Z-7
specimen. The blue line shows the PDF for Kt˚ values of surface points, considering only
points with negative heights and κσ . The orange points and curve represent the maximum Kt˚
value for each segmented defect and its PDF. The killer defect is marked in purple. PDFs are
calculated using the KDE method with Silverman’s rule of thumb for bandwidth selection.
Z Rank Percentage
Height κσ Kt˚ Kt˚
λc “ 800 µm λc “ 800 µm
Specimen ndef ects λc “ 800 µm rcurv “ 50 µm
rcurv “ 20 µm rcurv “ 20 µm
Z-1 725 323 234 215 51
Z-2 654 287 140 200 55
Z-3 710 181 198 187 67
Z-4 732 316 217 271 59
Z-5 752 26 11 1 100
Z-6 719 318 200 242 71
Z-7 739 301 87 176 67
Z-8 709 49 20 5 93
Z-9 730 53 31 28 82
Z-10 761 114 60 57 84
Median 728 234 114 182 69
Table 5.3 | Killer defect prediction scores (rank & percentage) for as-built Z speci-
mens using roughness, κσ and Kt˚ parameters. The Specimen column refers to the name
of individual specimens and the ndef ects column gives the total number of defects that were au-
tomatically segmented. Colors are given for guidance, the color tending to green for a successful
killer defect prediction while it tends to red for a failed prediction.
First, it can be noticed that the number of segmented defects is much higher for
as-built specimens (« 700) in comparison with specimens polished by PeP (ă 50). This is,
of course, linked to the much higher density of surface defects in as-built specimens.
A second observation is that ranks are much higher for as-built Z specimens compared to those
polished by PeP. Thus, for most as-built Z specimens, the method clearly fails to properly
predict the killer defect. In this context, it seems hazardous to try determining which
parameter among the roughness, κσ , and Kt˚ is the most relevant.
Nonetheless, results for as-built 45° specimens are much more encouraging than
those for as-built Z specimens. They are shown in the form of PDFs for one specimen in
Fig.5.22, followed by a more comprehensive analysis for all specimens in Tab.5.4.
As-built 45° specimens consistently achieve much lower ranks compared to as-built Z spec-
imens. For instance, the median rank for the Kt˚ parameter across 10 analyzed 45° specimens
is 21. The relatively better results obtained for the 45° specimens might be because defects in
45° specimens are larger. This may facilitate their identification using XCT with the resolution
employed here (voxel = 3 µm). The difference between Z and 45° specimens might also arise due
to the very different valley morphologies in either case.
1.0
0.5
0 0
Probability density
(defects max Kt*)
0.5
1 2 3 4 5
Figure 5.22 | Distributions of surface points and defects Kt˚ values for as-built 45°-7
specimen. The blue line shows the PDF for Kt˚ values of surface points, considering only
points with negative heights and κσ . The orange points and curve represent the maximum Kt˚
value for each segmented defect and its PDF. The killer defect is marked in purple. PDFs are
calculated using the KDE method with Silverman’s rule of thumb for bandwidth selection.
As for PeP specimens, it is also worth noting that although very high ranks are obtained for
some specimens, the percentage score can remain reasonably high. For instance, a rank
of 242 is obtained but a relatively high percentage of 71% is found for the Z-6 specimen.
Given all these observations, the following points might at least partially explain the inability
of the model to provide an accurate prediction of the killer defects in as-built specimens:
• The resolution of XCT scans is insufficient to properly capture surface defects.
• Due to the higher density of defects in as-built specimens, it seems logical that it is more
challenging to find the killer defect among «700 segmented defects than it was to find it
among «30 defects in the case of specimens polished by PeP.
• Other factors such as the microstructure may play a role and promote crack initiation at
spots where the stress concentration is not maximal.
• The model does not take into account the defects’ morphology. Yet, the particularly
elongated shape of surface valleys may play a significant role in crack initiation and/or
propagation. Properly taking into account the defects’ shape would probably require
improving the method used for defect segmentation together with the resolution of XCT
scans.
Table 5.4 | Killer defect prediction scores (rank & percentage) for as-built 45°
specimens using roughness, κσ and Kt˚ parameters. The Specimen column refers to the
name of individual specimens and the ndef ects column gives the total number of defects that
automatically were segmented. Colors are given for guidance, the color tending to green for a
successful killer defect prediction while it tends to red for a failed prediction.
κmin (µm-1)
< -0.09 -0.05 0 > 0.06
BD
60 µm
30 µm
Qualitatively, roughness parameters such as Sa and Sv , along with the maximum estimated
˚
stress concentration factor Kt,max , exhibited some correlation with the fatigue resistance
of as-built and PeP-polished specimens. In terms of the fatigue life reached at a given
maximum stress level, these parameters effectively differentiate between significantly
different surface states (like as-built vs PeP). However, they fall short in explaining the
variations in fatigue life among specimens with similar surface states, such as between
different as-built Z specimens. Surprisingly, a better correlation was still found for the Sa
parameter. Conversely, at a fixed fatigue lifetime, the maximum stress that specimens can
endure appears to correlate better with the parameters tested, particularly Kt˚ .
A perspective to improve these results for as-built specimens would be to take into account
the length of the surface valleys. Indeed, a correlation was also found between the fatigue
resistance of as-built specimens and the length of the initiation zone measured on the
fracture surface observed by SEM. The latter is related to the length of the valley at the
origin of fatigue failure.
The results obtained for the prediction of the killer defects using the Kt˚ parameter calculated
from XCT scans were encouraging for specimens polished by PeP. For 4 specimens out of the
8 where the killer defect could be identified on the XCT scan before failure, the killer defect
could be successfully predicted. For the 4 other specimens, the killer defect was among the
10 defects considered to be the most harmful based on their Kt˚ value. However, due to
the lack of resolution of XCT scans, the killer defect could not be identified on the surface
extracted from XCT scans for 6 specimens (out of the 14 that were analyzed).
The killer defect predictions have proved to be a far more complex task for specimens
with an as-built surface. Defects have much more complex geometries than in the case
of PeP specimens, making accurate segmentation more difficult. The high density of defects
(«700/specimen compared to «30/specimen for PeP specimens) also contributes to making
the task more difficult.
Hence, the killer defect predictions clearly failed for as-built Z specimens. The median rank
of the killer defect is thus 182 when using the Kt˚ parameter. Even though the killer defect
could not be successfully predicted in most cases, the results are much better for as-built
45° specimens. This could be due to the larger depth and different shape of valleys in 45°
specimens.
Still, further improvements could be achieved, for instance by taking into account the length
of the surface valleys. This would require improvements in the method used for the surface
defects segmentation, which has been done here essentially by thresholding surface points
based on the local curvature. Also, if higher resolutions are difficult to obtain, it would be
interesting to apply this methodology to specimens with larger surface defects which would
be better captured by XCT at an equivalent resolution. This could, for instance, be the case
by studying as-built samples manufactured by other AM processes such as E-PBF or Wire
Arc Additive Manufacturing (WAAM).
Conclusion
This work aimed at better understanding the impact of defects on fatigue properties of the
Ti64 alloy manufactured by Laser Powder Bed Fusion (L-PBF). The extensive literature on this
subject reveals that the large surface roughness inherited from this process significantly reduces
the fatigue properties. This roughness is caused by surface defects such as valleys that generate
stress concentrations and promote crack initiation under cyclic loading. Characterizing these
defects is challenging because additively manufactured surfaces can show complex 3D shapes
and topography. Traditional surface characterization methods often fall short in these scenarios,
requiring new developments.
Material characterization
Chapter 3 describes the samples used in this work. These are Ti64 mechanical specimens
manufactured by L-PBF using several build orientations: vertical (Z), at 45°, and horizontal
(for tensile specimens only). Samples were stress relieved for 2 h at 720 °C to remove all residual
stresses. After stress relief, the microstructure is composed of a mixture of fine α laths (with an
average thickness of 0.5 µm) along with a small fraction of β phase (2.2 ˘ 0.3%).
It results in a rather high yield strength (Y¯S “ 1098 MPa for Z specimens) but a rather
low elongation to failure (7.8% for Z specimens, and only 0.8% for horizontal specimens). The
strong anisotropy in elongation to failure seems to be partly due to the preferential failure at
prior β0 columnar grain boundaries.
The defects characterization showed that samples have a high density (ą 99.998%) and
that residual pores are both small (over 80% smaller than 30 µm) and have a high sphericity.
Most pores are located in a sub-surface layer at approximately 140 µm from the surface, which
corresponds to the boundary between contour laser scans and the bulk material (hatching).
Several surface defects could be identified on the 3D surface characterization, the defects that
are expected to cause the most damage in fatigue being valleys. Valleys were found to be deeper
and more tortuous in down-skin areas of specimens built at a 45° angle with respect to the built
plate, in comparison with vertically built specimens.
Perspectives
This work opens various avenues for future developments and a deeper understanding of
the impact of surface defects on fatigue properties. Below are some key areas for potential
improvements.
Figure 5.25 | Surface roughness and curvature characterization from 2.5D data. (a)
2.5D roughness measurement obtained using a chromatic confocal sensor (HiroX NPS) and a
lateral spacing for measurement of 1 µm. (b) Minimum curvature map, computed using rcurv “
10 µm. (c) Superposition of (a) and (b).
Figure 5.27 | Sharp features extraction using the approach of [COE 16], based on
normal vector field regularization. Sharp edges are successfully segmented in red, even
when some noise is added to the data (see the lower right half of the image).
Figure 5.28 | Comparing the killer defect prediction to the actual crack location
after failure in an IN718 bending fatigue specimen manufactured by L-PBF. The
surface subjected to fatigue loading is the down-skin part of a specimen built a 45°. The
characterization was done using XCT scans with a voxel size of 4 µm. (a) Enlarged view of the
Kt˚ map of the defect at point A, belonging to the crack path. Point B is the only location of
the surface where Kt˚ values superior to those in point A were found. (b) Internal view of the
surface Kt˚ map before fatigue testing. (c) Internal view of the surface Kt˚ map after fatigue
testing. The propagated crack is characterized by high Kt˚ values and can thus be seen in red.
(d) XCT slice revealing the notch defect at point A.
Measuring roughness on free-form surfaces involves calculating the distance between the
original rough surface and a reference surface. In this study, the original surface is defined by
what we call the sample mask or surface mask, as introduced in Chapter 2. The reference surface
is identified as the smooth mask in the same chapter. Additionally, we can introduce here the
concept of the smooth surface mask, which is to the smooth mask what the surface mask is to
the sample mask.
Different methods can calculate this distance, and while they often yield similar results, some
may introduce more errors, especially in specific scenarios. This appendix focuses on comparing
two distinct methods.
The method presented in Chapter 2 involves the use of a Signed Euclidean Distance Transform
(SEDT). Its advantage is that it is a 3D operation, directly available on software such as ImageJ.
Alternatively, a Nearest Neighbor Search (NNS) technique can be more appropriate in certain
cases. NNS seeks the closest point(s) in a point cloud for a given point. In our context, it can be
used to identify the nearest point on the smooth surface for each point on the (rough) surface.
The local height, or roughness, is then the distance between these two points. Practically, NNS
is often executed using a kd-tree data structure [SKR 19], as used in this study. In terms of
computing time, the NNS approach is similar to SEDT, but better than SEDT in terms of
memory usage. However, unlike SEDT or Gaussian filters, it is not a volumetric operation and
is not readily available in software like ImageJ without custom implementation. In this work,
the NNS computation was done using the scipy-spatial Python library. The SEDT method was
also conducted using Python, utilizing the edt library, which offers an efficient multi-threaded
implementation. Note that the Python implementation for 3D roughness measurement provided
in [STE 23] proposes the two methods for distance calculation. In the script, the SEDT method
is referred to as the ’edt’ method, while the NNS method is referred to as the ’kdtree’ method.
Fig.A.1a-e presents a comparison of roughness measurements obtained using SEDT and NNS
methods on a straightforward example: a perfect cylinder with a radius of 3.5 voxels. In this
scenario, both the sample mask and the smooth mask are identical, as they represent the same
perfect cylinder (shown in Fig.A.1a). Similarly, the surface mask and the smooth surface mask
are also the same (see Fig.A.1b).
| 197
Theoretical
sample shape
c d e
Theoretical result SEDT method NNS method
2.8 2.2 1.4 1 1 1 1.4 2.2 2.8
In this specific case, the error can be corrected by subtracting the mean height from all surface
voxels post-calculation, a process that will be referred to as zero-centering. This correction step
is recommended in Chapter 2, notably for situations like this.
For a more in-depth comparison, we used a cylinder with an artificial sinusoidal roughness
on its surface. The cylinder is 4 mm in height and has an average diameter of 1 mm, with the
sinusoidal roughness featuring a wavelength of λ “ 100 µm and amplitude A varying from 0 to
30 µm. Figure A.2 provides a schematic representation of a 2D segment of this surface. The
sample mask, used for calculations, is generated from the analytical formula of this shape, using
a voxel size of 2.5 µm.
λ = 100 µm A ∈ [0,30] µm
Bulk material
Figure A.2 | Illustration of a section of the sinusoidal roughness used for the example.
| 199
These findings suggest that while the choice between SEDT and NNS methods may not be
critical for surfaces with significant roughness, it becomes crucial for smoother surfaces. This is
for example the case for PeP-polished samples from Chapter 4, where the SEDT method was
found to cause substantial errors in some roughness measurements.
The error from using SEDT, as opposed to a method like NNS, will typically be around
the voxel size. While this may not seem substantial for single voxel height measurements or for
calculating roughness parameters like Sz or Sv , it can significantly affect average parameters such
as Sa . This is because a consistent 2.5 µm negative error (as seen in Fig.A.1d) will substantially
skew the Sa value.
2.5
Mean error (µm)
2.0
1.5
1.0
0.5
0.0 Theory
0 5 10 15 20 25 30 EDT
A (µm) EDT | Zero-centered
b
NNS
NNS | Zero-centered
15
Sa (µm)
10
0
0 5 10 15 20 25 30
A (µm)
Figure A.4 | Comparing roughness measurements from SEDT and NNS methods
with theoretical calculations. (a) Theoretical results with analytically calculated roughness
values. (b) Roughness measurements using the SEDT method. (c) Measurements from the NNS
method.
This effect is visible in Fig.A.4b, where using the SEDT method without zero-centering
results in Sa values markedly higher than theoretical ones, particularly at small A values. This
discrepancy persists even at larger A values. In contrast, the NNS method delivers Sa values
very close to the theoretical results.
In summary, the NNS method is preferable to SEDT where possible. If SEDT must be used,
the zero-centering step can reduce errors. However, caution is advised when characterizing very
smooth surfaces, particularly if Sa values are close to the voxel size.
| 201
The following methodology can be used to apply any 3D linear filter on a set of voxels. In
particular, it can be used to apply a Gaussian S-filter (or L-filter) to roughness values calcu-
lated following the workflow described in Section 2.2.2. It can be considered as a particular
application case of the concept of normalized convolution introduced by [KNU 93]. Unlike a
conventional convolution where all voxels are taken into account, normalized convolution can be
used to ignore certain voxels. Here, it will be used to ignore background voxels and thus filter
only the ones that carry actual information − i.e. the surface voxels that carry roughness values.
A linear filter is an operation where the value at a given voxel is replaced by a linear combi-
nation of the value at the given point and its neighbors. Each neighbor has a specific weight w,
i.e. the weight of its contribution to the final filtered value. The matrix assigning the weights
for the central voxel and its neighbors is called the filter kernel. For a uniform (= mean) filter,
all weights in the kernel will have the same weight. In the case of a Gaussian filter, weights
decrease as the distance to the central voxel increases following a Gaussian law, see Fig.B.1. In
both cases, weights are generally normalized, meaning the sum of all kernel weights equals 1.
For example, this ensures that applying a Gaussian filter to a uniform volume made of 1 will
result in a uniform volume made of 1. This is the standard "global" normalization illustrated in
Fig.B.1 and employed in common Gaussian filter implementations.
However, the situation is a bit different when applying a S-filter to a set of surface voxels. In
such a case, to compute the filtered value at a given voxel, only neighbor surface voxels should be
taken into account instead of all neighbor voxels. The solution would be to explicitly loop over
surface voxels only, and never visit background voxels. This can be done in programming lan-
guages such as C++. However, this still requires some programming skills, especially to achieve
reasonable computation times. Conversely, there are already many efficient implementations of
the standard Gaussian filter, including ones accelerated via GPU of FFT computations. The
idea here is thus to use such optimized implementations ingeniously to compute indirectly the
S-filter.
| 203
1 2 1
Discrete Gaussian kernel 2 4 2 Sum = 16
without normalization
1 2 1
"Local" normalization
Only surface voxels are taken into account
1 2
Apply standard Gaussian filter Apply standard Gaussian filter
to roughness to normalization volume
Pixel being
processed 0 1 0 0 0 0
0 6.7 0 0 0 0
0 8.9 0 0 0 0 0 1 0 0 0 0
0 0 8.6 7.9 0 0
Surface 0 0 1 1 0 0
pixel
0 0 0 0 7.2 0 0 0 0 0 1 0
0 0 0 0 8.1 8.9 0 0 0 0 1 1
Background
0 0 0 0 0 0 pixel 0 0 0 0 0 0
Figure B.1 Workflow proposed for the application of a 3D Gaussian S-filter. All calculations are
done using operations on digital volumes, just as the rest of the roughness computation workflow.
The example provided in the present figure is restricted to 2D, which means voxels are replaced
by pixels. This choice is made to facilitate understanding and visualization, considering the
adaptation to a 3D case is straightforward. Note that in steps 1 and 2 of the "local normalization"
in Fig.B.1, a non-normalized Gaussian filter is used for the sake of clarity only. The same steps
can therefore be followed using a conventional Gaussian filter.
The resulting value at each point is the sum of all the Gaussian kernel weights that effectively
come across surface voxels. It can be considered as a "local normalization factor", since by
dividing the filtered value obtained in step 1 by this value, we obtain a properly normalized
S-filter, see step 3 in Fig.B.1. This method is equivalent to use at each voxel a specific kernel.
The latter contains 0 at background voxels, and at other voxels, a weight decreasing according to
a Gaussian law with the distance to the central voxel. To be normalized, the sum of all weights
must be equal to one. The trick here is to achieve such a (non-linear) computation using only
linear filtration steps − i.e. using the same unique kernel for all voxels.
| 205
The following figures show EBSD characterizations on vertical specimens before and after
PeP treatment. Average α-laths thickness for the as-built (Z), Z|Vertical Vertical specimens
40min and Z|1h20
characterized were measured to be 0.609 µm, 0.593 µm and 0.594 µm respectively.
Acquisition and analysis were done by Quentin Gaillard. For a more detailed analysis of the
near-surface microstructure, see Section 3.2.2 or the PhD thesis of Quentin Gaillard [GAI 23a].
| 207
| 209
References | 211
212 | REFERENCES
REFERENCES | 213
214 | REFERENCES
REFERENCES | 215
216 | REFERENCES
REFERENCES | 217
218 | REFERENCES
REFERENCES | 219
220 | REFERENCES
REFERENCES | 221
222 | REFERENCES
REFERENCES | 223
224 | REFERENCES
REFERENCES | 225
226 | REFERENCES
REFERENCES | 227
228 | REFERENCES
RÉSUMÉ :
Manufacturing defects are known to significantly impact the fatigue properties of additively manufactured (AM)
components. Notably, the large surface roughness, typical of AM processes, leads to increased stress concentrations
that promote crack initiation, thereby reducing fatigue resistance.
Traditionally, surface roughness and related defects are evaluated using tools like white light interferometers. How-
ever, these instruments offer a limited, single-perspective and only partially three-dimensional analysis. Those limita-
tions do not enable the thorough characterization of the complex surfaces and hidden defects typical in AM components.
This study describes a methodology for performing a 3D surface analysis using X-ray Computed Tomography
(XCT) data. The method is illustrated on various samples, ranging from simple cylinders to more intricate architected
structures. It turns out to be very efficient at detecting critical surface defects, such as notches hidden by partially
melted powder particles.
The methodology is then applied to examine the effect of surface defects on the fatigue properties of Ti64 pro-
duced by Laser Powder Bed Fusion (L-PBF). This analysis includes both as-built surfaces and those subjected to
post-treatments, specifically investigating the impact of Plasma electrolytic Polishing (PeP) and surface oxygen con-
tamination (presence of an α-case layer) resulting from high-temperature heat treatment (860 °C).
Using XCT for 3D characterization, defects responsible for fatigue failure are identified, the latter being predom-
inantly surface valleys. The method’s ability to predict crack initiation locations is also evaluated, as well as its
potential to estimate the fatigue resistance of a specimen before testing.
MOTS-CLÉS: 3D surface analysis, Fatigue, Additive Manufacturing, X-ray Computed Tomography, Ti64,
Plasma electolytic Polishing, Alpha-case.
Laboratoires de recherche : MatéIS − UMR CNRS 5510 | SIMaP − UMR CNRS 5266