Metals 11 00731

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

metals

Article
Influence of Grain Orientation Distribution on the High
Temperature Fatigue Behaviour of Notched Specimen Made of
Polycrystalline Nickel-Base Superalloy
Benedikt Engel 1, *, Sebastian Ohneseit 2 , Lucas Mäde 3 and Tilmann Beck 4

1 Gasturbine and Transmission Research Center (G2TRC), University of Nottingham,


Nottingham NG7 2RD, UK
2 Institute of Applied Materials—Applied Materials Physics, Karlsruhe Institute of Technology,
76344 Eggenstein-Leopoldshafen, Germany; sebastian.ohneseit@kit.edu
3 Gas and Power Division, Department for Technology & Innovation, Siemens AG, 10553 Berlin, Germany;
lucas.maede@siemens.com
4 Institute of Materials Science and Engineering, University of Kaiserslautern, 67655 Kaiserslautern, Germany;
beck@mv.uni-kl.de
* Correspondence: benedikt.engel@nottingham.ac.uk

Abstract: Two different material batches made of random and textured orientated polycrystalline
nickel-base superalloy René80 were investigated under isothermal low cycle fatigue tests at 850 ◦ C
for a notched specimen geometry. In contrast to a smooth specimen geometry, no significant improve-
ment in fatigue behaviour of the notched specimen could be observed for the textured material. Finite
element simulations reveal an area along the notch where high stiffness evolves for the textured

 material, which lead to nearly similar shear stresses in the slip systems compared to a random
Citation: Engel, B.; Ohneseit, S.;
orientation distribution and therefore to no distinct differences in the lifetime.
Mäde, L.; Beck, T. Influence of Grain
Orientation Distribution on the High Keywords: nickel-base superalloy; LCF; grain orientation distribution; texture; resulting shear stress;
Temperature Fatigue Behaviour of polycrystalline finite element simulation; high temperature
Notched Specimen Made of
Polycrystalline Nickel-Base
Superalloy. Metals 2021, 11, 731.
https://doi.org/10.3390/met11050731 1. Introduction
The worldwide demand for the reduction of CO2 emissions during power generation
Academic Editor: Stefano Spigarelli
and the associated increased usage of renewable energy sources has lead to fluctuation
in the power supply and can cause problems with regard to power grid stability. In
Received: 12 April 2021
particular, the generation of wind and solar power is strongly dependent on the current
Accepted: 26 April 2021
weather conditions, which result in volatile energy input into the power grid. In order
Published: 29 April 2021
to fill in these gaps, gas power plants are used due to their ability to generate power in
short times. Modern gas turbines, as a main part of gas power plants, can be started from
Publisher’s Note: MDPI stays neutral
cold to maximum power output in less than 30 min [1]. Due to frequent starts, stops and
with regard to jurisdictional claims in
published maps and institutional affil-
load changes from part to full load, the materials in the hot gas section and especially
iations.
the turbine blades have to withstand significantly fluctuating mechanical and thermal
loads. Hence, thermo–mechanical fatigue, high-temperature fatigue behavior and creep
effects of the turbine material must be investigated in order to reach high efficiencies with
simultaneous economic design aspects. Since component testing is very complex and
expensive, generally, a standardized specimen of the used material is examined under
Copyright: © 2021 by the authors.
high-temperature fatigue conditions. To represent uniaxial stress states, specimens with
Licensee MDPI, Basel, Switzerland.
cylindrical gauge sections are used; moreover, notched specimens are used to determine the
This article is an open access article
distributed under the terms and
fatigue behavior of the material under multiaxial stress states as well as stress gradients to
conditions of the Creative Commons
represent complex component structures. Besides experimental material testing, numerical
Attribution (CC BY) license (https://
models are indispensable during the design process in order to predict the deformation and
creativecommons.org/licenses/by/ fatigue behavior under various load and temperature conditions. Based on experimental
4.0/). observed and measured mechanical material responses, models such as the commonly

Metals 2021, 11, 731. https://doi.org/10.3390/met11050731 https://www.mdpi.com/journal/metals


Metals 2021, 11, 731 2 of 21

used Ramberg–Osgood or the Chaboche model are used to describe the nonlinear stress
strain relationship for materials under high-temperature conditions [2–4].
Polycrystalline nickel-base superalloys, such as René80, are used as blades or vanes
in the rear hot gas sections of gas turbines with maximum operating temperatures up
to 982 ◦ C [5,6]. Due to the conventional vacuum casting process of the turbine blades,
different grains sizes and orientation distributions arise depending on the component
geometry and cooling conditions. Coarse grains and low grain numbers located in highly
stressed component areas in combination with high elastic anisotropies of nickel-base
superalloys lead to large scattering in fatigue lifetimes and mechanical properties [7–9],
which lead to high safety factors for deterministic design approaches. In order to achieve
higher efficiencies for the components, these conservative deterministic lifetime models are
successfully replaced by probabilistic approaches [10–13]. For this purpose, the influencing
factors and their statistical distributions on high-temperature fatigue behavior must be
identified and investigated for the material. Seibel et al. [11,14] investigated the fatigue
behavior of conventionally casted polycrystalline nickel-base superalloy René80 with an
assumed random grain orientation distribution under low cycle fatigue (LCF) conditions
at 850 ◦ C. For the same total strains, notched specimen shows a distinct increased fatigue
lifetime compared to the smooth specimen. This can be mainly attributed to the statistical
size effect, where the notched specimen has a significantly reduced highly loaded volume
and therefore the probability of grains which tend to crack initiation is decreased. In
addition, the formation of stress gradients, caused by the notch geometry, lead to decreased
stresses into the bulk material, which could result in lower crack growth rates and therefore
to longer lifetimes compared to cylindrical specimens. For different notch geometries, an
increase in lifetime with an increase of the notch factor could be shown. A similar behavior
could also be proven for different material such as steels [15].
Numerical evaluations in [16], considering the influence of grain orientation on the
mechanical behavior of nickel-base superalloys, reveal a relationship between the local
grain orientation and the resulting shear stress within the slip system, responsible for
plastic deformation. Due to high elastic anisotropies of nickel-base superalloys, the highest
shear stresses for the <111>{110} slip systems were achieved for grain orientations with
corresponding Young’s moduli between 180 GPa and 230 GPa (at 850 ◦ C), and Schmid
factors between 0.38 and 0.46 for uniaxial loading cases. Orientations with a maximum
in Schmid factors of 0.5 show distinct lower shear stresses, due to low correlated elastic
properties at around 120 GPa.
Engel, Mäde et al. showed in [17] the influence of the local grain orientation on the
probability of fatigue failure for smooth cylindrical specimen made of polycrystalline
René80 with a random and textured grain orientation distribution. By determining that
the grain orientation depended on the resulting shear stress within the slip systems of
the material by polycrystalline finite element simulation, reliable probabilistic lifetime
predictions could be made for randomly orientated material as well as textured material.
As a fatigue life describing parameter, a modified Schmid factor approach was introduced.
For the nodes of a polycrystalline finite element simulation, the modified Schmid factor is
defined as the quotient of von Mises stress and the resulting shear stress in the activated slip
systems. It could be shown that the determined texture of the investigated material lead to
a lower Young’s moduli for the cylindrical specimen and therefore to lower stresses during
uniaxial strain controlled tests compared to the randomly orientated material. Furthermore,
the texture influences the orientation of the slip system and lead to lower modified Schmid
factors. Thus follows the requirement of higher local stresses in the slip system which
results in a delay in crack initiation whereby higher fatigue lifetimes were achieved.
In addition to the results of different smooth-specimen grain orientation distributions
in [17,18], the following paper investigates the high-temperature LCF behavior of the
same material batch of nickel-base superalloy René80 for a notched specimen geometry.
Isothermal LCF tests at 850 ◦ C were carried out for a random and a textured grain ori-
entation distribution, and the lifetimes compared to the smooth specimen from previous
Metals 2021, 11, x FOR PEER REVIEW 3

Metals 2021, 11, 731 3 of 21

same material batch of nickel-base superalloy René80 for a notched specimen geome
Isothermal LCF tests at 850 °C were carried out for a random and a textured grain or
investigations.tation distribution,
To explain and the lifetimes
the observations compared
and differences in to
thethe smoothasspecimen
lifetimes well as thefrom prev
variation in mechanical behavior, polycrystalline finite element simulations for notchedas well as
investigations. To explain the observations and differences in the lifetimes
geometries werevariation
carried in
outmechanical
to analyze behavior, polycrystalline
the influence finite
of local grain element on
orientation simulations
the local for notc
geometries were carried out to analyze the influence
mechanical behavior during cyclic high-temperature fatigue tests. of local grain orientation on the l
mechanical behavior during cyclic high-temperature fatigue tests.
2. Materials and Methods
2. Materials
The composition of theand Methods
polycrystalline nickel-base superalloy René80 used in this
The
work is shown in the composition
following Tableof the polycrystalline
1 and was measured by nickel-base superalloy René80 used in
the manufacturer.
work is shown in the following Table 1 and was measured by the manufacturer.
Table 1. Composition of René80 in wt.%.
Table 1. Composition of René80 in wt.%.
Element Ni Cr Co Ti Mo W Al C B Zr
Element Ni Cr Co Ti Mo W Al C B Z
René80 Bal. 14.04 9.48 5.08 4.03 4.02 2.93 0.17 0.015 0.011
René80 Bal. 14.04 9.48 5.08 4.03 4.02 2.93 0.17 0.015 0.0

The material wasThevacuum


material casted from the
was vacuum same
casted meltthe
from and
sameheatmelt
treated
and after [19] at after [19
heat treated
Doncaster in Bochum/Germany
Doncaster in Bochum/Germany as bars of 150 mm in length with 20 mm in
as bars of 150 mm in length with 20 mm and 12 mm and 12 mm
diameter. The different
diameter.diameters of the
The different bars lead
diameters of to
thedifferent
bars leadcooling conditions
to different coolingand tem- and t
conditions
perature transients during
perature solidification
transients duringin the radial direction
solidification (x−direction,
in the radial see Figure 1).see Figur
direction (x−direction,
In addition, theIn
casting process
addition, as a standing
the casting processmould leads to mould
as a standing a further temperature
leads to a furthergradient
temperature gr
in the axial direction (y–direction).
ent in the Due(y–direction).
axial direction to the faster cooling
Due to theconditions in theconditions
faster cooling 12 mm bar,in the 12
grain growth isbar,
inhibited, which leads
grain growth to smaller
is inhibited, elongated
which leads tograins
smallercompared
elongated to grains
the 20 mm
compared to
bar, as Figure 120
shows.
mm bar, as Figure 1 shows. The white dashed line represents the smooth
The white dashed line represents the gauge section of the gauge section of
specimen, and smooth
the green dashed line
specimen, and the
the gauge section line
green dashed of the
thenotched specimen.
gauge section of the notched specim

(a) (b)

x
5 mm 3 mm

Figure 1.
Figure 1. Longitudinal cutLongitudinal
of the 20 mmcutbar
of (a)
theand
20 mm bar (a)
12 mm barand
(b).12 mm bar (b).

The preferred Thegrowth of the growth


preferred [100]–orientation towards the temperature
of the [100]–orientation gradients gradi
towards the temperature
known for nickel-based
known forsuperalloys
nickel-based leads to dendritic
superalloys leadssolidification along the edge
to dendritic solidification areas
along the edge a
of the bars [20]ofrepresented
the bars [20]asrepresented
blue dotted as lines.
blue dotted
Duringlines. During manufacturing
manufacturing of the sampleof the sample
geometries, thegeometries,
dendritic edgethe dendritic edge area
area is almost is almost
completely completely
removed from removed
the gauge from the gauge sec
section
of the 20 mm rod.of the 20 mm rod.
However, the However, the samples,
samples, which which were manufactured
were manufactured from the 12 mm fromrod,
the 12 mm
still show
still show dendritic dendritic
structures structures
within the gaugewithin the gauge
section, as seensection, as seen
in Figure 1. Asin described
Figure 1. As descri
in [17],solidification
in [17], the dendritic the dendritic andsolidification and the
the resulting resultingdirection
preferential preferential direction
of the grains,of the gra
with an average with an average
alignment = 25◦ , result
of ϑalignment of ϑ =in25°, result inmaterial
a different a different materialcompared
behavior behavior compare
to the specimen thefrom
specimen
the 20from
mm the
bar 20 mm bar(with
material material
random(withorientation).
random orientation).
To determineTo determine
grain size, the grains in the gauge section were evaluated by
the grain size, the grains in the gauge section were evaluated by the equivalent diameter the equivalent diameter u
a light microscope
using a light microscope and anprocessing
and an image image processing
software.software. Theof
The ratio ratio
theof the grain
grain sizes sizes wi
within the gauge thesection
gauge section of the specimen
of the specimen between between
the 12 mmthe 12 mm
bar andbar
20and
mm20bar mm bar is about 1:3
is about
1:3. In the following, the specimen taken from the middle of the 20 mm bar will be called
coarse-grained and consist of a random orientation distribution of the grains, since no
obvious crystallographic texture could be determined using electron backscatter diffraction.
Metals 2021, 11, x FOR PEER REVIEW 4 of 22

Metals 2021, 11, 731 4 of 21


the following, the specimen taken from the middle of the 20 mm bar will be called coarse-
grained and consist of a random orientation distribution of the grains, since no obvious
crystallographic texture could be determined using electron backscatter diffraction. The
The specimens
specimens madethe
made from from
12 the
mm12 mm
bar arebar are called
called fine-grain
fine-grain textured
textured material,
material, due todue
the to
the preferred
preferred solidification
solidification in direction
in direction of the resulting
of the resulting temperature
temperature gradient.
gradient. The René80
The René80 ma-
material
terial from
from the the isothermal
isothermal LCFresults
LCF tests tests results taken[14]
taken from from
for[14] for comparison
comparison was alsowas also
man-
ufactured and heat-treated by Doncaster but casted as solid plate with a thickness of 20 of
manufactured and heat-treated by Doncaster but casted as solid plate with a thickness
mm.20 The
mm.grain
The grain
sizes sizes
are inare in a good
a very very good accordance
accordance to thetocoarse-grain
the coarse-grain material
material as as
as well well
theasgrain
the grain orientation
orientation distribution
distribution waswasalsoalso described
described as random.
as random. In order
In order to compare
to compare thethe
results
results to to [14],
[14], thethe same
same notch
notch geometry
geometry as as well
well as as cylindrical
cylindrical gauge
gauge section
section were
were chosen
chosen
andandareare shown
shown in in Figure
Figure 2. 2.
TheThe notch
notch factor
factor waswas calculated
calculated to to a value
a value of of 1.62.
1.62.

Figure 2. Geometry of the smooth (a) and notched (b) specimen in accordance to [14,21].
Figure 2. Geometry of the smooth (a) and notched (b) specimen in accordance to [14,21].
Isothermal LCF tests were carried out at 850 ◦ C at a servo hydraulic MTS test rig with
Isothermal
a maximum load LCF capacity
tests wereofcarried
100 kNout at 850 °Cwith
combined at a servo hydraulicTrueHeat
an Hüttinger MTS test (TRUMPF
rig with
a maximum
Hüttinger GmbH,load capacity
Germany, of 100 kN combined
Freiburg) with an Hüttinger
MF 5000 induction TrueHeat
heating system. (TRUMPF
Strain-controlled
Hüttinger GmbH, Germany, Freiburg) MF 5000 induction heating
tests were performed at R = −1, using an MTS high-temperature extensometer type 632.53system. Strain-con-
trolled
with tests wererods,
ceramic performed at R = −1, using
a measurement lengthanofMTS
12 mmhigh-temperature
and a constantextensometer
air flow coolingtype to
632.53 with ceramic rods, a measurement length of 12 mm and a constant
prevent temperature influences on the measurement. All tests were carried out with a test air flow cooling
to frequency
prevent temperature influences
of 1 Hz. Details to theon theconducted
tests measurement. withAll
thetests
smoothweresamples
carriedcanout be
with a
found
testinfrequency
[17,21]. of 1 Hz. Details to the tests conducted with the smooth samples can be
found in [17,21].
Since the used type K ribbon thermocouple could not be applied in the notch due
Since the used
to geometrical type K ribbon
conditions, it wasthermocouple
attached above could
andnot be applied
calibrated for in thespecimen
each notch duetotothe
geometrical
temperature of 850 C by means of a wire thermocouple within the notch. to
conditions, ◦ it was attached above and calibrated for each specimen thethermal
The tem-
perature
gradient ofbetween
850 °C by means
these twoof a wire thermocouple
thermocouples was below within
8 C the
◦ notch.
for 850 ◦ C, The thermal
whereas gradi-
a controlling
entby between
only one these
wiretwo thermocouples
thermocouple was below
attached 8 °C for
at the notch was850 °C,
not whereas adue
practicable controlling
to contact
byproblems
only one during
wire thermocouple
testing. A drop attached at the notch
of the measured andwas not practicable
stabilized due to contact
stress amplitude of 2.35%
problems
indicates the failure of the notched specimen, which is equal to a drop of 2.5%2.35%
during testing. A drop of the measured and stabilized stress amplitude of for the
indicates
smooththe failure ifofa the
specimen notched specimen,
penny-shaped whichofis0.96
crack surface mmto2 in
equal a drop of 2.5% for
both specimen the is
types
assumed [22].
It should be noted here that due to confidential agreements with the project partners,
diagrams which contain lifetime and material data of the investigated material René80
Metals 2021, 11, x FOR PEER REVIEW 5 of 22

smooth specimen if a penny-shaped crack surface of 0.96 mm² in both specimen types is
Metals 2021, 11, 731 assumed [22]. 5 of 21
It should be noted here that due to confidential agreements with the project partners,
diagrams which contain lifetime and material data of the investigated material René80 are
normalized. Wöhler diagrams are labelled with “low” and “high”, which refers to life-
are normalized.
times Wöhler
and total strain, anddiagrams are labelled
stress amplitudes forwith “low”
the low andfatigue
cycle “high”, which
and highrefers to
cycle fa-
lifetimes and total strain, and stress amplitudes for the low cycle fatigue and high
tigue regime. Diagrams which contains material properties are normalized by the highest cycle
fatigue regime.
occurring Diagrams which contains material properties are normalized by the highest
value.
occurring value.
3. Modelling Approaches
3. Modelling Approaches
3.1. Modelling
3.1. Modelling thethe Notch
Notch Strain
Strain with
with Finite
Finite Element
Element Analysis
Analysisfor foran
anIsotropic
Isotropic Material
Assumption
Material Assumption
Dueto
Due tothe
theusage
usageofofthe
the MTS
MTS 12-mm
12-mm high-temperature
high-temperature extensometer
extensometer for both
for both speci-
specimen
men types, the axial strain within the notch must be determined numerically,
types, the axial strain within the notch must be determined numerically, in order to compare in order to
compare
the fatiguethe fatigue
tests tests to
to smooth smooth fatigue
specimen specimen fatigue
tests in straintestsamplitude
in strain amplitude Wöhler
Wöhler diagrams.
diagrams.
Due to the Due
notch togeometry,
the notch stress
geometry, stress concentrations
concentrations occur within occurthewithin the notch
notch root and leadroot
and
to leadplastic
local to local plastic deformation.
deformation. Therefore, Therefore, an isotropic
an isotropic elasticmaterial
elastic plastic plastic material
model, model,
based
based
on on a Ramberg–Osgood
a Ramberg–Osgood relationship
relationship (material
(material parameters
parameters provided
provided by Siemens),
by Siemens), for
for the
the coarse-grained
coarse-grained and fine-grain
and fine-grain textured
textured René80 René80 ◦
at 850at C, 850 was°C,implemented
was implemented into theinto the
finite
finite element
element solver ABAQUS
solver ABAQUS (Version (Version 2018, systemes,
2018, Dessault Dessault Vélizy-Villacoublay,
systemes, Vélizy-Villacoublay,
France). A
virtual
France).extensometer with a measurement
A virtual extensometer length of 12
with a measurement mm was
length of 12attached
mm wassymmetrically
attached sym-
around
metricallythe around
notch ofthea full specimen
notch of a fullmodel.
specimen As amodel.
globalAs loading,
a globaltheloading,
end facetheof end
the model
face of
was displaced
the model wasstep by step,
displaced where
step by step,the where
front face was locked
the front face was in locked
the axial direction.
in the For
axial direc-
each step, the measured axial strain on the extensometer was determined
tion. For each step, the measured axial strain on the extensometer was determined as well as well as the
axial strain within the notch root. The following Figure 3 shows the
as the axial strain within the notch root. The following Figure 3 shows the relationship relationship between
the axial strain
between applied
the axial straintoapplied
the extensometer global (respectively
εglobal (respectively
to the extensometer to the specimen) and the
to the specimen)
corresponding strain in the
and the corresponding notch
strain rootnotch
in the εnotchroot notchcoarse-
for the for theand fine-grain
coarse- textured material
and fine-grain textured
(normalized with the maximum
material (normalized occurringoccurring
with the maximum global strain εglobal
global ). ϵ_global).
strain

Figure3.
Figure 3. Normalized
Normalizedrelationship
relationshipbetween
betweenthe
theglobal
globalstrain
strainatatthe
theextensometer
extensometerεglobal
εglobal and
andthe
thelocal
local
strain at the notch root ε for the coarse- and fine-grain textured material.
strain at the notch root εnotch for the coarse- and fine-grain textured material.
notch

To
To estimate the
therequired
requiredstrain
strainamplitude
amplitude within
within thethe notch
notch root,root,
fittedfitted quadratic
quadratic func-
functions (from Figure 3) were used to adjust the total strain control value at the fatigue
tions (from Figure 3) were used to adjust the total strain control value at the fatigue test
test controlling
controlling system.
system.

3.2.
3.2. Modelling
Modelling the
the Local
Local Material
Material Behaviour
Behaviour with
with Finite
Finite Element
ElementAnalysis
Analysisfor
forPolycrystalline
Polycrystalline
Nickel-Base Superalloys
Nickel-Base Superalloys
In order to simulate the material behaviour in dependence of different grain ori-
entation distributions, the open-source software NEPER (Version 3.3, by Romain Quey,
MINES Saint-Étienne, France) was used to generate random grain morphologies using
the 3D Voronoi tessellation method [23–25]. Since the notched section is of great im-
portance for the mechanical behaviour of the specimen, a cuboid with the dimension
10 mm × 11 mm × 5.5 mm and two notches, according to Figure 4, was modelled. The
However, because of the low notch factor of 1.62, the axial stress component is signifi-
cantly higher than the remaining stress components and therefore these differences in
stress state can be neglected.
According to the grain size distribution mentioned in Section 2, the coarse-grain
Metals 2021, 11, 731 model with random orientation distribution contains of 100 grains, which result in 6 ofan
21
equivalent grain diameter of 2.76 ± 0.55 mm, evaluated using the NEPER grain statistics
tool. The fine-grain textured model contains of 1000 grains with an equivalent grain di-
ameter
thicknessof 1.03
of the± 0.16
model mm. was Due to thetovery
chosen high
ensure computing
that the surface time for generation,
of the two notchesmeshing
is equal
and
to the circumferential surface of the round notch on the test specimendistributions
calculation, three models with three textured grain orientation each
in order to avoid
were created for the fine-grain textured sample. The coarse-grained models
statistical size effects. The cuboid type of model was selected because it is not practicable cover four
different models with three different grain orientation distributions
to perform rotationally symmetrical cuts on cylindrical models within the used version of each. For both types
of models,
NEPER. Duethetovalue for the characteristic
the considerations of a flatlength
notched of specimen
the elements forfinite
in the the average
elementcell size
models,
was set to rcl = 0.3 (relative element length) within the NEPER meshing
differences in the stress state of the notch root occur when compared to a circumferential tool. As a result,
the mesh specimen,
notched consists of asa uniform
used in distribution
the experiment.of approximately
For example,250,000 quadratic
the tangential tetrahedral
stress compo-
elements (C3D4) for the coarse-grain model and approximately 700,000
nent of the circumferential notched specimen is not existent in the flat notched specimen. quadratic tetrahe-
dral elements
However, (C3D4)
because of the forlow
thenotch
fine-grain
factor textured model.
of 1.62, the Both component
axial stress polycrystalline specimen
is significantly
models are shown
higher than in Figurestress
the remaining 4. Note, the edgesand
components within the notch
therefore theseappear in the visualization
differences in stress state
and
can are removed by the meshing procedure.
be neglected.

(a) (b)

50 Grains 1000 Grains

Figure 4. Polycrystalline models of the notched specimen, coarse-grain with 100 grains (a), fine grain with 1000 grains (b).
Figure 4. Polycrystalline models of the notched specimen, coarse-grain with 100 grains (a), fine grain with 1000 grains (b).
According to the grain size distribution mentioned in Section 2, the coarse-grain model
A detailed description of the model properties of the smooth specimen as well as
with random orientation distribution contains of 100 grains, which result in an equivalent
generation is described in [17]. In order to expand the simulation database, six coarse-
grain diameter of 2.76 ± 0.55 mm, evaluated using the NEPER grain statistics tool. The
grain smooth models with random orientation distributions and six fine-grain smooth
fine-grain textured model contains of 1000 grains with an equivalent grain diameter of
1.03 ± 0.16 mm. Due to the very high computing time for generation, meshing and calcula-
tion, three models with three textured grain orientation distributions each were created
for the fine-grain textured sample. The coarse-grained models cover four different models
with three different grain orientation distributions each. For both types of models, the value
for the characteristic length of the elements for the average cell size was set to rcl = 0.3
(relative element length) within the NEPER meshing tool. As a result, the mesh consists of
a uniform distribution of approximately 250,000 quadratic tetrahedral elements (C3D4) for
the coarse-grain model and approximately 700,000 quadratic tetrahedral elements (C3D4)
for the fine-grain textured model. Both polycrystalline specimen models are shown in
Figure 4. Note, the edges within the notch appear in the visualization and are removed by
the meshing procedure.
A detailed description of the model properties of the smooth specimen as well as
generation is described in [17]. In order to expand the simulation database, six coarse-grain
smooth models with random orientation distributions and six fine-grain smooth models
with a textured orientation distribution were added and analysed. Table 2 summarises all
conducted numerical simulations with the corresponding orientation distributions.
Metals 2021, 11, 731 7 of 21

Table 2. Overview of the different numerical simulations.

Grain Number of Models Number of


Number of
Specimen Type Orientation with Different Grain Simulations
Grains
Distribution Morphologies per Model
Notched, flat 100 coarse random 4 3
Notched, flat 1000 fine textured 3 3
Smooth, round 49 coarse random 6 1
Smooth, round 500 fine textured 6 1

In order to create local material properties, each grain was rotated using rotational
matrices U (U is a function of the Euler angles ϕ1 , ϕ2 and ϑ) according to an orientation
distribution in a pre-processing step and the respective data written to the input file. The
grains in the coarse-grain René80 batch show no preferential direction in orientation, which
is why the rotational matrices U used in this pre-processing step are distributed according
to an isotropic measure, mathematically given by the Haar measure at the SO3 group of
rotations [10]. The rotational matrices U which were assigned to the fine-grain textured
models were generated by random distribution of the Euler angles ϕ1 , ϕ2 where ϑ kept
constant with 25◦ to represent the examined texture. To calculate the elastic properties of the
grains in dependence of their local coordinate system, a global, anisotropic, linear-elastic
material law was defined in ABAQUS. As there is no data concerning the elastic constants
of René80 in dependence of the temperature in the literature, elastic values for an IN738LC
were taken from [26]. Since both the composition and content of the γ’ phase are very
similar in IN738LC and René80, it can be assumed that the elastic behaviour of both alloys
are qualitatively comparable. A linear interpolation from the data at 800 ◦ C and 898 ◦ C
results in the elastic constants C11 = 225.83 MPa, C12 = 161.45 MPa and C44 = 98.79 MPa for
a temperature of 850 ◦ C. During the simulation, the local material properties in dependence
of the local grain orientation were transferred to the globally defined material law and
the grains interact according to their local stiffness. All simulations were carried out at
T = 850 ◦ C in displacement control, in order to achieve consistency to the LCF experiments.
The nodes of one face were locked only in the axial direction, in order to enable transverse
contraction, whereas the nodes of the opposite face were displaced by 0.01 mm. Due to the
cyclic stability of the material behaviour during high-temperature LCF testing, only one
single loading cycle was simulated.
A python-based tool was developed to calculate the maximum resulting shear stress
within the slip systems (i,j) on each node of the finite element models in dependence of the
local stress tensor σ and the assigned local grain orientation U. Since the crystallographic
structure of nickel-base superalloys is face-centred cubic (fcc), dislocation movement
occurs usually in slip systems of the type {111}<110> if a critical value τcrit is exceeded.
The resulting shear stress on each node is determined by calculating the shear stresses for
all 12 slip systems, where the slip system with the highest value is the activated one—see
Equation (1).
τres,ij = U·ni σ·U·si,j (1)
Details for the calculation procedure can be found in [10,17,21]. Since all simulations
are purely elastic, only linear elastic shear stresses are calculated within the slip systems.
Still, their distribution at the model surface, i.e., the notch root, gives an indication about
the expected onset of plasticity and therefore the possibility of a fatigue crack initiation.

4. Results
4.1. Experimental Results of Isothermal LCF Tests at 850 ◦ C for Smooth and Notched Specimen
In the following section, the results of the total strain-controlled isothermal LCF ex-
periments at 850 ◦ C of the notched specimen are presented in comparison to the results
of the smooth specimen. Figure 5 shows the influence of grain orientation distribution
and specimen geometry on the high-temperature LCF behaviour of René80 as the total
4. Results
4.1. Experimental Results of Isothermal LCF Tests at 850 °C for Smooth and Notched Specimen
Metals 2021, 11, 731
In the following section, the results of the total strain-controlled isothermal LCF ex-
8 of 21
periments at 850 °C of the notched specimen are presented in comparison to the results of
the smooth specimen. Figure 5 shows the influence of grain orientation distribution and
specimen geometry on the high-temperature LCF behaviour of René80 as the total strain
strain
amplitudeamplitude Wöhler
Wöhler curve.
curve. Fornotched
For the the notched specimen,
specimen, the the
shownshown
totaltotal strain
strain ampli-
amplitude
tude
εa,t,notch represents
represents
εa,t,notch the strain
the total total strain amplitude
amplitude withinwithin the notch
the notch rootSection
root (see (see Section 3.1),
3.1), deter-
determined by finite element simulation using a Ramberg–Osgood
mined by finite element simulation using a Ramberg–Osgood relationship. The investi-relationship. The in-
vestigated coarse-grained specimens are displayed in green, whereas,
gated coarse-grained specimens are displayed in green, whereas, in addition, the results in addition, the
results for coarse-grain
for coarse-grain smoothsmooth
and and notched
notched specimens
specimens (taken
(taken from[14])
from [14])are
areshown
shown in in grey.
grey.
Smooth
Smoothand andnotched
notchedspecimens
specimens made
made from
fromthethe
fine-grain textured
fine-grain material
textured are are
material marked
marked in
blue. Smooth
in blue. specimens
Smooth are shown
specimens as circular
are shown datadata
as circular points, whereas
points, notched
whereas specimens
notched are
specimens
represented
are represented by a by
rhombus. Run Run
a rhombus. out tests are marked
out tests with with
are marked an arrow.
an arrow.

Figure5.5.Total
Figure Totalstrain
strainamplitude
amplitudeWöhler
Wöhler curve
curve forfor smooth
smooth and
and notched
notched specimen
specimen with
with different
different grain
grain orientation distributions for isothermal LCF tests
orientation distributions for isothermal LCF tests at 850 C. ◦at 850 °C.

Figure55clearly
Figure clearlyshows,
shows,for forall
alltotal
totalstrain
strainamplitudes
amplitudesand andgrain
grainorientation
orientationdistribu-
distribu-
tions,that
tions, thatfatigue
fatiguelifetimes
lifetimesofofthethenotched
notchedspecimen
specimenare arehigher
highercompared
comparedtotothe thesmooth
smooth
specimen.As
specimen. Asalready
alreadyexplained
explainedin in[17,21],
[17,21],smooth
smoothspecimens
specimenswith withaafine-grain
fine-graintextured
textured
materialshow
material showhigher
higherlifetimes
lifetimes compared
compared to to
thethe coarse-grain
coarse-grain material,
material, especially
especially for low
for low to-
tal strains.
total However,
strains. However, for for
thethe
notched
notched specimens,
specimens, no unambiguous
no unambiguous improvement
improvement regarding
regard-
fatigue life can
ing fatigue life be
canfound for the
be found for fine-grain
the fine-graintextured material.
textured material.It isItnoticeable
is noticeablethatthat
over a
over
relatively large range of low total strains, both failure and run out occur
a relatively large range of low total strains, both failure and run out occur for the notched for the notched
geometry,
geometry,which which leads distinctlifetime
leads to distinct lifetimescattering.
scattering.AtAt high
high totaltotal strains,
strains, the the lifetime
lifetime scat-
scatter appears
ter appears slightly
slightly decreased.
decreased.
Figure
Figure66showsshows the the measured stress amplitude
amplitude in inorder
ordertotorepresent
representthe theinfluence
influenceof
ofstiffness
stiffness ofof
thethe individual
individual specimens.
specimens. ForFor
thethe notched
notched specimen,
specimen, thethe stress
stress amplitude
amplitude cor-
corresponds to the maximum stress in the
responds to the maximum stress in the notch root (σa,notchnotch root (σ ).
a,notch ).
For the notched specimen, the measured stress amplitudes are in general lower com-
pared to the smooth specimen for the same total strains. While the smooth specimen shows
a clearly lower stress amplitude for the textured fine-grain material, no clear influence
of grain orientation on the stress amplitude for the notched specimen can be determined.
Additionally, the scatter in stress amplitude seems, for the smooth specimen, much higher
compared to the notched specimen, where the highest scatter could be found for the
coarse-grain smooth specimen.
Metals 2021, 11, 731 9 of 21
Metals 2021, 11, x FOR PEER REVIEW 9 of 22

Figure6.6. Stress amplitude


Figure amplitude Wöhler
Wöhlercurve
curvefor
forsmooth
smoothand
andnotched
notchedspecimen with
specimen different
with grain
different grain
orientation distributions.
orientation distributions.

4.2. Microscopic Analysis


For the notched of the Fracture
specimen, Aurfaces stress amplitudes are in general lower com-
the measured
pared to the smooth specimen for the
The fracture surfaces of the notched specimen same total willstrains. Whileinthe
be evaluated thesmooth
following, specimen
where
ashows
detaileda clearly
analysislower
of thestress amplitude
fracture surfaceforforthe
thetextured
smooth fine-grain
specimen can material,
be foundno clear influ-
in [18]. In
ence of the
general, grain orientation
typical on the stress
high-temperature amplitude
fatigue crackingforbehaviour
the notched specimenconsisting
is observed, can be deter-of a
mined. and
smooth Additionally,
oxidized partthewhich
scatter in stressthe
represents amplitude seems,
fatigue crack for the
surface, smooth
which specimen,
was developed
much higher
during cyclingcompared
and a rough to non-oxidized
the notched specimen,
part which where the highest
represents scatterfracture
the residual could be found
surface
(caused by the shutdown
for the coarse-grain smoothof the heating system after specimen failure). Moreover, scanning
specimen.
electron microscope (SEM) images show the typical honeycomb structure and prominent
4.2. Microscopic
cracks Analysis
in the residual of the Fracture
fracture Aurfaces the detailed examination of the fatigue
area. Conversely,
crackThesurface clearly shows the crack initiation
fracture surfaces of the notched specimen sites at thebesurface
will evaluated withinthethe
typical lens
following,
around the initiation spot. Subsequently, river patterns are visible in
where a detailed analysis of the fracture surface for the smooth specimen can be found in the fatigue rupture
surface
[18]. In by SEM investigations,
general, as visible in Figure
the typical high-temperature fatigue7. cracking behaviour is observed, con-
Independent of specimen geometry, the
sisting of a smooth and oxidized part which represents fractured surfaces showcrack
the fatigue an influence
surface, of the
which
total strain amplitude. At low total strain amplitudes, single crack
was developed during cycling and a rough non-oxidized part which represents the resid- initiation with large
fatigue surfaces
ual fracture were(caused
surface observed, as seen
by the in Figure
shutdown 8a.heating
of the Almostsystem
half ofafter
the fracture
specimen surface is
failure).
represented by the fatigue
Moreover, scanning crackmicroscope
electron originating (SEM)
from the single show
images crack initiation
the typical site,honeycomb
while the
other half is dominated by final rupture surface. In contrast, at high total strain amplitudes,
structure and prominent cracks in the residual fracture area. Conversely, the detailed ex-
multiple crack initiation sites distributed all around the notch root were observed, as shown
amination of the fatigue crack surface clearly shows the crack initiation sites at the surface
in Figure 8b. High strains and therefore higher stresses lead to multiple crack initiation
with the typical lens around the initiation spot. Subsequently, river patterns are visible in
sites and faster crack growth. As a result, the fatigue surfaces are significantly smaller and
the fatigue rupture surface by SEM investigations, as visible in Figure 7.
irregularly shaped. Partially, the merging of cracks can be observed. The red arrows in
Figure 8 illustrate the crack initiation site.
Metals 2021, 11, 731 10 of 21
Metals 2021, 11, x FOR PEER REVIEW 10 of 22

200 µm

1 mm

Figure 7. Fatigue and residual fracture surface of a notched specimen. 200 µm


Independent of specimen geometry, the fractured surfaces show an influence of the
total strain amplitude. At low total strain amplitudes, single crack initiation with large
fatigue surfaces were observed, as seen in Figure 8a. Almost half of the fracture surface is
represented by the fatigue crack originating from the single crack initiation site, while the
other half is dominated by final rupture surface. In contrast, at high total strain ampli-
tudes, multiple crack initiation sites distributed all around the notch root were observed,
as shown in Figure 8b. High strains and therefore higher stresses lead to multiple crack
1 mm
initiation sites and faster crack growth. As a result, the fatigue surfaces are significantly
smaller and irregularly shaped. Partially, the merging of cracks can be observed. The red
Figure
arrows7. Fatigue and8residual
in Figure fracture
illustrate surface
the crack of a notched
initiation site. specimen.

Independent of specimen geometry, the fractured surfaces show an influence of the


a)
(a) total strain amplitude. At low(b)total strain amplitudes, single crack initiation with large
fatigue surfaces were observed, as seen in Figure 8a. Almost half of the fracture surface is
represented by the fatigue crack originating from the single crack initiation site, while the
other half is dominated by final rupture surface. In contrast, at high total strain ampli-
tudes, multiple crack initiation sites distributed all around the notch root were observed,
as shown in Figure 8b. High strains and therefore higher stresses lead to multiple crack
initiation sites and faster crack growth. As a result, the fatigue surfaces are significantly
smaller and irregularly shaped. Partially, the merging of cracks can be observed. The red
arrows in Figure 8 illustrate the crack initiation site.

a)
(a) (b)

2 mm
3 mm 30 µm
3 mm

Figure 8. Crack
Crack initiation
initiation sites
sites of
of a notched specimen
specimen fatigued
fatigued with
with (a) low total strain amplitude, (b) high total strain
amplitude
amplitude at
at 850 ◦ C.
850 °C.

4.3. Cyclic Deformation Behaviour of the Notched Specimen under Total Strain Control
The development of stress amplitude σa,notch for the notched specimen over the cycle
number is presented in Figure 9 (total strain amplitudes are normalized).
As expected, the stress amplitude σa,notch decreases with a decrease of total strain
amplitude, which results in longer lifetimes. All curves show a stable cyclic behaviour
during their full lifetime until a crack initiation occurs; therefore, their cyclic behaviour
3 mm2 mm 30 µm
3 mm
is comparable to the smooth specimen [18]. After crack initiation, some samples show
a steep drop in stress amplitude σa,notch . However, stress amplitude increases are also
Figure 8. Crack initiation sites of a notched
possible. specimen
After [27], fatigued
this effect canwith (a) low total
be explained bystrain amplitude,
the position (b) crack
of the high total strainto the
relative
amplitude at 850 °C.
extensometer. Since the duration of the crack growth phase is distinctly smaller compared
to the total lifetime, the effect of the orientation of the crack to the extensometer can
be neglected.
Metals 2021, 11, x FOR PEER REVIEW 11 of 22

4.3. Cyclic Deformation Behaviour of the Notched Specimen under Total Strain Control
Metals 2021, 11, 731 11 of 21
The development of stress amplitude σa,notch for the notched specimen over the cycle
number is presented in Figure 9 (total strain amplitudes are normalized).

Figure 9.
Figure 9. Development
Developmentof ofstress
stressamplitude
amplitudeover
overlifetime during
lifetime strain-controlled
during LCF
strain-controlled testing
LCF at at
testing
850 °C
◦ for coarse-grain and fine-grain textured material.
850 C for coarse-grain and fine-grain textured material.

4.4. Results of the Finite


As expected, Element
the stress Simulation
amplitude σa,notch decreases with a decrease of total strain am-
plitude,
For the determination of the elasticAll
which results in longer lifetimes. curves
axial show
strains a stable
within thecyclic behaviour
notch, results during
of the
their fullsimulations
isotropic lifetime until a crack
based initiation
on the occurs; therefore,
Ramberg–Osgood their cyclic
relationship behaviour
(mentioned is compa-
in Section 3.1)
rable
are to theinsmooth
shown Figure specimen
10. The axial [18].displacement
After crack initiation,
of 0.1% insome samples show
the y–direction a steep
in the drop
following
in stress amplitude
illustrations σa,notchto. However,
was chosen compare the stress amplitude
results with the increases are also
anisotropic possible.and
approaches After
to
[27], this
avoid effect
plastic can be explained
deformation by the position
(the polycrystalline of the crack
simulations arerelative to the
calculated extensometer.
fully elastic). In
Metals 2021, 11, x FOR PEER REVIEW 12 of 22
Since to
order thedifferentiate
duration ofthe thetwocrack growth
material phase the
batches, is distinctly
respectivesmaller
Young’s compared to thefrom
moduli (taken total
the Ramberg–Osgood
lifetime, the effect of therelationship)
orientationwere assigned
of the crack totothetheextensometer
models. can be neglected.

(a)4.4. Results of the Finite Element Simulation (b)


For the determination of the elastic axial strains within the notch, results of the iso-
tropic simulations based on the Ramberg–Osgood relationship (mentioned in Section 3.1)
are shown in Figure 10. The axial displacement of 0.1% in the y–direction in the following
illustrations was chosen to compare the results with the anisotropic approaches and to
avoid plastic deformation (the polycrystalline simulations are calculated fully elastic). In
order to differentiate the two material batches, the respective Young’s moduli (taken from
the Ramberg–Osgood relationship) were assigned to the models.

Figure 10. Isotropic


Figure10. Isotropic elastic simulation
simulation coarse-grain
coarse-grain(a)
(a)material
materialand
and
forfor
thethe fine-grain
fine-grain textured
textured material
material (b), (b),
usingusing a
a Ram-
berg–Osgood approach.
Ramberg–Osgood approach.

The
Thevon
vonMises
Misesstress
stressdistribution
distributionreveals
revealsthe
theanticipated
anticipatedinhomogeneous
inhomogeneousdistribution
distribution
within
withinthe thenotch
notchwhere
wherethe themaximum
maximumvon vonMises
Misesstress
stressisislocated
locatedwithin
withinthethenotch
notchroot,
root,
with
with284.5
284.5MPaMPaforforthe
thecoarse-grain
coarse-grainmaterial
materialand
and273.2
273.2MPa
MPafor forthe
thefine-grain
fine-graintextured
textured
material.
material.Due Duetotothe
theslightly
slightlylower
lowerYoung’s
Young’smoduli
moduliofofthethefine-grain
fine-graintextured
texturedmaterial,
material,the the
local
local stresses are slightly decreased compared to the coarse-grained material. Inaddition,
stresses are slightly decreased compared to the coarse-grained material. In addition,
the
theisotropic
isotropicsimulation
simulationshows
showsa ageometrically
geometricallycaused
causedslight
slightreduction
reductionofofthethestress
stressatatthe
the
edges of the notch.
edges of the notch.
The results of the finite element analysis for the polycrystalline notched specimen at
850 °C and an applied total strain of 0.1% will be presented in the following. Figure 11
shows the distribution of the von Mises stress for a model with coarse-grain material (100
grains, random orientation distribution) and a model with the fine-grain textured material
The von Mises stress distribution reveals the anticipated inhomogeneous distribution
within the notch where the maximum von Mises stress is located within the notch root,
with 284.5 MPa for the coarse-grain material and 273.2 MPa for the fine-grain textured
Metals 2021, 11, 731 material. Due to the slightly lower Young’s moduli of the fine-grain textured material, 12 of the
21
local stresses are slightly decreased compared to the coarse-grained material. In addition,
the isotropic simulation shows a geometrically caused slight reduction of the stress at the
edges of the notch.
The
Theresults
resultsofofthe
thefinite
finiteelement
elementanalysis
analysisfor
forthe
thepolycrystalline
polycrystallinenotched
notchedspecimen
specimenatat
850 ◦ C and an applied total strain of 0.1% will be presented in the following. Figure 11
850 °C and an applied total strain of 0.1% will be presented in the following. Figure 11
shows
showsthethe distribution
distribution of of the
the von
vonMises
Misesstress
stressfor
for a model
a model with
with coarse-grain
coarse-grain material
material (100
(100 grains, random orientation distribution) and a model with the fine-grain textured
grains, random orientation distribution) and a model with the fine-grain textured material
material (1000 grains).
(1000 grains).

(a) (b)

1
1
2

Figure 11. Distribution of von Mises stress for notched specimen for the coarse-grain material (100 grains, (a)) and
Figure 11. Distribution of von Mises stress for notched specimen for the coarse-grain material (100 grains, (a)) and fine-
fine-grained textured material (1000 grains, (b)).
grained textured material (1000 grains, (b)).
It can be clearly seen that compared to an isotropic approach the stress distribution
It can be clearly seen that compared to an isotropic approach the stress distribution
is highly inhomogeneous within the notch area, as well as in the bulk material caused by
is highly inhomogeneous within the notch area, as well as in the bulk material caused by
grains with different orientations and therefore varying elastic properties. Within the notch
root local maxima of stress, located close to grain boundaries, areas that 1 indicate can be
found. At the same time, areas with significantly reduced stresses can occur even in the
notch root (areas 2). A statistical evaluation of all simulations revealed that, on average, the
maximum von Mises stress for the coarse-grained and randomly orientated specimen is
508 ± 58 MPa (evaluated from 12 simulations). The average maximum von Mises stress for
the fine-grain and textured material is with 634 ± 45 MPa significantly increased (evaluated
from nine simulations). It should be noted here that both models were calculated with the
same applied total strain of 0.1% and the same material model; only the grain orientation
distribution and grain numbers were different.
Comparing the von Mises stress results to the isotropic approach presented in Figure 10
reveals distinct higher von Mises stresses for the polycrystalline model. However, they
appear on a much smaller scale within the notch root, whereas, for the isotropic approach,
the whole notch root shows a maximum in von Mises stress.
Figure 12 shows the distribution of the strain component E22 in the loading direction
(y–direction) for the coarse-grain and fine-grain textured material.
Similar to the von Mises stress distribution, the strain distribution is also inhomoge-
neously distributed along the notch area, which results in a distinct strain scatter between
0.1% and 0.3%, where local maxima (area 1) as well as minima (area 2) can be found in the
notch root. Moreover, some models show, in contrast to isotropic calculated models, that
strain maxima can also occur outside the notch root, as area 3 shows. For all nine calculated
models, the maximum strain E22 in loading direction for the fine-grain textured material is
slightly higher with 0.279 ± 0.015% than for the coarse-grain randomly orientated material
with 0.271 ± 0.033% (12 models).
The calculation of the resulting shear stress distribution is carried out with the python
tool mentioned in Section 3.2. Figure 13 shows the distribution of the maxima in the
resulting shear stress within the slip system of the grains and gives an indication of where
plastic deformation, and therefore fatigue crack initiation, occurs favourably.
only the grain orientation distribution and grain numbers were different.
Comparing the von Mises stress results to the isotropic approach presented in Figure
10 reveals distinct higher von Mises stresses for the polycrystalline model. However, they
appear on a much smaller scale within the notch root, whereas, for the isotropic approach,
the whole notch root shows a maximum in von Mises stress.
Metals 2021, 11, 731 13 of 21
Figure 12 shows the distribution of the strain component E22 in the loading direction
(y–direction) for the coarse-grain and fine-grain textured material.

(a) (b)

2
1

Figure
Metals 12.xDistribution
2021, 11, of the strain component E22 in loading direction for notched specimen for the coarse-grain material14 of 22
FOR PEER REVIEW
Figure
(100 12. Distribution
grains, of the strain
(a)) and fine-grained component
textured E22(1000
material in loading
grains,direction
(b)). for notched specimen for the coarse-grain material
(100 grains, (a)) and fine-grained textured material (1000 grains, (b)).

max. tres (a) Similar to the von Mises stress distribution,


max. tres (b) the strain distribution is also inhomoge-
neously distributed along the notch area, which results in a distinct strain scatter between
0.1% and 0.3%, where local maxima (area 1) as well as minima (area 2) can be found in the
notch root. Moreover, some models show, in contrast to isotropic calculated models, that
strain maxima can also occur outside the notch root, as area 3 shows. For all nine calcu-
lated models, the maximum strain E22 in loading direction for the fine-grain textured ma-
terial is slightly higher with 0.279 ± 0.015% than for the coarse-grain randomly orientated
material with 0.271 ± 0.033% (12 models).
The calculation of the resulting shear stress distribution is carried out3 with the python
tool mentioned in Section 3.2. 1Figure 13 shows the distribution of the maxima in the re-
sulting shear stress within the3 slip system of the grains and gives an 2 indication of where
plastic deformation, and therefore fatigue crack initiation, occurs favourably.

Figure 13.Distribution
Figure13. Distributionofofthe
themaximum
maximumshear
shearstress
stresswithin
withinthe
theslip
slipsystem
systemofofthe
thegrains
grainsfor
forthe
thecoarse-grain
coarse-grainmaterial
material(100
(100
grains, (a)) and fine-grained textured material (1000 grains, (b)).
grains, (a)) and fine-grained textured material (1000 grains, (b)).

Due to the inhomogeneous distribution of stresses and strains within the notch, caused
Due to the inhomogeneous distribution of stresses and strains within the notch,
by the polycrystalline modelling approach, the maximum resulting shear stresses are also
caused by the polycrystalline modelling approach, the maximum resulting shear stresses
inhomogeneously distributed. Local maxima can be found within the notch root (area 2)
are also inhomogeneously distributed. Local maxima can be found within the notch root
and also outside the notch root (see area 1). Since the shear stress distribution also depends
(area 2) and also outside the notch root (see area 1). Since the shear stress distribution also
on the stress distribution, areas with low shear stresses can also be found in the notch
depends on the stress distribution, areas with low shear stresses can also be found in the
root, as shown for the areas 3. In these areas, a predominantly elastic deformation occurs,
notch root, as shown for the areas 3. In these areas, a predominantly elastic deformation
whereas plastic deformation can be expected first in the areas with high resulting shear
occurs, whereas plastic deformation can be expected first in the areas with high resulting
stresses. Figure 13 already reveals local higher-shear stresses for the fine-grain textured
shear stresses. Figure 13 already reveals local higher-shear stresses for the fine-grain tex-
material which can be confirmed for all simulations. The average value for the maximum
tured material which can be confirmed for all simulations. The average value for the max-
resulting shear stress is 287 ± 9.8 MPa for the fine-grained textured material, while the
imum resulting
average value forshear stress is 287randomly
the coarse-grain ± 9.8 MPaorientated
for the fine-grained
material istextured
236 ± 26material,
MPa. while
the average
In conclusion, the notched models with a fine-grain textured material26reveal,
value for the coarse-grain randomly orientated material is 236 ± MPa. on
In conclusion, the notched models with a fine-grain textured material reveal,
average, higher mechanical properties, i.e., stress, strains and shear stresses, compared on av-
to notched models with a random orientation distribution (coarse-grain) for the sameto
erage, higher mechanical properties, i.e., stress, strains and shear stresses, compared
notchedcondition.
loading models with a random orientation distribution (coarse-grain) for the same load-
ing condition.

4.5. Shear Stress Distribution of the Smooth Specimen


In addition to [17], further smooth-specimen simulations were carried out in order to
increase the database. Figure 14 shows the distribution of shear stress for coarse-grain and
fine-grain textured models at 850 °C and an applied total strain of 0.25%.

max. tres (a) max. tres (b)


tured material which can be confirmed for all simulations. The average value for the max-
imum resulting shear stress is 287 ± 9.8 MPa for the fine-grained textured material, while
the average value for the coarse-grain randomly orientated material is 236 ± 26 MPa.
In conclusion, the notched models with a fine-grain textured material reveal, on av-
erage, higher mechanical properties, i.e., stress, strains and shear stresses, compared to
Metals 2021, 11, 731 14 of 21
notched models with a random orientation distribution (coarse-grain) for the same load-
ing condition.

4.5.4.5. Shear
Shear Stress
Stress Distribution
Distribution of the
of the Smooth
Smooth Specimen
Specimen
In In addition
addition to to [17],
[17], further
further smooth-specimen
smooth-specimen simulations
simulations were
were carried
carried outout in order
in order to to
increase
increase thethe database.
database. Figure
Figure 14 14 shows
shows thethe distribution
distribution of of shear
shear stress
stress forfor coarse-grain
coarse-grain andand
fine-grain
fine-grain textured
textured models
models at 850 ◦ C°C
at 850 andand
anan applied
applied total
total strain
strain of 0.25%.
of 0.25%.

max. tres (a) max. tres (b)

Figure 14. Distribution of the maximum shear stress within the slip system for smooth specimen with coarse-grain material
(49 grains, (a)) and fine-grained textured material (500 grains, (b)).

For the simulation of the smooth specimen, areas with low resulting shear stress
as well as areas with locally significant increased shear stresses occurring on the surface
can be clearly seen. An evaluation of the average maximum resulting shear stress of all
calculated models (including the results from [17]) shows, for the coarse-grain, random
orientated specimen, an average resulting shear stress of 397 ± 25 MPa. The average value
for fine-grain textured material is clearly lower with about 315 ± 21 MPa.
A summary of the finite element simulations reveals for the fine-grain textured mate-
rial higher local mechanical properties (stress, strain and shear stresses) when compared to
the coarse-grain material for the case of a notched specimen. Interestingly, in the case of a
smooth sample, the fine-grain textured material led to distinct lower mechanical properties.
The following section will explain where these differences in mechanical behaviour result
from and how they affect the fatigue behaviour.

5. Discussion
A general difference of lifetime behaviour between the smooth and notched specimen
can be statistically attributed to the size effect. For the smooth specimen, the possible area
for crack initiation, i.e., the gauge section surface, is about 6.5 (the whole surface of the
notch) times higher for the notched specimen. However, just around a third of the notch
surface lies within the notch root; therefore, the surface relation between the notched and
smooth specimen where fatigue crack probably occurs is much higher. Despite more grains
on the surface of the fine-grain textured smooth specimen which from a statistical aspect
should increase the possibility of a crack initiation, a longer lifetime can be observed for the
same total strains in comparison to the coarse-grain smooth specimen. The reason lies in
the preferred orientation of the grains in combination with the global uniaxial stress state.
This results in lower Young’s moduli of the specimen compared to a randomly orientated
material. It follows for strain-controlled test conditions, lower stress amplitudes and lower
shear stresses within the slip systems and therefore longer lifetimes [17]. A significantly
improved lifetime of the fine-grain textured material, however, cannot be determined
for the notched samples. Due to the multiaxial character of the notch and the resulting
multiaxial stress states, an evaluation of Young’s moduli as a ratio of stress and strain
in loading direction is not appropriate. To determine the influence of stiffness by using
finite element, the reaction forces in the loading direction of all nodes were calculated and
Metals 2021, 11, 731 15 of 21

summarized to calculate the resulting force on the specimen caused by the displacement
of 0.1%. Related to the cross section within the notch, the averaged nominal stress which
occurs for a displacement of 0.1% for the fine-grain textured material is 240 ± 1.5 MPa.
For the randomly orientated coarse-grain material, the average stress in the notch can be
calculated to 190 ± 7 MPa. It could be assumed here that, for the coarse-grain material, an
insufficient number of grains with perhaps stiff orientations could influence the results. As
mentioned in [28], a minimum of six to eight grains are required within the cross section
to minimize the influence of different grain orientations. For the coarse-grain notched
specimen, more than 10 grains are located in the cross section of the notch for every
model. In addition, the number of simulations (12 in total) with random grain orientation
distributions, generated with the Haar measure, ensures a sufficient amount of data to
provide secure statistical evaluations.
Thus, the averaged values reveal a general higher stress required for the textured
material to achieve a total strain of 0.1% than for the randomly oriented material. Therefore,
the notched specimens show a contrary behaviour to the smooth specimens with regard to
stiffness. Since all models, notched and smooth, were calculated with the same material
Metals 2021, 11, x FOR PEER REVIEW 16 of 22
models as well as orientation distribution, the main the differences in mechanical behaviour
can be attributed to the influence of the notch.

5.1. Development of Stiffness along


5.1. along the
the Notch
Notch
The following
The following section explains where the stiffness differences arise from by investi-
gating the
gating the elements
elements of of the
the finite
finite elements
elements models
models andand assuming
assuming thatthat elements
elements can can have
have
different elastic
different elastic properties
propertiesduedueto to their
their local
local orientation
orientationin inrelation
relationtotothetheloading
loadingaxis.axis. For
For
all the
all the explanations
explanations given given below,
below, an an anisotropic
anisotropicmaterial
materialbehaviour
behaviourisisassumed
assumeddue duetoto the
the
texture
texture and and differences
differences inin the
the alignment
alignment of of the
the grains
grains (i.e.,
(i.e., elements)
elements) between
between the the local
local
principal
principal stress axis and and loading
loadingaxis,axis,but
butthethemicrostructure
microstructureofofthe thematerial
material is is assumed
assumed to
to
bebe homogenous
homogenous (non-considerationofofgrains).
(non-consideration grains).Therefore,
Therefore,aa grain
grain interaction caused by by
differently
differently shapedshaped andand orientated
orientated grains
grains isis not
not considered
considered here.
here. For the smooth specimen,
the
the loading
loading axis axis aligns
aligns with
with the
the axis
axis ofof maximum
maximum principal
principal stress
stress σσII for
for all
all elements
elements (σ (σIIII,
σ
σIII << Iσ
III<<σ ).I ). Assumingthe
Assuming thedetermined
determinedcrystallographic
crystallographic texture
texture for allall elements,
elements,with withϕφ1 ,1,ϕφ22
random ◦
random and ϑϑ constant constant with
with 2525°,, the
the Young’s
Young’s moduli,
moduli, i.e.,
i.e., stiffness,
stiffness, are
are nearly
nearly uniformly
uniformly
distributed with an average value of 128 ± 0.9 GPa, calculated for 850 ◦ C. This outcome
distributed with an average value of 128 ± 0.9 GPa, calculated for 850 °C. This outcome
results
results from
from Monte
Monte Carlo
Carlo simulations
simulations in in which
which aa single
single element
element waswas rotated
rotated 10,000
10,000 times
times
while
while the stiffness in loading direction (direction of maximum principal stress)was
the stiffness in loading direction (direction of maximum principal stress) wasevalu-
eval-
ated.
uated.For Forthe thenotched
notchedspecimen,
specimen,the thedirection
directionofofmaximum
maximumprincipal
principalstress
stressisisnot
notconstant
constant
and
and changes
changes for for each
each element
element along
along the
the notch due to
notch due to its
its geometry,
geometry, as as Figure
Figure 15 15 illustrates.
illustrates.
Loading axis

Figure 15. Directions


Figure 15. Directions of
of maximum
maximum principle
principle stress
stress along
alongthe
thenotch
notchcompared
comparedto
toloading
loadingdirection
directiony.
y.

It can be clearly seen that along the notch, an angle ψ between the maximum principal
stress and global loading axis forms. If now a constant texture, i.e., constant orientation of
each element is assumed, the principal stress direction of the elements aligns tangentially
Metals 2021, 11, 731 16 of 21

It can be clearly seen that along the notch, an angle ψ between the maximum principal
stress and global loading axis forms. If now a constant texture, i.e., constant orientation of
each element is assumed, the principal stress direction of the elements aligns tangentially
along the notch with regard to the global loading axis y. This assumption is valid for
notches with a circular or elliptical shape (no sharp notches). For elements in the notch root,
the difference angle is ψ = 0◦ and therefore the results of the Monte Carlo simulation to
determine the possible stiffness at this point are similar to the results for a uniaxial case in
the smooth specimen (the direction of principal stress align with the loading direction). For
calculating the stiffness for elements with angles ψ 6= 0◦ , the following method was applied.
Instead of calculating the stiffness for the constant texture with changing directions of the
maximum principal stress (to represent the notch), the principal stress direction was kept
constant, and the orientation of the elements were changed. For an angle of 5◦ between the
loading direction and principal stress axis, the element was rotated about ψ = 5◦ , in such
Metals 2021, 11, x FOR PEER REVIEW
a way that the principle stress axis and global loading direction align. For the textured 17 of 22
◦ ◦
material with ϑ = 25 follows as the new angle ϑ + ψ = 30 . Monte Carlo simulations
were carried out for angles ψ = 5◦ , 10◦ , . . . , 60◦ (3000 samples each). Figure 16 shows the
development of a stiffnessofincrease
shows the development in dependence
a stiffness of the angle ψ
increase in dependence of for
theaangle
textured
ψ fororientation
a textured
distribution with ϑ = 25 ◦ for all elements along the notch (ϕ , ϕ = random).
orientation distribution with ϑ = 25° for all elements along 1the notch
2 (φ1, φ2 = random).

Figure16.
Figure 16.Development
Developmentofofstiffness
stiffnessalong
alongthe
thenotch
notchfor
forthe
thetextured
texturedmaterial.
material.

Theanalysis
The analysisof ofthe
theMonte
MonteCarlo
Carlosimulation
simulationreveals
revealsananincrease
increaseininstiffness
stiffnessuntil
untilaa
maximum value with 200 ± 18 GPa is reached for ψ = ◦
25° at the notch.
maximum value with 200 ± 18 GPa is reached for ψ = 25 at the notch. With a further With a further
increasein
increase in ψ,
ψ, the stiffness decreases
decreases with
withsimultaneously
simultaneouslyincreasing
increasingscatter,
scatter,which
whichcan be
can
explained
be explained byby
plotting thethe
plotting inverse polepole
inverse figure for different
figure angles
for different ψ over
angles the Young’s
ψ over mod-
the Young’s
uli at 850
moduli at 850 ◦ C,presented
°C, as in Figure
as presented 17. 17.
in Figure
Possible orientations (black dotted line) with a fixed angle of ψ = 0◦ (equal to the
determined texture with ◦
[111]ϑ = 25 ) show low corresponding Young’s moduli and are located
[111]
in a region where the Young’s moduli itself is nearly homogeneously 250
distributed, which
results in a negligible scatter. With an increase in ψ, the possible orientations cross through
areas where the width of possible values for Young’s moduli increases,200 which also increases
the scatter. As shown in Figure 16, the highest average values occur for ψ between 20◦ and
30◦ . This is due to the case that some orientations align close to the [111] 150axis, which result
in the highest Young’s moduli values of about 250 GPa. For ψ > 30◦ , the scatter remains
nearly constant with decreasing average value. As can be seen at the pole figures, with
further increasing ψ, the possible Young’s moduli values move closer100 to areas close to the
[100] and [110] orientations, which results in a decrease of the average Young’s moduli. For
these orientations, the scatter remains quite high, since the possible orientations cross areas
with a broad range of possible Young’s moduli.

[001] [011] [001] [011]

Figure 17. Inverse pole figure plot for different orientation distributions with constant value of ψ over the Young’s moduli
at 850 °C.
The analysis of the Monte Carlo simulation reveals an increase in stiffness until a
maximum value with 200 ± 18 GPa is reached for ψ = 25° at the notch. With a further
increase in ψ, the stiffness decreases with simultaneously increasing scatter, which can be
Metals 2021, 11, 731 17 of
explained by plotting the inverse pole figure for different angles ψ over the Young’s 21
mod-
uli at 850 °C, as presented in Figure 17.

[111] [111]
250

200

150

100

Metals 2021, 11, x FOR PEER REVIEW 18 of 22


[001] [011] [001] [011]
Figure 17. Inverse pole figure plot for different orientation distributions with constant value of ψ over the Young’s moduli
Figure◦17. Inverse pole figure plot formoduli.
Young’s differentFor
orientation distributionsthe
these orientations, withscatter
constant value ofquite
remains ψ over the Young’s
high, moduli
since the possible
at 850 C.
at 850 °C. orientations cross areas with a broad range of possible Young’s moduli.
Fromthis
From thisbehaviour
behaviour follows
follows aa huge
huge stiffness discrepancy with with an
an average
average 7070GPa
GPa
higherPossible
higher stiffness
stiffness orientations
at (black
25◦ compared
at ψψ == 25° compared dotted line)
to 128
to 128 GPa
GPawith
at aψ fixed
at ψ
◦ . angle
== 00°. of ψbe
It should
It should be= noted
0° (equal
noted thatto
that the
this
this
determined
behaviour
behaviour texture
isisonly
onlyvalidwith
valid ϑ elements
for
for =elements
25°) show onlow
on the corresponding
the surface of
surface of the Young’s
the notch
notch moduli
root.
root. For and arein
For elements
elements located
inthe
the
in a region
direction of where
the bulk the Young’s
material moduli itself
(x–direction), is
the nearly homogeneously
orientation changes of
direction of the bulk material (x–direction), the orientation changes of principal stress distributed,
principal stresswhich
are
are
resultson
shown
shown in aaacross
on negligible
crosssectionscatter.
section Withthe
through
through annotch
the increase
notch in ψ, the
in Figure
in Figure 18.possible orientations cross through
18.
areas where the width of possible values for Young’s moduli increases, which also in-
creases the scatter. As shown in Figure 16, the highest average values occur for ψ between
20° and 30°. This is due to the case that some orientations align close to the [111] axis,
which result in the highest Young’s moduli values of about 250 GPa. For ψ > 30°, the scat-
ter remains nearly constant with decreasing average value. As can be seen at the pole
figures, with further increasing ψ, the possible Young’s moduli values move closer to ar-
y eas close to the [100]y and [110] orientations, which results in a decrease of the average

Figure18.
Figure Crosssection
18.Cross sectionofofelements
elementsthrough
through the
the notch
notch with
with directions
directions of
of principle
principle stress.
stress.

For the first couple of elements close to the notch, a distinct difference angle ψ between
For the first couple of elements close to the notch, a distinct difference angle ψ be-
the global loading axis y and principal stress direction can be identified. With increasing
tween the global loading axis y and principal stress direction can be identified. With in-
distance from the notch (in x–direction), the maximum principal stress aligns more and
creasing distance from the notch (in x–direction), the maximum principal stress aligns
more with the global loading direction. Transferred to the results for the elements in the
more and more with the global loading direction. Transferred to the results for the ele-
notch, a stiffness gradient follows with high stiffness at the notch root and decreasing
ments in the notch, a stiffness gradient follows with high stiffness at the notch root and
values into the bulk material. Figure 19 shows the development of stiffness into the bulk
decreasing values ainto
material through the
cross bulk calculated
section material. Figure 19 shows
by Monte the development
Carlo simulations. of stiffness
The cross section
into the bulk material through a
◦ cross section calculated by Monte Carlo
is taken horizontally at ψ = 25 since it is the found maximum—see Figure 16. simulations. The
cross section
As can isbetaken horizontally
seen on at ψof= the
the direction 25° principal
since it isstress
the found maximum—see
for each chosen elementFigure
in
16.
Figures 18 and 19, a uniaxial stress state is already reached after a couple of elements,
which is less than a millimetre into the bulk material. This can lead to high stiffness
differences between the notch area and the bulk material which reach up to 104 GPa in
the worst case. The majority of bulk material, on the other hand, has only low stiffness
due to the predominant uniaxial stress state in combination with the examined texture
(comparable to the stress state in the smooth specimen).

Begin of uniaxial stress state


creasing distance from the notch (in x–direction), the maximum principal stress aligns
more and more with the global loading direction. Transferred to the results for the ele-
ments in the notch, a stiffness gradient follows with high stiffness at the notch root and
decreasing values into the bulk material. Figure 19 shows the development of stiffness
Metals 2021, 11, 731 into the bulk material through a cross section calculated by Monte Carlo simulations.
18 ofThe
21
cross section is taken horizontally at ψ = 25° since it is the found maximum—see Figure
16.

Begin of uniaxial stress state

Figure19.
Figure 19.Development
Developmentof
ofstiffness
stiffnessinto
intothe
thebulk
bulkmaterial
materialfor
forthe
thefine-grain
fine-graintextured
texturedmaterial.
material.

Interestingly, the on
highest resulting stiffness (associated with aneach
anglechosen ◦ ) lies still
ψ = 25element
As can be seen the direction of the principal stress for in
in the range
Figures of the
18 and 19,highest stresses,
a uniaxial stresscalculated for thereached
state is already isotropic approach
after a couple caused by the
of elements,
notch
whichand therefore
is less than a in the critical
millimetre area
into thefor crack
bulk initiation.
material. This Incansummary, it can
lead to high be stated
stiffness dif-
for the notched
ferences between specimen
the notch with
areathe examined
and the bulktextured
materialorientation
which reach distribution
up to 104 GPa that inhigh
the
stiffness appears
worst case. in the highly
The majority of bulkloaded andontherefore
material, the othercritical areas
hand, has of the
only lownotch duedue
stiffness to ato
stiffness increase along
the predominant the notch.
uniaxial This results
stress state in high local
in combination with stresses as well astexture
the examined shear stresses,
(compa-
as seen by the finite element simulations
rable to the stress state in the smooth specimen). in Section 4.3. The bulk material, where the
loading direction and maximum principal stress direction align, show significantly lower
stiffness due to the texture. Thus follows for the examined textured in combination with a
notched specimen geometry, that the total stiffness can be seen as a composite of the low
stiffness of the bulk and potentially high stiffness of the notched area.
For a real considered polycrystalline and anisotropic material, the local mechanical
behaviour is much more complex. Besides the previous described differences in align-
ment between maximum principal stress and global loading axis caused by the geometric
character of the notch, further effects occur. Especially, the interaction between grains
of different sizes and orientations where resulting constraints lead to inhomogeneous
stress states along the notch as well as in the bulk material, as seen in the finite element
simulations. Due to the complexity, grain interactions are not considered for the evaluation
of stiffness properties.

5.2. Influence of Grain Orientation Distribution on the Shear Stress in Slip Systems
Besides the influence on stiffness, the texture of the fine-grain textured material also
leads to a preferred orientation of the slip systems. To determine the onset of plasticity,
the maximum resulting shear stress within the slip system was calculated according to
Equation (1) and the modified Schmid factor obtained by dividing the resulting shear stress
by the von Mises stress of the node. Due to the node-based formulation of the python
code, the polycrystalline finite element models can be considered, whereby the local grain
interaction is taken into account for the resulting shear stress evaluation. While for the
smooth sample all nodes on the surface were evaluated, for the notch sample only nodes in
the area of the notch root (location of possible crack initiation) were considered. Figure 20
shows the distribution of the modified Schmid factor for the smooth and notched specimen
for the two different investigated grain orientation distributions.
stress by the von Mises stress of the node. Due to the node-based formulation of the py-
thon code, the polycrystalline finite element models can be considered, whereby the local
grain interaction is taken into account for the resulting shear stress evaluation. While for
the smooth sample all nodes on the surface were evaluated, for the notch sample only
Metals 2021, 11, 731
nodes in the area of the notch root (location of possible crack initiation) were considered.
19 of 21
Figure 20 shows the distribution of the modified Schmid factor for the smooth and
notched specimen for the two different investigated grain orientation distributions.

(a) (b)

Figure
Figure20.
20.Distribution
Distributionof of modified
modified Schmid
Schmid factor for random-orientated
random-orientated coarse-grain
coarse-grainand
andfine-grain
fine-graintextured
texturedsmooth
smooth(a)
(a)
(according to [17]) and notched specimen (b).
(according to [17]) and notched specimen (b).

The
The averaged
averagedmodified
modifiedSchmid
Schmidfactor
factorfor
forthe
thesmooth
smoothspecimen
specimenisis0.4310.431± ±
0.05 forfor
0.05 the
the fine-grain textured material and therefore significantly
fine-grain textured material and therefore significantly lower lower than for the coarse-grain
for the coarse-grain
materialwith
material 0.468± ±
with0.468 0.05.
0.05. Figure
Figure 20b20b shows
shows the distribution
the distribution of theofmodified
the modified Schmid
Schmid factor
factor
for for all simulations
all simulations carriedcarried outawith
out with a notched
notched model.model.
It can It
becan be clearly
clearly seenthe
seen that thatdistri-
the
distributions
butions of theofmodified
the modified
SchmidSchmid factor
factor showshow similar
similar profiles
profiles for for
thethe notched
notched specimen
specimen for
for both orientation distributions. The average modified Schmid factor for the fine-grain
textured material is 0.448 ± 0.05 and nearly similar to the value for the coarse-grained
randomly orientated material with 0.454 ± 0.06. The nearly similar modified Schmid factor
distribution can be explained by the change of the direction of the principal stress along the
notch. For the smooth specimen, the loading direction is constant and lead by the constant
texture to an improvement of the modified Schmid factor. For the notched specimen, the
texture is also constant, but the direction of maximum principle stress changes along the
notch, which leads to a variation in Schmid factors due to constant texture with varying
principal stress direction (analogue to the stiffness). Since nearly similar modified Schmid
factor distributions were found for the notched specimen, the increased values of the
resulting shear stress of about 50 MPa (see Figure 13) for the fine-grained textured material
can be attributed to the evaluated higher stiffness along the notch root. The increased
stiffness in combination with slightly higher strains (see Figure 12) leads to higher stresses
in the notch root and finally to higher local shear stresses.
A slightly shorter lifetime associated with the slightly higher shear stress of the fine-
grain textured and notched specimen cannot be clearly evaluated regarding the stress and
strain Wöhler curves shown in Figures 5 and 6. Due to the low amount of data, it is very
difficult to make reliable statements about differences in lifetime due to the large scatter.

6. Conclusions and Outlook


In the present work, the influence of a random and textured grain orientation distribu-
tion on the mechanical as well as fatigue behaviour was investigated for the polycrystalline
nickel-base superalloy René80. Isothermal LCF tests at 850 ◦ C with notched specimen show
a stable cyclic behaviour over the lifetime independent of the applied strain amplitude.
Analogous to the fatigue behaviour of smooth specimen, low total strain leads to single
crack initiation, whereas high total strain results in multi-crack initiation sites. Overall, for
same total strain amplitudes, the notched specimen reveals distinct higher lifetimes com-
pared to the smooth specimen, which can be mainly attributed to a significantly smaller
highly loaded volume with distinct smaller grain numbers. In contrast to the smooth
specimen, where the fine-grain textured improved the fatigue life due to lower Young’s
moduli and modified Schmid factor, no improvement was found for the notched specimen.
This can mainly be attributed to the development of an area with high stiffness along the
notch, caused by constant grain texture with varying principal stress directions generated
by the notch geometry. Under loading, this results in slightly higher stresses and therefore
Metals 2021, 11, 731 20 of 21

to higher shear stresses in the notch root area compared to the random orientated material.
Since the distribution of modified Schmid factors show no differences for the notched
random and textured orientated specimen, the increased shear stress results mainly from
the increased stiffness development along the notch. According to minor differences in
shear stress for the random and textured material with about 50 GPa, no deviations in
lifetime could be observed, which can be attributed to the small number of tests as well as
to the occurring scatter in lifetime.
In summary, it can be stated that the examined material texture is beneficial for
uniaxial fatigue loading in combination with smooth specimens due to its improvement in
elastic properties as well as the onset of plasticity. However, for notched specimens, these
advantages disappear and a nearly similar fatigue behavior evolves between the textured
and random-orientated material.
Using the presented modeling approach, it is furthermore possible to design grain
orientation distributions which generate minima in stress as well as resulting in shear stress
for complex geometries, to improve the component performance under high-temperature
LCF conditions.

Author Contributions: B.E. wrote the present paper and executed the material modelling. Experi-
mental work and analysis was carried out by B.E. and supported S.O. L.M., and T.B. were mainly
involved in discussing and interpreting the results as well as in supervising the work. All authors
have read and agreed to the published version of the manuscript.
Funding: The investigations were conducted as part of the joint research program COOREFLEX
TURBO in the frame of AG Turbo. The work was supported by the Federal Ministry for Economic Af-
fairs and Energy as per resolution of the German Federal Parliament under grant number 03ET7041K.
Data Availability Statement: Data not available due to commercial restrictions.
Acknowledgments: The authors gratefully acknowledge AG Turbo and Siemens AG for their support
and permission to publish this paper. The responsibility for the content lies solely with its authors.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Walsh, P.P.; Fletcher, P. Gas Turbine Performance; Blackwell Science Ltd.: Oxford, UK, 2004.
2. Ramberg, W.; Osgood, W.R. Description of Stress-Strain Curves by Three Parameters; National Advisory Committee for Aeronautics:
Edwards, CA, USA, 1943.
3. Zhao, L.G.; Tong, J.; Vermeulen, B.; Byrne, J. On the uniaxial mechanical behaviour of an advanced nickel base superalloy at high
temperature. Mech. Mater. 2001, 33, 593–600. [CrossRef]
4. Zhang, Z.; Yu, H.; Dong, C. LCF behavior and life prediction method of a single crystal nickel-based superalloy at high
temperature. Front. Mech. Eng. 2015, 10, 418–423. [CrossRef]
5. Neidel, A.; Riesenbeck, S.; Ullrich, T.; Völker, J.; Yao, C. Hot cracking in the HAZ of laser-drilled turbine blades made from René
80. Mater. Test. 2005, 47, 553–559. [CrossRef]
6. Siavashani, R.S.; Novinrooz, A.J.; Taherkhani, A. The Effect of Different Continuous Cooling Rates on Lattice Constant and
Morphology of the Precipitates in Nickel-Base Super Alloy Rene 80. J. Basic. Appl. Sci. 2013, 3, 14–19.
7. Donachie, M.J.; Donachie, S. Superalloys: A Technical Guide, 2nd ed.; ASM International: Russell Township, OH, USA, 2002.
8. Kulawinski, D. Biaxial-Planare Isotherme und Thermo-Mechanische Ermüdung an polykristallinen Nickelbasis-Superlegierungen; Logos
Verlag: Berlin, Germany, 2015.
9. Holländer, D. Experimentelles Verfahren zur Charakterisierung des Einachsigen Ermüdungsverhaltens auf Basis Miniaturisierter
Prüfkörper und Anwendung auf Hochtemperatur-Legierungen der Energietechnik. Ph.D. Thesis, Logos Verlag, Berlin, Germany,
2017.
10. Moch, N. From Microscopic Models of Damage Accumulation to the Probability of Failure of Gas Turbines. Ph.D. Thesis,
Universitätsbibliothek Wuppertal, Wuppertal, Germany, 2019.
11. Schmitz, S.; Seibel, T.; Beck, T.; Rollmann, G.; Krause, R.; Gottschalk, H. A probabilistic model for LCF. Comput. Mater. Sci. 2013,
79, 584–590. [CrossRef]
12. Gottschalk, H.; Schmitz, S.; Seibel, T.; Rollmann, G.; Krause, R.; Beck, T. Probabilistic Schmid factors and scatter of low cycle
fatigue (LCF) life. Materialwissenschaft und Werkstofftechnik 2015, 46, 156–164. [CrossRef]
13. Mäde, L.; Kumar, K.; Schmitz, S.; Gundavarapu, S.; Beck, T. Evaluation of component-similar rotor steel specimens with a local
probabilistic approach for LCF. Fatigue Fract. Eng. Mater. Struct. 2019, 91, 319. [CrossRef]
Metals 2021, 11, 731 21 of 21

14. Seibel, T. Einfluss der Probengröße und der Kornorientierung auf die Lebensdauer einer polykristallinen Ni-Basislegierung bei
LCF-Beanspruchung. Ph.D. Thesis, Forschungszentrum Jülich, Jülich, Germany, 2014.
15. Mäde, L.; Schmitz, S.; Gottschalk, H.; Beck, T. Combined notch and size effect modeling in a local probabilistic approach for LCF.
Comput. Mater. Sci. 2018, 142, 377–388. [CrossRef]
16. Engel, B.; Beck, T.; Moch, N.; Gottschalk, H.; Schmitz, S. Effect of local anisotropy on fatigue crack initiation in a coarse grained
nickel-base superalloy. In Proceedings of the 12th International Fatigue Congress (FATIGUE 2018), Poitiers, France, 27 May–1
June 2018. [CrossRef]
17. Engel, B.; Mäde, L.; Lion, P.; Moch, N.; Gottschalk, H.; Beck, T. Probabilistic Modeling of Slip System-Based Shear Stresses and
Fatigue Behavior of Coarse-Grained Ni-Base Superalloy Considering Local Grain Anisotropy and Grain Orientation. Metals 2019,
9, 813. [CrossRef]
18. Engel, B.; Beck, T.; Schmitz, S. High temperatue Low Cylce Fatigue of the Ni-base Superallor René80. In Proceedings of the Eighth
International Conference on Low Cycle Fatigue (LCF8), Dresden, Germany, 27–29 June 2018.
19. Safari, J.; Nategh, S. On the heat treatment of Rene-80 nickel-base superalloy. J. Mater. Process. Technol. 2006, 176, 240–250.
[CrossRef]
20. Takaki, T.; Sakane, S.; Ohno, M.; Shibuta, Y.; Aoki, T.; Gandin, C.-A. Competitive grain growth during directional solidification of
a polycrystalline binary alloy: Three-dimensional large-scale phase-field study. Materialia 2018, 1, 104–113. [CrossRef]
21. Engel, B. Einfluss der Lokalen Kornorientierung und der Korngröße auf das Verformungs- und Ermüdungsverhalten von
Nickelbasis Superlegierungen. Ph.D. Thesis, Technische Universität Kaiserslautern, Kaiserslautern, Germany, 2019.
22. ISO International Organization for Standardization. Metallic Materials—Fatigue Testing—Axial-Strain-Controlled; ISO International
Organization for Standardization: Geneva, Switzerland, 2017.
23. Quey, R.; Dawson, P.R.; Barbe, F. Large-scale 3D random polycrystals for the finite element method: Generation, meshing and
remeshing. Comput. Meth. Appl. Mech. Eng. 2011, 200, 1729–1745. [CrossRef]
24. Ghazvinian, E.; Diederichs, M.S.; Quey, R. 3D random Voronoi grain-based models for simulation of brittle rock damage and
fabric-guided micro-fracturing. J. Rock Mech. Geotech. Eng. 2014, 6, 506–521. [CrossRef]
25. Quey, R.; Renversade, L. Optimal polyhedral description of 3D polycrystals: Method and application to statistical and synchrotron
X-ray diffraction data. Comput. Meth. Appl. Mech. Eng. 2018, 330, 308–333. [CrossRef]
26. Hermann, W.; Sockel, H.G.; Han, J.; Bertram, A. Elastic Properties and Determination of Elastic Constants of Nickel-Base
Superalloys by a Free-Free Beam Technique. In Proceedings of the Eighth International Symposium on Superalloys, Champion,
PA, USA, 22–26 September 1996; pp. 229–238.
27. Gollmer, M. Schadensakkumulationsverhalten der Superlegierung René 80 unter Zweistufiger Low Cycle Fatigue Beanspruchung.
Ph.D. Thesis, Technische Universität Kaiserslautern, Kaiserslautern, Germany, 2018.
28. Hyde, T.H.; Sun, W.; Williams, J.A. Requirements for and use of miniature test specimens to provide mechanical and creep
properties of materials: A review. Int. Mater. Rev. 2007, 52, 213–255. [CrossRef]

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy