Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Version of Record: https://www.sciencedirect.

com/science/article/pii/S0921509317317112
Manuscript_c985a8967f2ced019cd55a430a80a4d1

Is there an optimal grain size for creep resistance in Ni-based disk


superalloys?
Louis Thébaud1,2, Patrick Villechaise2, Coraline Crozet1, Alexandre Devaux1, Denis
Béchet1, Jean-Michel Franchet3, Anne-Laure Rouffié3, Michael Mills4, Jonathan
Cormier2,*
1 Aubert & Duval, Site des Ancizes, Research and Development, BP1, 63770, Les Ancizes, France
2 Institut Pprime, UPR CNRS 3346, Physics and Mechanics of Materials Department, ISAE-ENSMA, BP 40109,
86961, Futuroscope-Chasseneuil Cedex, France
3 Safran SA, SafranTech – Materials & Process Division, Rue des Jeunes Bois – Châteaufort – CS 80112, 78772,

Magny-Les-Hameaux, France
4 Department of Materials Science and Engineering, The Ohio State University, Columbus, OH, USA

*Corresponding author: jonathan.cormier@ensma.fr

Abstract
The high temperature creep properties of next generation cast and wrought AD730 superalloy
have been investigated taking into consideration three microstructural parameters: the grain size,
the presence of grain boundaries and the γ′ precipitates size and distribution. Definitive analysis
of the influence of the grain boundaries and γ′ precipitates size distribution has been enabled by
the study of single crystalline versions of the polycrystalline alloys studied. At high temperature
(equal to or in excess of 850°C), the grain size controls creep properties. Comparisons between
polycrystalline and single crystalline specimens indicate that the grain boundaries provide a
strengthening effect, especially in the small strain regime. At intermediate temperature (700°C),
the γ′ precipitates size is the main creep-rate controlling parameter. In this temperature domain,
creep strength seems to be mainly controlled by dislocation motion. A very striking grain
boundary strengthening mechanism is observed at small creep strain and intermediate
temperature.

Keywords: Superalloy; Creep; Grain boundaries; γ′ precipitation

1) Introduction
Nickel-based superalloys are used for the manufacturing of several hot components of aero-
engines, such as polycrystalline alloys for turbine disks, and single crystalline alloys for blades and
vanes. Their well-known and excellent properties, especially in creep at high temperature, make
them the best candidates for these kinds of applications [1, 2].
While turbine disks are usually composed of polycrystalline microstructures due to their high
resistance to disk burst [3] and cyclic fatigue loading [4-6], single crystals are preferred for
applications such as turbine blades or vanes, thanks to their superior creep and oxidation
resistance resulting from the absence of any “deleterious” grain boundaries [1, 2].
The creep behavior of polycrystalline and single crystalline microstructures of various superalloys
has been widely studied. It appears that, in addition to the alloy chemistry, microstructural
parameters play a predominant role on the creep properties. Hence, crystalline orientation,
porosity, and the size of strengthening precipitates are the main creep controlling parameters of

© 2018 published by Elsevier. This manuscript is made available under the Elsevier user license
https://www.elsevier.com/open-access/userlicense/1.0/
single crystalline microstructures [1, 2, 7-9]. The creep behavior of polycrystalline superalloys are
mainly determined by the grain size and the γ′ size (primarily the secondary γ′ size based on
present literature) [1, 2, 10-13].
Due to the differences between polycrystalline and single crystalline superalloys, especially in
terms of chemistry and γ′ volume fraction (single crystalline alloys usually have a higher γ′ volume
fractions than polycrystalline ones), it is almost impossible to study the influence of all these
microstructural parameters independently and objectively.
In this paper, we comprehensively investigate the influence of two microstructural parameters
(grain size and γ′ size distribution) using the nickel-based superalloy AD730, a newly designed
polycrystalline superalloy developed for hot turbine parts such as turbine disks and seal rings [14-
16]. A unique aspect of this study is the possibility of examining the influence of these parameters
independently, thanks to the use of both polycrystalline and single crystalline versions of the
same alloy. Given potential industrial applications of AD730 alloy, a special attention will be paid
to the role of grain boundaries and intragranular γ′ precipitation to the first stages of creep
deformation and minimum creep rate.

2. Experimental Procedure
2.1. Material
Two polycristalline and four single crystalline microstructures have been used for this study, all
coming from the same ingot of AD730 provided by the Aubert & Duval company. Hence, all
specimens used here have the same chemical composition which is given in Table 1. The
polycrystalline microstructures were provided by Aubert & Duval as a forged bar whose diameter
is 87 mm. Blanks were cut along the longitudinal direction of the bar, at mid-radius, prior to heat
treatment. The single crystals (SX) bars have been specifically casted by SAFRAN Aircraft
Engines at Gennevilliers plant for this study, with a nearly [001] crystallographic orientation
(primary misorientation from the perfect [001] orientation less than 5 degrees).
Table 1 - Chemical composition (in wt.%) of AD730
Element Ni C Cr Mo W Al Co Ti Nb Fe B Zr
Wt. % Balance <0.02 16 3 2.7 2.3 8.5 3.5 1.1 4 0.01 0.03

In order to be able to study both the grain size and the γ′ size (and distribution) independently,
six microstructures were created. Two microstructures are polycrystalline: the first one, denoted
as Fine Grain (FG), is developed to achieve high tensile strength, low cycle fatigue, and disk burst
resistance (average grain size: 10 µm) [17]. The second one is coarse grained (CG), and is
designed for creep resistance. These microstructures are obtained by subsolvus and supersolvus
solution treatments, followed by one step or two step aging (Table 2).
The four other microstructures are single crystalline (SX). The first SX microstructure was
designed to understand the role and importance of grain boundaries in mechanical (creep)
properties. Hence, it was given the same heat treatment as the CG microstructure so that the only
difference is the presence or the absence of grain boundaries. This microstructure is denoted as
SX-CG.
The three final microstructures were used to evaluate the impact of the size and distribution of
the γ′ precipitation. It was then decided to perform a heat treatment composed of two steps: one
supersolvus solution treatment (the same for each sample) followed by different controlled
cooling rates. These microstructures are denoted as SX-40, SX-100 and SX-300, the number
designation indicating the cooling rate used after the solution treatment (in °C.min-1).
The heat treatments associated to each microstructures are detailed in the Table 2 below. It is
worth mentioning that before solution treatment 1 (see Table 2), a first homogenization heat
treatment at 1180°C for 24 hours has been applied to SX specimens to homogenize as much as
possible the chemistry across the dendritic structure of these specimens.

Table 2 - Heat treatments applied for every microstructures used for this study

Microstructural investigations have been performed using etched specimens (mirror polished
then etched with a solution made of 1/3 HNO3 + 2/3 HCl (vol. parts) at 4°C for 6 seconds). A
JEOL 7000F field emission gun scanning electron microscope (FEG-SEM) (accelerating voltage:
25 kV, working distance: 6 mm) has been used to reveal the multiscale aspect of these materials.
Figure 1 presents typical images at different magnifications for each microstructure from the
grain structure to the finest γ′ precipitation. A quantitative analysis, performed on several images
using image processing algorithms via Visilog software [18] leads to the mean characteristics
summarized in Table 3.

The FG microstructure is composed of primary γ′ precipitates with an average size of about 1


µm, and fine secondary γ′ particles (~ 40 nm diameter), while both CG and SX exhibit a bimodal
distribution of secondary γ′ precipitates, with coarse cuboidal γ′ particles having a ~300 nm size
and fine spherical γ′ with an average diameter of about 35 nm. SX-40 and SX-100 present a
bimodal γ′ precipitation: the secondary γ′ precipitates have a diameter of around 160 nm and 80
nm, while the tertiary precipitates are about 8 nm and 6 nm diameter respectively. SX-300
demonstrates a monododal distribution of γ′ precipitates, of about 35 nm.
Figure 1 - SEM observations of every microstructures used in this study; (a-c) FG
microstructure; (d-f) CG microstructure; (g-i) SX-CG microstructure; (j)(k)(l) SX-40, SX-
100 and SX-300 microstructures respectively; (a)(d)(g) Granular microstructure; (b)(e)(h)
Coarse γ′ precipitation; (c)(f)(i)(j)(k)(l) Fine γ′ precipitation
Several characteristics of these microstructures are detailed in Table 3.
Table 3 - Stereological characteristics of every microstructures used in this study (size
and volume fractions) and associated tensile properties at 700°C/5.10-3 s-1. Note that the
interparticle spacing measurements is the average distance between secondary γ′
particles extracted from size distributions.
2.2. Mechanical Testing
Mechanical tests using FG and CG samples were performed using cylindrical specimens having a
13 mm gage length, and a 4.3 mm diameter. SX samples have a 14 mm gage length and a 4 mm
diameter. Before mechanical testing, specimens were low-stress polished with SiC papers up to a
4000 grade. Monotonic creep tests in air were performed using a conventional creep device in
tension mode and under constant load. Elongation was continuously monitored using a Linear
Variable Displacement Transducer (LVDT).

The test temperatures range from 700°C to 900°C. These temperatures are high for this kind of
alloy (developed initially to withstand temperatures up to 750°C [14-16, 19]). However, these
conditions enable a better understanding of the role of the microstructural parameters and allow
a proper evaluation of the influence of oxidation on creep properties. The applied stresses range
from 50 to 900 MPa, depending on the testing temperature. Additional tensile tests were
performed at 700°C under controlled displacement mode (the strain rate chosen is 5.10-3 s-1) using
an Instron 8562 electromechanical machine. Elongation was monitored using a high temperature
extensometer.

3. Results and discussion


3.1. Impact of grain size on creep strength
The influence of the grain size on creep properties has been studied through the comparison of
the two polycrystalline microstructures. Representative creep curves are presented in Figure 2.
This figure shows that the effect of grain size on the creep strength depends on the temperature.
At 850°C, the FG microstructure presents very poor creep properties compared to the CG one
(e.g. creep life 43 times lower than CG at σ = 250 MPa). The CG creep strain rate is also much
lower than that for FG in these conditions: ε(FG) = 100*ε(CG) (Figure 3). Under these
conditions, the creep deformation mechanisms at low creep strains should be controlled by
diffusion processes, given the Norton stress exponents around 2 and 4 for FG and CG
microstructures, respectively (see Figure 3). The poor creep properties of FG microstructure may
result from a large contribution of grain boundary sliding in this temperature regime [20-22], as
suggested by the change of grain shape after creep tests (not shown here). In contrast, when T =
700°C, the contribution of the grain size to the creep properties seems to be significantly
reduced. Figure 2(d) shows that the creep life as well as the creep strain rate of both FG and CG
microstructures are very similar. Other creep results can be found in a previous article from the
authors [19]. Figure 3, which represents the minimum creep strain rate as a function of the
applied stress, reveals that at 700°C, the similarity of CG and FG microstructure creep resistance
persists over a wide range of applied stresses.

Hence, the influence of the grain size on creep properties is complex, and other microstructural
parameters must also play a key role. The secondary γ′ precipitates size is similar in FG and CG
microstructures (~ 35 / 40 nm, see Table 3), as well as the interparticle spacings between
secondary precipitates (30 / 40 nm, see Table 3). The microstructural parameter controlling the
creep deformation mechanisms (i.e. the minimum creep strain rate) in these conditions might
then be the γ′ size. The influence of this parameter will be investigated in deeper detail in section
2.3.
Figure 2 - Most representative creep curves for several conditions – Comparison between
CG, FG and SX-CG microstructures
(a) 900°C/150 MPa; (b) 900°C/250 MPa; (c) 850°C/250MPa; (d) 700°C/800 MPa

10-5
Minimum creep rate [s-1]

10-6

10-7
Change of slope

CG - 900°C SX-CG - 900°C


10-8 FG - 850°C CG - 850°C
SX-CG - 850°C FG - 700°C
CG - 700°C SX-CG - 700°C
SX-40 - 700°C SX-100 - 700°C
SX-300 - 700°C
10-9
100 Initial Applied Stress [MPa] 1 000
Figure 3 - Norton's Plot: Creep strain rate as a function of the applied stress
3.2. Contribution of grain boundaries to the creep resistance
3.2.1. Creep at high temperatures (850°C and 900°C)
The influence of the presence of grain boundaries has been evaluated using ~ <001> AD730
single crystalline specimens heat treated exactly as the CG microstructure. Figure 2 presents
several creep curves allowing the comparison of SX-CG and CG microstructures, for several
temperatures. Creep tests performed at 900°C on CG and SX-CG microstructures indicate that
depending on the applied stress, and hence testing time, the single crystalline alloy does not
necessarily present the best creep properties. At low applied stress (150 MPa), SX-CG presents a
longer creep life (creep life of the SX specimen is ~ 2.5 times that of the polycristalline CG).
However, despite a longer creep life, the SX specimen also presents a higher minimum creep
strain rate compared to the CG microstructure (see fig. 2(a)). When the applied stress is higher
(250 MPa), SX-CG creep life is lower than the polycristalline microstructure. As for σ = 150
MPa, SX-CG creep strain rate is still higher. When the temperature decreases (T = 850°C), the
creep properties of both CG and SX-CG microstructures are getting closer (Figure 2(c)): the
creep life and the minimum creep strain rates are very similar, despite the minimum creep rate of
SX-CG still remain higher than the CG one in the investigated stress range at 850°C (see Fig. 3).

Comparisons of the creep strain rate evolutions of the single crystalline microstructure with the
coarse grained one having the same intragranular γ′ precipitation (Figure 4) indicate that the
evolution of the strain during the test is strongly linked with the presence or the absence of grain
boundaries. More specifically, a strong decrease of the creep strain rate is noticed at the beginning
of the tests for CG microstructure (900°C and 850°C). However, an almost continuous increase
of the creep strain rate is observed for SX-CG, from the beginning of the test to the end.

Figure 4 - Creep strain rate as a function of the creep strain for the creep tests performed at 900°C
(a) and 850°C (b); Comparison between SX-CG and CG microstructures

This result clearly demonstrates the importance of the grain boundaries, which can be considered
under these conditions as providing a mechanical strengthening. Indeed, in the polycrystalline
version of AD730, the work hardening during the first stages of creep at both 900°C and 850°C
(see Figure 4(a) and Figure 4(b) respectively) seems to be enhanced by the presence of grain
boundaries. This explains the decrease in creep strain rate. In the single crystalline version, the
absence of grain boundaries allows the plastic deformation to develop in the entire volume of the
tested sample without encountering significant barriers or variation in stress state. This beneficial
effect of grain boundaries to creep properties is not observed in typical Ni-based single crystalline
superalloys used for blade applications since most of the strengthening in such alloys comes from
the high volume fraction of γ′ precipitates [1, 23-25]. Grain boundaries, if present in such alloys,
would be “weak links” in a strong microstructure and would lead directly to intergranular crack
initiation, as shown recently by Stinville et al. [26] and Mataveli Suave et al. [27] during transverse
creep testing of a high γ′ volume fraction Ni-based superalloys in a bi-crystalline and directionally
condition, respectively.

The better performance of the SX-CG microstructure in comparison to the CG one at


900°C/150 MPa can be explained through the observation of the longitudinal cut of the CG
specimen (Figure 5). These SEM observations show the impact of oxidation during a creep test at
high temperature. Several secondary cracks along the gage length indicate that the grain
boundaries are clearly the weakest points regarding oxidation. In these harsh conditions, the
single crystalline microstructure performs better than the polycrystalline one in terms of creep life
due to the absence of any grain boundaries exposed to oxidation, even if its creep strain rate is
higher.

Figure 5 - SEM observation of a longitudinal cut of CG crept specimen at 900°C –


Evidences of the influence of oxidation

3.2.2. Creep at 700°C


At intermediate temperatures (700°C) (see Figure 2(d) and Figure 3), SX-CG demonstrates
inferior creep properties when compared to the two polycrystalline microstructures (CG and
FG). Despite the fact that at this temperature the γ′ volume fraction is about 36 % [28], i.e. higher
than for the previous creep conditions at 850°C and 900°C at which the γ′ volume fraction is 32
% and 27 % respectively, SX-CG always exhibits a faster strain rate and shorter creep life. This is
more clearly illustrated in Figure 6 showing the evolution of the creep strain rate as a function of
the creep strain. It is hence shown that under these creep conditions, the γ′ precipitation is not
the only contributing parameter to the creep strength and here again, grain boundaries are
beneficial to the creep strength, especially at low creep strains below 1%. According to Figure 3,
Norton exponents higher than 16 are observed for all microstructures, indicating that the
minimum creep rate is be controlled by diffusion processes. The creep rate should be controlled
in these conditions, at least partly, by the dislocation activity. This assumption is also supported
by the fact that γ′ precipitates coarsening at 700°C is very slow at 700°C and hardly visible within
the timeframe of our experiments [29]. Hence, it is highly suspected that grain boundaries are
obstacles to the dislocation movement and they contribute in a spectacular way at 700°C to the
creep rate decrease at the beginning of the tests.

Figure 6 - Creep strain rate as a function of the creep strain for the creep tests performed
at 700°C/800 MPa; Comparison between SX-CG and CG microstructures
One last point to consider here is the fact that the γ′ precipitation is not homogeneous in the
single crystalline microstructures, due to the dendritic solidification (Figure 1(g)). Despite long
homogenization heat treatments having been applied to the single crystalline specimens (24h at
1180°C, followed by two subsequent solution treatments, see Table 2), these heterogeneities still
remain. The γ′ volume fraction is then slightly lower in the primary dendritic arms compared to
the interdendritic areas. Dendritic areas are consequently expected to deform more rapidly than
the interdendritic zones. At high temperature (850°C and above), the coalescence of the γ′
precipitates is significant (Figure 7). The precipitates coarsening is responsible for the progressive
increase of the creep strain rate during almost all the creep test, as shown in Fig. 4. The
heterogeneities between dendritic and interdendritic areas are then progressively erased as the
creep duration and temperature increase, leading to almost similar creep strengths between SX-
CG and CG microstructures. This γ′ microstructure evolution helps to explain their similar creep
properties. At 700°C, there is almost no evolution of the γ′ precipitates [29]. Thus, the
heterogeneities mentioned earlier, which remain in these conditions, certainly contribute to the
poor creep properties of SX-CG compared to CG microstructure.
To conclude, there is a strong competition between the strengthening mechanisms due to the
grain boundaries, the intragranular γ′ precipitates, and the impact of oxidation at grain
boundaries. Depending on the applied temperature and stress conditions, the absence of grain
boundaries may be beneficial or deleterious.
Figure 7 - γ′ precipitation after creep tests on SX-CG microstructure; (a)(b) 700°C
(exposure time: 950 h) ; (c) 850°C (550 h) ; (d) 900°C (650 h)

3.3. γ′ precipitates size and distribution


The previous results suggest that at 700°C, the grain size is not the main creep controlling
parameter. According to previous studies [11, 13, 30-32], one creep controlling parameter in
these conditions could be the γ′ precipitates size and distribution. However, in these studies, the
γ′ precipitates size and distribution is not the only variable parameter. The grain boundaries are
indeed playing a complex role in creep properties, by providing a strengthening effect and by
modifying the intragranular γ′ volume fraction (depending on the heat treatment, a population of
the γ′ precipitates may be localized along the grain boundaries). In the present study, the effect of
γ′ size/distribution has been analyzed by using three single crystalline microstructures SX-40, SX-
100 and SX-300 (Figure 1, Table 2 and Table 3). These microstructures allow the investigation of
the influence of the γ′ precipitates size independently from the grain size or the presence of grain
boundaries.
The results of the creep tests performed on these microstructures are presented in Figure 8 and
their minimum strain rates as a function of the applied stress are also added in the Norton
diagram in Figure 3.
Figure 8 - Creep curves at 700°C/700 MPa and 750 MPa using SX-40, SX-100 and SX-300
microstructures
These curves demonstrate that at 700°C, there is a strong influence of the γ′ size on the creep
properties. Indeed, the creep life of SX-40 (which has the largest secondary γ′ precipitates
(average secondary γ′ size of ~ 160 nm)) is 14 times lower than the creep life of SX-300 (average
secondary γ′ size of ~ 35 nm). Moreover, a completely different behavior is observed between
these two microstructures: the creep strain of SX-40 increases continuously from the beginning
of the test and exhibits very little strain hardening compared to the two other considered SX
microstructures, as observed in Fig. 9. In comparison, creep strain of the SX-300 microstructure
remains very low during most of the test (below 0.5 % up to 80 % of the creep life), and
increases dramatically only at the end of the test. This behavior applies for the two investigated
applied stresses.
The performance of SX-100 is different from the two other SX microstructures; its creep
behavior is highly dependent to the applied stress. When σ = 750 MPa, SX-100 behavior is
similar to SX-40, while it is close to SX-300 when σ = 700 MPa. In either case, the influence of
the γ' size is very strong in these conditions. These creep tests confirm that at 700°C, this
microstructural parameter is the creep controlling one in AD730.

The analysis of the creep strain rates during the creep tests performed on SX-40, SX-100 and SX-
300 suggests that the poor creep performance of SX-40 is linked to an absence of strong
strengthening during the first stage of creep tests. This phenomenon has already been noticed for
the other single crystalline microstructure SX-CG (Figure 4). However, SX-300 presents a
significant decrease of its creep strain rate at the beginning of the test. Hence, the strengthening
effect of the (very) fine γ′ precipitates (associated with the small interparticle spacings between
secondary γ′ precipitates) is prominent for this precipitate structure, and helps to compensate for
the absence grain boundaries in terms of the early stages of hardening.
Figure 9 - Creep strain rate as a function of the creep strain; (a) 700°C/750 MPa; (b)
700°C/700 MPa
The creep strain rate of the SX-100 microstructure for σ = 750 MPa is similar to that of SX-40
(small decrease of the creep strain rate at the beginning of the test), while it is very similar to that
of SX-300 when σ = 700 MPa. The variation of the applied stress might induce a change in the
creep deformation mechanism for this microstructure, which requires additional investigation.
Observations of longitudinal cuts of crept samples allow the investigation of the shape of the
fracture surface. As shown in Figure 10, fracture surfaces of the SX-40 specimens are very planar,
whereas it is much more tortuous for the finest microstructure (SX-300). This observation is true
for both microstructures at both 700 and 750 MPa. However, a change in fracture mode is
noticed for SX-100 specimens: the fracture surface of the sample tested at 700°C/750 MPa is
very planar/crystallographic with one of the two fracture planes close to {111} plane (Figure
10(b)), while it is more tortuous when the applied stress decreases (700 MPa, Figure 10(e)).

Figure 10 - Longitudinal cuts of crept samples after rupture; Back Scattered imaging
mode; (a)(b)(c) 700°C/750 MPa; (d)(e)(f) 700°C/700 MPa; (a)(d) SX-40; (b)(e) SX-100;
(c)(e) SX-300
General features of the deformation processes have been analyzed using SEM imaging on these
longitudinal cuts (Figure 11). An evident shearing of the γ′ precipitates is noticed for all the
microstructures, but a higher density of slip bands (see arrows in Fig. 11) are observed in the SX-
40 microstructure compared to SX-100 and SX-300. This might indicate that the differences in
creep properties between the three microstructures comes from a change in deformation
mechanism, or at least in the intensity of localized shearing that is possible for the coarser
precipitate microstructure. Finer length-scale analysis of the differences in deformation
mechanisms will be the subject of future investigation.

Figure 11 - Microstructure of the crept samples after rupture - 700 °C/700 MPa; (a) SX-40;
(b) SX-100; (c) SX-300

An evaluation of the critical resolved shear stress for precipitate cutting and of the Orowan shear
stress has been made by using models presented by Galindo-Nava et al. [33]. These authors have
developed a model taking into account a multimodal γ′ precipitate population. A dislocation can
penetrate a field of γ′ precipitates by either bypassing (the so-called Orowan bypassing
mechanism), or cutting the precipitates. In the latter case, the dislocations that are dissociated
forming an antiphase boundary (APB) can cut the γ′ precipitate via a weak pair-coupled or strong
pair-coupled mode. The Galindo-Nava model presents a unified equation accounting for both
weak and strong pair-coupled cases. The critical shear stress τp for particle shearing is described
in Eq. 1:

γ APB  s
l1
t
l1 
 Equation 1
τp =  ws
+ wt
Λ1 + 2 rs Λ1 + 2 r t 
2b s t

where γAPB is the antiphase boundary energy, b is the magnitude of the Burgers vector, ri (i = s or
t, for secondary or tertiary γ′ precipitate) is the mean radius of secondary and tertiary γ′
precipitate, wi is the particle number fraction of secondary and tertiary γ′ precipitate, l1i is the
segment length of the leading dislocations acting in the cutting of a particle and Λ1i the effective
length of the leading dislocation driving particle cutting.
The equation used to calculate the Orowan shear stress τOro comes from Reppich’s pioneering
works [34-37]:

2 µb Equation 2
τOro = ws 3 Ls

where µ is the shear modulus, and Ls is the mean particle spacing.


The interested reader is referred to [33] for additional details regarding this model.
The raw data used for these calculations are presented in Table 4 below.
Table 4 - Data used in the calculation of critical stresses
Antiphase boundary energy γAPB [J.m2] 0.265
Magnitude of the Burgers vector [nm] [17] 0.254
Shear modulus µ [GPa] [17] 76.3

The result of the calculations are presented in Figure 12. According to this diagram, the
deformation rate limiting mechanism for SX-300 under the tested creep conditions is the cutting
of the γ′ precipitates. On the other hand, regarding SX-40, both Orowan bypassing and cutting of
the γ′ precipitates seem to be possible. However, these calculation indicate that the main creep
rate controlling mechanism seems to be Orowan bypassing due to a lower resolved shear stress
required to activate this deformation mode. This is consistent with the highest creep strain rate
noticed for SX-40 compared to SX-300. It is also noted that for SX-40 and SX-100
microstructures, even if Orowan bypassing seems to be the rate controlling deformation
mechanism, these calculations suggest that γ′ cutting is also possible. This is consistent with the
high number of shearing bands observed on the longitudinal cuts of crept specimens (see Figure
11).

Figure 12 - Critical Resolved Shear and Orowan stresses as a function of the precipitate
diameter – The resolved shear stresses corresponding to the applied stresses have been
reported on this diagram
Deformation substructure analysis by TEM are needed to better understand the differences in
deformation mechanisms between the three microstructures and to confirm (or not) the
assumptions presented. For instance, there is substantial evidence that shearing of precipitates
through the formation of stacking faults and microtwins [38-44] is also important in Ni base
superalloys under creep conditions, so that the assumption of APB shearing made in model of
Galindo-Nava et al. [31] has yet to be validated. TEM observations would also provide insight to
the apparent change in creep behavior of SX-100 specimen when the applied stress decreases
from 750 MPa to 700 MPa at 700°C.

4. Conclusions
In this study, the influence of three microstructural parameters (grain size, the presence of grain
boundaries, and the γ′ size distribution) on the creep resistance, especially on the minimum creep
rate, has been comprehensively studied using the nickel-based superalloy AD730. The following
main conclusions have been established:
• Creep tests performed using two polycristalline microstructures have shown that the grain
size is not always the main creep controlling parameter: If it is clearly controlling the
creep properties at 850°C, creep tests at 700°C showed that both FG and CG
microstructures yield similar properties. Hence, there is no optimal grain size that can
provide maximized creep properties over a wide range of temperature and applied
stresses.
• The most striking result of this study is that grain boundaries contribute to a notable
strengthening of AD730 alloy, especially at 700°C.
• It has been demonstrated unambiguously using SX specimens having the same γ′
precipitates distribution as the polycrystalline coarse grain reference alloy that with
increasing the temperature (up to 850°C and more), a strong competition between grain
boundary strengthening and intergranular oxidation occurs.
• The use of single crystals has also allowed an objective investigation of the influence of
the secondary γ′ size and distribution. Creep results clearly indicate that for the finest
precipitate populations, superior creep properties (i.e. lower minimum strain rate) are
obtained. This is potentially linked to a transition in creep deformation mechanisms,
which has been evaluated with the aide of models developed by Galindo-Nava et al. [33].
• Finally, the present study clearly shows that single crystalline specimens are not always
stronger in creep compared to their polycrystalline counterparts. Indeed, a remarkable
improvement of creep strength at smaller creep strain (<2%) is observed for
polycrystalline samples, even when tested up to 900°C.

Acknowledgments
The authors would like to thank Aubert & Duval for providing the materials, for L.T. PhD grant
and for their continuous interest in this study. SAFRAN SA is gratefully acknowledged for
financial support, for the continuous collaboration for over 10 years with Institut Pprime (UPR
CNRS 3346) on Ni-based superalloys activities and for casting the single crystals of AD730TM. A
part of this work has been carried out in the frame of the OPALE industrial chair, co-funded by
the Safran group and the French ANR (grant number ANR-14-CHIN-0002). MJM would like to
acknowledge the support of the National Science Foundation DMREF program under grant
#1534826.

References
[1] T.M. Pollock, S. Tin, Journal of Propulsion and Power, 22 (2006) 361-374.
[2] R.C. Reed, The Superalloys - Fundamentals and Applications, Cambridge University Press,
Cambridge, UK, 2006.
[3] M. Mazière, J. Besson, S. Forest, B. Tanguy, H. Chalons, F. Vogel, European Journal of
Mechanics A/Solids, 28 (2009) 428-432.
[4] M. Clavel, A. Pineau, Materials Science and Engineering, 55 (1982) 157-171.
[5] M. Clavel, A. Pineau, Materials Science and Engineering, 55 (1982) 173-180.
[6] Andre Pineau, Stephen D Antolovich, Engineering Failure Analysis, 16 (2009) 2668-2697.
[7] P. Caron, T. Khan, Materials Science and Engineering, 61 (1983) 173-194.
[8] P. Caron, Y. Ohta, Y.G. Nakagawa, T. Khan, in: S. Reichman, D.N. Duhl, G. Maurer, S.
Antolovich, C. Lund (Eds.) Superalloys 1988, TMS, Seven Springs, PA, USA, 1988, pp. 215-224.
[9] Pierre Caron, Tasadduq Khan, Aerospace Science and Technology, 3 (1999) 513-523.
[10] Kr Bain, Ml Gambone, Jm Hyzak, Mc Thomas, in: S. Reichman, D.N. Duhl, G. Maurer, S.
Antolovich, C. Lund (Eds.) Superalloys 1988, TMS, Seven Springs, PA, USA, 1988, pp. 13-22.
[11] P.R. Bhowal, E.F. Wright, E.L. Raymond, Metallurgical Transactions A, 21 (1990) 1709-
1717.
[12] F. Torster, G. Baumeister, J. Albrecht, G. Lütjering, D. Helm, M.A. Daeubler, Materials
Science and Engineering: A, 234 (1997) 189-192.
[13] M.P. Jackson, R.C. Reed, Material Science and Engineering A, 259 (1999) 85-97.
[14] A. Devaux, E. Georges, P. Heritier, in: E.A. Ott, J.R. Groh, A. Banik, I. Dempster, T.P.
Gabb, R. Helmink, X. Liu, A. Mitchell, G. Sjöberg, A. Wusatowska-Sarnek (Eds.) 7th
International Symposium on Superalloy 718 and Derivatives, TMS, Pittsburgh, PA, USA, 2010,
pp. 223-235.
[15] A. Devaux, E. Georges, P. Heritier, Advanced Materials Research, 278 (2011) 405-410.
[16] A. Devaux, B. Picqué, M.F. Gervais, E. Georges, T. Poulain, P. Héritier, in: E.S. Huron, R.C.
Reed, M.C. Hardy, M.J. Mills, R.E. Montero, P.D. Portella, J. Telesman (Eds.) Superalloys 2012,
TMS, Seven Springs, Champion, PA, USA, 2012, pp. 911-919.
[17] A. Devaux, A. Helstroffer, J. Cormier, P. Villechaise, J. Douin, M. Hantcherli, F. Pettinari-
Sturmel, in: E. Ott, A. Banik, X. Liu, I. Dempster, K. Heck, J. Andersson, J. Groh, T. Gabb, R.
Helmink, A. Wusatowska-Sarnek (Eds.) 8th International Symposium on Superalloy 718 and
Derivatives, TMS, Pittsburgh, PA, USA, 2014, pp. 521-535.
[18] J.R. Vaunois, J. Cormier, P. Villechaise, A. Devaux, B. Flageolet, in: E.A. Ott, J.R. Groh, A.
Banik, I. Dempster, T.P. Gabb, R. Helmink, X. Liu, A. Mitchell, G. Sjöberg, A. Wusatowska-
Sarnek (Eds.) 7th International Symposium on Superalloy 718 and Derivatives, TMS, Pittsburgh,
PA, USA, 2010, pp. 199-213.
[19] L. Thébaud, P. Villechaise, J. Cormier, F. Hamon, C. Crozet, A. Devaux, J.-M. Franchet, A.-
L. Rouffié, A. Organista, in: M. Hardy, U. Glatzel, B. Griffin, B. Lewis, C. Rae, V. Seetharaman,
S. Tin, E. Huron (Eds.) Superalloys 2016, TMS, Seven Springs, Champion, PA, USA, 2016, pp.
877-886.
[20] I.M. Lifshitz, Soviet Physics JETP, 17 (1963) 909-920.
[21] T.G. Langdon, Journal of Materials Science, 41 (2006) 597-609.
[22] A. Wisniewski, J. Beddoes, Materials Science and Engineering, A 510-511 (2009) 266-272.
[23] T. Murakumo, T. Kobayashi, Y. Koizumi, H. Harada, Acta Materialia, 52 (2004) 3737-3744.
[24] T. M. Pollock, A. S. Argon, Acta Metallurgica Materialia, 40 (1992) 1-30.
[25] M. Bensch, C.H. Konrad, E. Fleischmann, C.M.F. Rae, U. Glatzel, Materials Science and
Engineering: A, 577 (2013) 179-188.
[26] J.C. Stinville, K. Gallup, T.M. Pollock, Metallurgical and Materials Transactions A, 46A
(2015) 2516-2529.
[27] L. Mataveli Suave, J. Cormier, P. Villechaise, D. Bertheau, G. Benoit, F. Mauget, G.
Cailletaud, L. Marcin, in: M. Hardy, U. Glatzel, B. Griffin, B. Lewis, C. Rae, V. Seetharaman, S.
Tin, E. Huron (Eds.) Superalloys 2016, TMS, Seven Springs, Champion, PA, USA, 2016, pp. 747-
756.
[28] A. Devaux, L. Berglin, L. Thébaud, R. Delattre, C. Crozet, O. Nodin, in: MATEC Web of
Conferences, EDP Sciences, 2014, pp. 01004.
[29] F. Masoumi, M. Jahazi, D. Shahriari, J. Cormier, Journal of Alloys and Compounds, 658
(2016) 981-995.
[30] D. Locq, M. Marty, P. Caron, in: T.M. Pollock, R.D. Kissinger, R.R. Bowman, K.A. Green,
M. Mclean, S. Olson, J.J. Schirra (Eds.) Superalloys 2000, TMS, Seven Springs, Champion, PA,
USA, 2000.
[31] D. Locq, P. Caron, S. Raujol, F. Pettinari-Sturmel, A. Coujou, N. Clément, in: K.A. Green,
T.M. Pollock, H. Harada, T.E. Howson, R.C. Reed, J.J. Schirra, S. Walston (Eds.) Superalloys
2004, TMS, Seven Springs, Champion, PA, USA, 2004, pp. 179-187.
[32] A. Laurence, J. Cormier, P. Villechaise, T. Billot, J.-M. Franchet, F. Pettinari-Sturmel, M.
Hantcherli, F. Mompiou, A. Wessman, in: E. Ott, A. Banik, X. Liu, I. Dempster, K. Heck, J.
Andersson, J. Groh, T. Gabb, R. Helmink, A. Wusatowska-Sarnek (Eds.) 8th International
Symposium on Superalloy 718 and Derivatives, TMS, Pittsburgh, PA, USA, 2014, pp. 333-348.
[33] E.I. Galindo-Nava, L.D. Connor, C.M.F. Rae, Acta Materialia, 98 (2015) 377-390.
[34] W. Huther, B. Reppich, Zeitschrift fur Metallkunde, 69 (1978) 628-634.
[35] B. Reppich, Acta Metallurgica, 30 (1982) 87-94.
[36] B. Reppich, Particle strengthening, Wiley ‐ VCH Verlag GmbH & Co. KGaA, 1993.
[37] S. Steuer, Z. Hervier, S. Thabart, C. Castaing, T.M. Pollock, J. Cormier, Material Science and
Engineering A, 601 (2014) 145-152.
[38] M.G. Ardakani, M. Mclean, B. Shollock, Acta materialia, 47 (1999) 2593-2602.
[39] D.M. Knowles, Q.Z. Chen, Materials Science and Engineering, A340 (2003) 88-102.
[40] G.B. Viswanathan, P.M. Sarosi, M.F. Henry, D.D. Whittis, W.W. Milligan, M.J. Mills, Acta
Materialia, 53 (2005) 3041-3057.
[41] A. Guimier, J.L. Strudel, in: Second International Conference on the Strength of Metals and
Alloys, American Society for Metals, Pacific Grove, California, USA, 1970, pp. 1145-1149.
[42] B. Décamps, A.J. Morton, M. Condat, Philosophical Magazine A, 64 (1991) 641-668.
[43] W.W. Milligan, S.D. Antolovich, Metallurgical and Materials Transactions A, 22 (1991) 2309-
2318.
[44] T.M. Smith, R.R. Unocic, H. Deutchman, M.J. Mills, Materials at High Temperatures, 33
(2016) 372-383.

You might also like