You are on page 1of 8

Materials Science & Engineering A 696 (2017) 323–330

Contents lists available at ScienceDirect

Materials Science & Engineering A


journal homepage: www.elsevier.com/locate/msea

Temperature dependence of tensile properties and deformation behaviors of MARK


nickel-base superalloy M951G

Luqing Cuia,b, Huhu Sua,b, Jinjiang Yua, , Jinlai Liua, Tao Jina, Xiaofeng Suna
a
Institute of Metal Research, Chinese Academy of Sciences, Shenyang 110016, China
b
School of Materials Science and Engineering, University of Science and Technology of China, Hefei 230026, China

A R T I C L E I N F O A B S T R A C T

Keywords: Tensile behaviors of M951G alloy have been studied from 20 °C to 1100 °C. It shows that the higher W and Hf
M951G alloy additions enhance the tensile strength of M951G alloy greatly compared to M951 alloy from RT to 800 °C, which
Tensile behaviors has been analyzed from the points of deformation microstructures, the strong effect of solid solution
Continuous stacking faults strengthening and precipitation strengthening in detail. The continuous stacking faults present in the
Microstructures
microstructures at low temperatures, which have a significant influence on work hardening and rarely been
reported in the traditional casting superalloy previously. The primary deformation mechanisms include shearing
of the γ′ precipitates by a/2〈110〉 dislocations below 600 °C and by-passing of γ′ precipitates by a/2〈110〉 matrix
dislocations above 900 °C. At intermediate temperatures, it shows a transition from shearing to by-passing. At
last, the relationship between tensile properties and microstructures has been reasonably explained.

1. Introduction lifetime is increased from 15 to 120 h [11]. The results from Yang et al.
show that with the increase of W additions, the stacking fault energy of
Nickel-base superalloys are unique high temperature materials used γ phase significantly decreases, and the steady state creep rate reduces
in gas turbine engines [1,2], which are due to their excellent resistance simultaneously [12]. In GH4586 alloy, when the amount of W increases
to mechanical and chemical degradation [2]. The outstanding proper- from 2% to 5%, the yield strength is increased from 860 to 910 MPa
ties are mainly owing to their special microstructures, which consist of when tensile at 800 °C [13].
ordered γ′ precipitates with L12 structure embedded in a fcc (face Tensile tests of many superalloys have been carried out [14], while
centered cubic) nickel-base solid solution matrix [3]. the influence of the transition elements and temperature on their
Tensile test is a basic test for many metal materials and the results outstanding properties are still unclear. In order to further study the
are important reference for other experimental tests. In Ni-base super- effects of W and Hf on the mechanical property, we designed a new
alloys, the three main deformation mechanisms during tensile test are nickel-base superalloy, which is based on M951 [14] by adding 3% W
dislocation shearing γ′ precipitates, dislocation climbing and disloca- and 1.5% Hf additionally, and named as M951G. Therefore, in the
tion by-passing γ′ precipitates via Orowan loop [4–6]. When γ′ present work, we purpose to explore the tensile deformation mechan-
precipitates sheared by dislocations during tensile deformation, differ- isms at different temperatures and analyze the effects of W and Hf on
ent dislocation configurations will be generated such as anti-phase tensile property.
boundary (APB) and stacking fault [5,6]. Deformation microtwin [7]
and complex reaction between dislocations [8] also can be observed 2. Experimental procedure
during the tensile test.
Transition elements (such as W, Mo, Hf, Co, Cr etc.) are the most The nominal chemical compositions of M951G alloy are listed in the
important solid solution elements in superalloy, which can greatly Table 1. The master alloy was remelted into the tensile samples in a
improve the mechanical properties of the alloys [9,10]. Among these vacuum induction furnace. The mold preheating and pouring tempera-
elements, Hf and W are of special interest as they are known to enhance ture were 900 and 1450 °C, respectively. The samples were heat treated
transverse creep properties and reduce the stacking fault energy (SFE) by 1210 °C×4 h, AC→1100 °C×4 h, AC→870 °C×24 h, AC (AC: air
of γ phase. Adding about 2% Hf to MAR-M200 alloy, the elongation is cooling).
increased from 1.6% to 9.8% when creep at 870 °C/420 MPa and the Then, the metallographic sample was mechanically polished and


Corresponding author.
E-mail address: jjyu@imr.ac.cn (J. Yu).

http://dx.doi.org/10.1016/j.msea.2017.04.065
Received 15 December 2016; Received in revised form 12 April 2017; Accepted 14 April 2017
Available online 14 April 2017
0921-5093/ © 2017 Published by Elsevier B.V.
L. Cui et al. Materials Science & Engineering A 696 (2017) 323–330

Table 1 700 °C and then rapidly decreases to zero at 900 °C. Above 1000 °C, the
Nominal compositions and γ′ content of M951 [14] and M951G alloys (wt%). value of n is below zero, indicating the trend of the recovery softening is
larger than that of work hardening. The yield strength (YS) of M951G
Alloys W Mo Nb Co Cr Al Hf others Ni Vγ′ (%)
alloy progressively decreases from 20 °C to 600 °C, then gradually
M951 3.5 3 2.2 5 9 6 0 0.091 Bal 49.4 increases with the increase of temperature and reaches a peak at 800 °C.
M951G 6.5 3 2.2 5 9 6 1.5 0.091 Bal 54 After that, it drops sharply. The ultimate tensile strength (UTS) of
M951G almost remains the same from room temperature to 600 °C,
then increases with the temperature increasing and reaches its max-
etched in a solution of 20 g CuSO4+100 ml HCl+5 ml H2SO4+80 ml
imum at 700 °C. The UTS-temperature curve has the same trend above
H2O. The microstructure was analyzed using a JMS-6301F field-
700 °C as that of YS-temperature curve above 800 °C. The elongation
emission scanning electron microscope (SEM). The volume fraction
(EL) slightly increases from room temperature to 600 °C, then drops
and the size of γ′ precipitates were measured and calculated by the
quickly to the minimum at 700 °C corresponding to the maximum of
Image-Pro Plus software. The partitioning behavior of alloy elements
UTS, after that it increases rapidly with temperature until 1100 °C. The
after heat treatment between the γ and γ′ phases in M951G alloy
trend of the reduction of area (R/A) is similar to that of the elongation.
determined by a CAMECA SX-100 Electron Probe Microanalysis
It also can be observed that the yield strength of M951G alloy is much
(EPMA).
higher than M951 alloy below 900 °C, as shown in Fig. 2c.
Tensile specimens with a diameter of 5 mm and a gauge length of
25 mm were machined. The tensile tests were carried out on an AG-
3.3. Deformation microstructures
250KNE mechanical test machine from room temperature to 1100 °C
with a constant strain rate of 0.75 mm/min to rupture. During the test,
3.3.1. Low temperatures
the temperature variation was maintained within ± 2 °C. Two identical
Similar deformation microstructures are observed at low tempera-
specimens were tested at each temperature.
tures (blow 600 °C), which are consist of tangled dislocations in the γ
After the tensile test, the discs for the transmission electron
channel and shearing of γ′ precipitates by pairs of a/2〈101〉 matrix
microcopy (TEM) observation were cut perpendicularly to the loading
dislocations (Fig. 3a and c) and continuous stacking faults (Fig. 3b and
axis and 5 mm away from the fracture surfaces. The discs were thinned
d). The continuous stacking faults can be occasionally observed in the
to 50 µm, and then subjected to twin-jet polishing in a solution of
single crystal superalloys [15], however, to the best of our knowledge,
ethanol with 10 vol% perchloric acid at −25 °C and 20 V. The
this type of dislocation configuration has rarely been reported in the
microstructures of all specimens were examined using a JEM 2100
traditional casting superalloy previously. The formation of the contin-
TEM operating at 200 kV.
uous stacking faults in M951G alloy will be discussed in Section 4.1.1.
At 20 °C, the dislocations distribute unevenly throughout the specimen
3. Results and a large number of dislocations tangle in the γ matrix channel,
which lead to localized strain. This is in good agreement with the
3.1. Microstructures before tensile test relatively low tensile plasticity as shown in Fig. 2d. At 600 °C, an
interesting and special dislocation configuration is observed, as shown
The microstructures and the size distribution of the γ′ precipitates in Fig. 3c. Several obvious characteristics can be observed in the
for as-cast and heat treated M951G alloys are shown in Fig. 1. The γ′ dislocation configuration: (i) the line directions of these dislocations
precipitates of the as-cast sample display irregular morphology com- usually have 90° angles reorientation, indicating the occurrence of
pared to the heat-treated samples, as shown in Fig. 1a and b. The as-cast cross-slip; (ii) the dislocations are very long in the TEM specimen,
M951G alloy contains about 60% volume fraction of the γ′ phase and a which indicates that these dislocations are in the (001) plane. (iii) Most
mean size of the γ′ precipitates of 0.48 µm. After standard heat of the dislocations with Burgers vector b=1/2[101]. These dislocations
treatment, the microstructure consists of cuboidal γ′ precipitates with characteristics show that this type of dislocation configuration may
the same average size as the as-cast one and spherical γ′ particles of form by the cross-slip of a/2〈101〉 matrix dislocations on {111} slip
65–75 nm mean diameter. The total volume percent of γ′ phase is about planes in the horizontal γ matrix.
54%, smaller than that of as-cast condition. The high annealing
temperature (aged at 1100 °C) may be responsible for the partial 3.3.2. Intermediate temperatures
dissolution of the γ′ phase, since the temperature range of the γ′ At intermediate temperatures, it shows a transition of deformation
precipitates solution is 1136 °C to 1200 °C. The size distribution of the mechanism from shearing to by-passing during tensile deformation.
γ′ precipitates in heat treated samples is more homogeneously than the Pairs of a/2〈101〉 dislocations and stacking faults cutting into γ′
as-cast ones, as shown in Fig. 1a and b, which is also supported by the precipitates can be seen, but no continuous stacking fault present at
FWHM in Fig. 1c and d. intermediate temperatures. Some dislocation loops also can be occa-
sionally observed, indicating the bypassing of the γ′ particles by
3.2. Tensile behaviors dislocations occurs.
At 700 °C, a number of slip bands present in the microstructure (as
The tensile results of M951G alloy are illustrated in Fig. 2. Fig. 2a shown in Fig. 4a), consisting of high density of tangled dislocations,
shows the true stress-strain curves of M951G alloy at different which lead to considerable strain localization. Interaction between
temperatures. At 20 °C, 600 °C and 700 °C, an obvious work hardening dislocations from different slip systems results in significant work
takes place after the yield point. At 760 °C and 800 °C, only a weak hardening, which is in consistent with the maximum of the strain-
work hardening is observed after the yield point, then the stress hardening exponent at 700 °C, as shown in Fig. 2b. Unlike the
increases little by little with the temperature increasing until rupture. dislocation configurations at 700 °C, the density of stacking faults at
At 900 °C, the work hardening and recovery softening achieve a state of 800 °C is much higher, as shown in Fig. 4b. These stacking faults have
equilibrium, so the stress maintains a constant until rupture. The flow been studied by many researchers [4]. Caron et al. [16] proposes a
stress at 1000 °C and 1100 °C gradually decreases after yield point until mechanism that these stacking faults form by the single perfect matrix
rupture. The strain-hardening exponent (n), tensile strength, and dislocation decomposition in the following way:
plasticity of M951G alloy are obtained from the tensile curves. The
a/2 < 101 > → a/3 < 211 > + a/6 < 121 > + SF inγ′ (1)
strain-hardening exponents of M951G alloy at different temperatures
are shown in Fig. 2b. It is evident that n value increases to maximum at It is generally believed that the a/3〈211〉 leading dislocation shears

324
L. Cui et al. Materials Science & Engineering A 696 (2017) 323–330

Fig. 1. The SEM images and the size distribution of the γ′ precipitates of M951G alloy. (a) As-cast sample. (b) Heat treated sample. (c)(d) The average edge length and FWHM of the γ′
precipitates in as-cast and heat treated samples, respectively.

into the γ′ precipitates leaving a stacking fault behind it, and the a/ 〈100〉 direction.
6〈121〉 Shockley dislocation remains at the γ/γ′ interface. Additionally, The configurations of the dislocations after tensile tests at 1000 °C
different extended directions of the stacking faults (marked by “I” and and 1100 °C are similar. At 1100 °C, it can be observed that the
“II”) in the same γ′ precipitate are also observed in Fig. 4c, which dislocation networks in the Fig. 5d is more regular and denser than
certainly indicates that dislocations with different slip systems are those in Fig. 5b and Fig. 5c. The dislocation networks can be classified
activated. Complex reactions have occurred between stacking faults “I” into two types, i.e. hexagonal (marked by “I”) and tetragonal (marked
and “II”, and a sessile dislocation has formed, the formation process will by “II”) networks. The tetragonal networks are the same in Fig. 5b. The
be discussed in Section 4.1.3. line directions of some dislocations in hexagonal networks approxi-
mately orientate in the γ/γ′ misfit. It is worth to note that the number of
3.3.3. High temperatures the dislocations in the γ′ precipitates at 1100 °C is smaller than that at
Different to the deformation mainly via shearing of γ′ precipitates 900 °C and 1000 °C. It can be rationalized as follows: (i) The flow stress
by dislocations at low temperatures, few stacking faults can be observed at 1100 °C is lower compared with other temperatures. (ii) The channel
at 900 °C, and no stacking faults present at 1000 °C and 1100 °C. width is much larger due to the greater dissolution of the γ′ precipitates
Dislocation loops can be frequently observed in Fig. 5a, c and d, at 1100 °C, thus the Orowan bypassing resistance decreased. (iii) The γ/
indicating that dislocation bypassing the γ′ precipitates becomes the γ′ interfacial dislocation networks at 1100 °C is more regular and
main deformation mechanism. denser, which also have some resistance to the slipping of dislocations.
Some dislocation networks are observed at 900 °C in Fig. 5b.
Morphologically, most of them are tetragonal dislocation networks, 4. Discussion
and the line directions of the dislocation nearly lie in the [110] or [110],
which deviate from the well-defined misfit orientations ([100] and 4.1. Stacking fault
[010]) to some degrees. The dislocation networks are usually formed
during early tensile creep deformation under high temperature and low Fig. 6 shows tensile properties and corresponding deformation
stress. Tensile test is a brief deformation process, which is similar to the microstructures of M951G alloy at various temperatures. In the present
deformation behavior in the initial creep stage. Although the tensile test work, stacking faults can be observed below 900 °C, but their appear-
time is very short, the flow stress is much higher compared with the ances and quantities vary remarkably with temperatures. The stacking
creep stress at this temperature. So the dislocation networks still can faults can be divided into two kinds according to their appearances, i.e.
form in the tensile test. However, the tensile test time is too short for continuous stacking fault as the low temperature type (as shown in
these dislocations to change their line directions from the 〈110〉 to the Fig. 6b and d), isolated stacking fault as the high temperature type (as

325
L. Cui et al. Materials Science & Engineering A 696 (2017) 323–330

Fig. 2. The tensile results of M951G alloy. (a) Typical tensile true stress-strain curves at various temperatures. Strain-hardening exponent (b), tensile strength (c) and tensile plasticity (d)
with temperature for M951G alloy.

shown in Fig. 6e and f). In addition, complex reactions relate to the formation mechanisms of the continuous stacking fault in Ni-base
stacking faults also have been observed in Fig. 6g at 800 °C. superalloys previously. The model provided by Décamps et al. [17] is
widely accepted. Décamps believed that these shearing processes are
4.1.1. Continuous stacking fault initiated by a dissociated a/2〈110〉 matrix dislocation, and the propa-
The continuous stacking faults can be observed at low temperatures, gation of the leading partial creates this shearing process. A decrease in
as shown in Figs. 6b and 6d. Many researchers [15,17] have studied the the SFE of γ matrix or the channel width and an increase of the applied

Fig. 3. TEM images showing deformation microstructures of M951G at low temperatures: (a)(b) RT, (c)(d) 600 °C. Beam direction close to the [001] zone axial.

326
L. Cui et al. Materials Science & Engineering A 696 (2017) 323–330

Fig. 4. TEM images showing deformation microstructures of M951G alloy at intermediate temperatures: (a) 700 °C, (b)(c) 800 °C. Beam direction close to the [001] zone axial.

stress would favor this mechanism [17]. The channel width of M951G is the continuous stacking fault. Consequently, no continuous stacking
about 80 nm, moreover, plenty of spherical γ′ particles precipitate in faults occur above 600 °C.
the matrix channel, as shown in Fig. 1b, illustrating the width between
the γ′ precipitates is relatively narrow. In addition, a number of
4.1.2. Isolated stacking fault
transition elements are added to the M951G alloy, especially Co and
As shown in Fig. 6, the type of the stacking fault changes from
W, which can significantly reduce the SFE of γ matrix [12,13]. There-
continuous stacking fault to isolated stacking fault when the tempera-
fore, the SFE of the γ matrix should be very low. What's more, the flow
ture exceeds 600 °C. And the density of stacking faults increases to the
stress in the tensile test is much higher than the creep stress at the same
maximum at 800 °C, which is in contrast with the increase of the SFE in
temperature. Thus, it can be speculated that the continuous stacking
γ′ phases with an increase in temperature [19]. It can be rationalized
faults in Fig. 6 form in this model. It has been reported that the SFE of
from the following viewpoints. The deformation mechanism is shown in
many disordered face centered cubic materials increases by a factor of 2
Eq. (1). From the viewpoint of energy, a/2〈110〉 cannot dissociate, as
when temperature increases from 0 °C to 350 °C [18]. Therefore, it is
the Ea/2〈110〉 < Ea/6〈112〉+Ea/3〈112〉, regardless of SFE of γ′ phase. It
confirm that the SFE of γ matrix increases significantly above 600 °C.
needs some necessary energy to keep the reaction balance. However,
The thermodynamic conditions cannot be satisfied for the formation of
the misfit energy between the γ and γ′ phases can just supply the energy

Fig. 5. TEM images showing deformation microstructures of M951G alloy at high temperatures: (a)(b) 900 °C, (c) 1000 °C, (d) 1100 °C. Beam direction close to the [001] zone axial.

327
L. Cui et al. Materials Science & Engineering A 696 (2017) 323–330

Fig. 6. The tensile properties and corresponding deformation microstructures of M951G alloy at various temperatures. (a)(b) RT, (c)(d) 600 °C, (e) 700 °C, (f)(g) 800 °C and (h) 1100 °C.

needed for the reaction. Generally, the thermal expansion coefficient of dislocations shear into the γ′ precipitate and react at the intersection of
ordered solid solution is smaller than disordered one [20], so the misfit (111) and (111) glide plane in the following way (as shown in Fig. 7c):
becomes more negative with the increase of temperature. This leads to
a/3[112] + a/3[211] → a /3[101] (2)
the decomposition of a/2〈101〉 matrix dislocation at γ/γ′ interface from
700 °C, and more stacking faults present in γ′ precipitates at 800 °C. The geometric and thermodynamic conditions can be satisfied for
However, when the temperature is above 900 °C, the SFE of the γ′ phase Eq. (2). By calculation, the line direction and the glide plane of the a/
is too high for the a/2〈110〉 perfect dislocation to dissociate. Conse- 3[101] dislocation are [101] and (010), respectively. Thus the a/3[101]
quently, the stacking fault in the γ′ precipitate does not occur. dislocation is a pure edge dislocation. It should be noted that the a/
3[101] dislocation is something like Lomer-cottrell dislocation lock,
4.1.3. The reactions relate to the stacking faults which is frequently observed in the fcc metal. Therefore, the newly
As shown in Fig. 6g, complex reactions have occurred between formed dislocation (called quasi-Lomer-cottrell dislocation lock) can
stacking faults “I” and “II”, resulting in the formation of a sessile neither slip on the (111) plane nor (010) plane. The reason why the
dislocation. A possible formation process is schematic illustrated in Burgers vector of the newly formed dislocation differs from sessile
Fig. 7. Two matrix dislocations with different Burgers vector dissociate dislocation of Lomer-cottrell dislocation lock may be due to the crystal
at the γ/γ′ interfaces (as shown in Fig. 7b) and the dissociation structure difference between LI2 superstructure of γ′ phase and ordinary
mechanism is the same as shown in Eq. (1). The a/3〈112〉 leading fcc metal, which results in the different decomposition of a/2〈101〉

Fig. 7. Schematic illustration of the formation process of the sessile dislocation in the γ′ phase in Fig. 4c.

328
L. Cui et al. Materials Science & Engineering A 696 (2017) 323–330

700 °C, as shown in Fig. 6. With the help of thermal activation the a/
2〈101〉 matrix dislocations in the γ′ precipitates may cross-slip from the
{111} plane to the {100} plane and form Kear-Wilsdorf lock, which
would increase the strength of alloy [22]. Furthermore, the high density
of dislocations tangle in γ matrix channel and fierce interaction
between dislocations from different slip systems at 700 °C, as shown
in Fig. 6e, also have a significant influence on the work hardening.
As shown in Fig. 6, the M951G alloy shows an obvious low ductility
at 700 °C. The behavior of intermediate temperature brittleness (ITB)
has been studied by many researchers previously. There are plenty of
reasons which can lead to ductility minima of alloys, such as strain
localization [23], a change of the deformation mechanisms [24] and
instability of γ′ phase during exposure at high temperatures [25]. As
illustrated in Fig. 6, the primary deformation mechanism blow 600 °C is
the shearing of γ′ phase by a/2〈101〉 matrix dislocations; the disloca-
tions bypassing the γ′ precipitates becomes the main mode above
Fig. 8. The element segregation between γ and γ′ phases in the M951G alloy determined
900 °C, and both deformation mechanisms are simultaneously observed
by EPMA. K=Cγ′/Cγ and K−1=Cγ/Cγ′ defined as the ratio of the elements in the γ′ phase at intermediate temperatures. At 700 °C, numbers of slip bands
to the γ phase and the γ phase to γ′ phase. consisting of high density of tangled dislocations are present in γ
matrix channel (as shown in Fig. 6e), which lead to inhomogeneous
dislocation. The special dislocation configuration has an important deformation and significant localized strain. Consequently the ductility
impact on the work hardening. minimum at 700 °C is due to strain localization and a change of the
deformation mechanisms.
As shown in Fig. 6, the yield strength reaches the maximum at
4.2. The yield strength of M951 and M951G alloys
800 °C. With the increase of temperature from 700 °C to 800 °C, the
strength of γ′ precipitates increases [21] as well as the SFE of γ′ phase
As shown in Fig. 6, the yield strength of M951G alloy is much higher
enhances [9,26]. Then it is difficult for a/3〈112〉 super-partial disloca-
than M951 below 800 °C. It can be interpreted from the following
tions to cut into the γ′ precipitates at 800 °C. In order to maintain the
viewpoints: (i) At lower temperatures (below 600 °C), compared with
same applied strain rate at 700 °C and 800 °C, the yield strength at
M951 alloy [14], numbers of continuous stacking faults present in the
800 °C is higher than that at 700 °C. What's more, the plenty of stacking
microstructure of M951G alloy. The significant reduction of SFE by
faults in γ′ precipitates and the formation of the quasi-Lomer-cottrell
adding W and Hf is favorable for the formation of continuous stacking
dislocation lock at 800 °C also create barriers for cross-slip. Therefore,
faults, which could block the mobile dislocations and increase the
the yield strength at 800 °C higher compared with that at 700 °C.
resistance of cross-slip. At higher temperatures (600–800 °C), the high
The tensile strength of M951G alloy rapidly decreases when
density of tangled dislocations in γ matrix channel (as shown in Fig. 6e)
temperature above 800 °C, as shown in Fig. 6. Due to the coarsening
and lots of stacking faults in γ′ precipitates (as shown in Fig. 6f)
of γ′ phase present at local areas and partial dissolution of the γ′
significantly impede the cross-slip of dislocations in the {111} planes.
precipitates, the strength drops obviously at higher temperatures. On
Furthermore, the formation of quasi-Lomer-cottrell dislocation lock also
the other hand, the dominate deformation mechanism has changed
has some impact on work hardening at intermediate temperatures. (ii)
from shearing to bypassing above 800 °C. The local Orowan resistance
The higher γ′ content (about 54%) of M951G alloy compared with
of the dislocations slipping in channels is as follows [3]:
M951 alloy, as shown in Table 1, also has an impact on tensile strength.
(iii) As shown in Fig. 8, in the present study, transition elements W and 2 Gb
τ=
Hf tend to segregate to the γ matrix and γ′ precipitates respectively. 3 h (3)
Thus, the higher W and Hf additions (as shown in Table 1) in the
where G is the shear modulus, b the Burgers vector, and h is the channel
M951G alloy are expected to result in a stronger effect of solid solution
width. Due to the higher degree dissolution of the γ′ phase, the channel
strengthening. Therefore, the yield strength of M951G alloy is much
width at 1100 °C is much larger than at other temperatures, therefore
higher than M951 below 800 °C. However, when the temperature above
the yield strength is very low at 1100 °C.
900 °C, partial γ′ phase has dissolved and the dominant deformation
mechanism of both M951 and M951G alloys changes from shearing to
5. Conclusions
bypassing. Hence the strength intervals between M951 alloy and
M951G alloy to become increasingly narrow above 800 °C.
The microstructures of M951G alloy were characterized by scanning
electron microscopy and transmission electron microscopy after tensile
4.3. Relationship between tensile properties and deformation tests at different temperatures, and the corresponding deformation
microstructures mechanisms have been analyzed in detail. The following conclusions
can be drawn from this work:
As shown in Fig. 6, the tensile properties and deformation micro-
structures vary significantly with the change of temperatures. At low 1) The stacking faults in present study can be divided into two types
temperatures, the majority of plastic deformation takes place in γ according to their appearances, i.e. continuous stacking fault as the
phase. It has been reported that the strength of the γ phase decreases low temperature type (20–600 °C), isolated stacking fault as the high
from room temperature to 600 °C [21], which may be responsible for temperature type (700–900 °C).
the reduction of the yield strength from room temperature to 600 °C. 2) The higher additions of W and Hf enhance the tensile strength of
The deformation at room temperature is very inhomogeneous and large M951G alloy below 800 °C, which has been rationalized from the
numbers of dislocations tangle in the γ matrix channel as shown in points of deformation microstructures, the strong effect of solid
Fig. 6a, which lead to significant strain localization. Thus, the tensile solution strengthening and precipitation strengthening.
plasticity at room temperature is very low. 3) The reactions relate to the stacking faults from different slip systems
It can be seen that the ultimate strength reaches the maximum at

329
L. Cui et al. Materials Science & Engineering A 696 (2017) 323–330

occur at 800 °C, and a sessile dislocation is formed, which has some (2007) 5802–5812.
[11] D.N. Duhl, C.P. Sullivan, Some effects of hafnium additions on mechanical
influence on work hardening. properties of a columnar-grained nickel-base superalloy, J. Met. 23 (1971) 38–40.
4) The primary deformation mechanisms include shearing of the γ′ [12] N.S. Yang, Y.H. Wei, H.Z. Fu, W.G. Li, R.D. Tang, Influence of W and Co on high
precipitates by a/2〈110〉 dislocations below 600 °C and by-passing temperature creep behaviour of Ni-20Cr base alloys, Acta Metall. Sin. 17 (1981)
44–49.
of γ′ precipitates by a/2〈110〉 matrix dislocations above 900 °C. At [13] G.P. Zhao, B.K. Zhao, Behavior of W and Hf in GH586 alloy, Acta Metall. Sin. 31
intermediate temperatures, it shows a transition from shearing to (1995) S193–S197.
by-passing. [14] Z.W. Lian, J.J. Yu, et al., Temperature dependence of tensile behavior of Ni-based
superalloy M951, Mater. Sci. Eng. A 489 (2008) 227–233.
[15] G.B. Viswanathan, P.M. Sarosi, M.F. Henry, D.D. Whitis, W.W. Milligan, M.J. Mills,
Acknowledgments Investigation of creep deformation mechanisms at intermediate temperatures in
René 88 DT, Acta Mater. 53 (2005) 3041–3057.
[16] P. Caron, T. Khan, P. Veyssière, On precipitate shearing by superlattice stacking
This work was financially supported by the National Natural Science
faults in superalloys, Philos. Mag. A 57 (1988) 859–875.
Foundation of China (NSFC) under Grant nos. 50931004 and [17] B. Décamps, S. Raujol, et al., On the shearing mechanism of gamma' precipitates by
51571196. The authors are grateful for those supports. a single a/2〈110〉 dissociated matrix dislocations in Ni–based superalloys, Philos.
Mag. A 84 (2004) 91–107.
[18] A. Sengupta, S.K. Putatunda, L. Bartosiewicz, J. Hangas, P.J. Nailos, M. Peputapeck,
References F.E. Alberts, Tensile behavior of a new single crystal nickel-based superalloy
(CMSX-4) at room and elevated temperatures, J. Mater. Eng. Perform. 3 (1994)
[1] Q.H. Wu, J. Zhang, Y.S. Luo, Composition and mechanical property of DD6 664–672.
superalloy revert, Mater. Sci. Forum 788 (2014) 488–492. [19] X.G. Wang, J.L. Liu, T. Jin, X.F. Sun, Tensile behaviors and deformation mechan-
[2] M. Kolbe, K. Neuking, G. Eggeler, Dislocation reactions and microstructural isms of a nickel-base single crystal superalloy at different temperatures, Mater. Sci.
instability during 1025 °C shear creep testing of superalloy single crystals, Mater. Eng. A 598 (2014) 154–161.
Sci. Eng. A 234 (1997) 877–879. [20] L. Müller, T. Link, M. Feller-Kniepmeier, Temperature dependence of the thermal
[3] T.M. Pollock, A.S. Argon, Creep resistance of CMSX-3 nickel base superalloy single lattice mismatch in a single crystal nickel-base superalloy measured by neutron
crystals, Acta Metall. Mater. 40 (1992) 1–30. diffraction, Scr. Metall. Mater. 26 (1992) 1297–1302.
[4] L. Rémy, A. Pineau, B. Thomas, Temperature dependence of stacking fault energy in [21] P. Beadmore, R.G. Davies, T.L. Johnston, On the temperature dependence of the
close-packed metals and alloys, Mater. Sci. Eng. 36 (1978) 47–63. flow stress of nickel-base alloys, Trans. AIME 245 (1969) 1537–1545.
[5] H. Rouault-Rogez, M. Dupeux, M. Ignat, High temperature tensile creep of CMSX-2 [22] W. Österle, D. Bettge, B. Fedelich, H. Klingelhöffer, Modelling the orientation and
Nickel base superalloy single crystals, Acta Metall. Mater. 42 (1994) 3137–3148. direction dependence of the critical resolved shear stress of nickel-base superalloy
[6] A. Sengupta, S.K. Putatunda, et al., Tensile behavior of a new single-crystal nickel- single crystals, Acta Mater. 48 (2000) 689–700.
based superalloy (CMSX-4) at room and elevated temperatures, J. Mater. Eng. [23] Z. Chu, J. Yu, X. Sun, H. Guan, Z. Hu, Tensile property and deformation behavior of
Perform. 3 (1994) 73–81. a directionally solidified Ni-base superalloy, Mater. Sci. Eng. A 527 (2010)
[7] C.G. Tian, G.M. Han, C.Y. Cui, X.F. Sun, Effects of Co content on tensile properties 3010–3014.
and deformation behaviors of Ni-based disk superalloys at different temperatures, [24] D. Bettge, W. Oeesterle, J. Ziebs, Temperature dependence of yield strength and
Mater. Des. 88 (2015) 123–131. elongation of the nickel-base superalloy IN 738 LC and the corresponding
[8] D. Qi, B. Fu, K. Du, T. Yao, C. Cui, J. Zhang, H. Ye, Temperature effects on the microstructural evolution, Z. Met. 86 (1995) 190–197.
transition from Lomer-Cottrell locks to deformation twinning in a Ni-Co-based [25] S.M. Copley, B.H. Kear, G.M. Rowe, The temperature and orientation dependence of
superalloy, Scr. Mater. 125 (2016) 24–28. yielding in Mar-M200 single crystals, Mater. Sci. Eng. 10 (1972) 87–92.
[9] F. Diologent, P. Caron, On the creep behavior at 1033 K of new generation single- [26] J. Courbon, F. Louchet, On the mechanisms of formation of superlattice stacking
crystal superalloys, Mater. Sci. Eng. A 385 (2004) 245–257. faults in L12 precipitates embedded in a F.C.C. matrix, Phys. Status Solidi A 137
[10] S. Ma, L. Carroll, T.M. Pollock, Development of γ phase stacking faults during high (1993) 417–428.
temperature creep of Ru-containing single crystal superalloys, Acta Mater. 55

330

You might also like