You are on page 1of 16

International Journal of Plasticity 25 (2009) 822837

Contents lists available at ScienceDirect

International Journal of Plasticity


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

High temperature strength and failure of the Ni-base


superalloy PM 3030
Michel Nganbe a,*, Martin Heilmaier b
a
b

University of Ottawa, Department of Mechanical Engineering, 161 Louis-Pasteur, Ottawa, ON, Canada K1N 6N5
Otto-von-Guericke Universitt Magdeburg, Institut fr Werkstoff- und Fgetechnik, D-39016 Magdeburg, Germany

a r t i c l e

i n f o

Article history:
Received 10 April 2008
Received in nal revised form 12 June 2008
Available online 21 June 2008
Keywords:
Strengthening mechanisms
Fatigue
Fracture
Microstructure
Optimization

a b s t r a c t
The strength, fatigue life and fracture behavior of the oxide dispersion strengthened (ODS) nickel-base superalloy PM 3030 are investigated. The high Al content in PM 3030 leads to the formation of
coherent c0 particles and, thus, to additional precipitation strengthening. A coarse and elongated grain structure (R34) and two isotropic batches with mean grain sizes of 1 lm (R90) and 17 lm
(R901315) are considered. Compressive constant strain rate tests
and high cycle fatigue (HCF) tests are performed. Optical, scanning
and transmission electron microscopy (OM, SEM and TEM) are carried out. The properties are compared with those of the solely
oxide dispersion strengthened Ni-base alloy PM 1000 [Estrin, Y.,
Heilmaier, M., Drew, G., 1999. Creep properties of an oxide dispersion strengthened nickel-base alloy: the effect of grain orientation
and grain aspect ratio. Mater. Sci. Eng. A 272(1), 163173]. It is
found that additional c0 hardening provides an increase in quasistatic strength by about a factor 2 and in HCF life by about a factor
102103 at temperatures up to 850 C. When fatigue life is compared at a xed ratio of stress amplitude-to-yield or ultimate compressive strength, R34 shows a fatigue life similar to that of PM
1000 at lower temperature (e.g. 600 C) indicating that the quasistatic strength advantage is proportionally translated into
improved fatigue performance; for higher temperatures (850 C)
however, R34 shows a shorter fatigue life as compared to PM
1000. Grain size reduction, as exemplied with the ne grain R90
batch, also provides an increase in strength up to the equicohesion
temperature (TE) [Dieter, G.E., 1986. Mechanical Metallurgy. SI
Metric Ed. McGraw-Hill Book Company, London]. Above TE, faster
diffusion and grain boundary sliding [Raj, R., Ashby, M.F., 1971.
On the grain boundary sliding and diffusional creep. Metall. Trans.

* Corresponding author. Tel.: +1 613 562 5800; fax: +1 613 652 5177.
E-mail address: mnganbe@uottawa.ca (M. Nganbe).
0749-6419/$ - see front matter 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijplas.2008.06.005

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

823

2, 11131127; Spingarn, J.R., Nix, W.D., 1978. Diffusional creep and


diffusionally accommodated grain rearrangement. Acta Metall. 26,
13891398] lead to a drastic drop in strength for the R90 material.
In contrast, the batch with intermediate grain size (R901315)
shows strength comparable to that of R34 up to 850 C. Furthermore, R901315 shows improved crack tolerance compared to its
coarse grain counterpart R34. Due to premature crack initiating
coarse oxide particles however, R901315 does not show any
improvement in elongation to failure during tensile tests. Eliminating those coarse particles is expected to improve the ductility and
toughness of this isotropic batch.
2008 Elsevier Ltd. All rights reserved.

1. Introduction
Components in high temperature applications such as gas turbines commonly experience hot gas
corrosion and erosion as well as static and dynamic loading. Examples of loading contributions include
bending and torsion due to owing high pressure gas as well as tension due to centrifugal forces. Oxide
dispersion strengthened (ODS) nickel-base superalloys have been considered promising candidate
materials to fulll these multiple challenges. Due to their excellent thermal stability, oxide dispersoids
have to be generally incorporated into a metal matrix by mechanical alloying in order to provide creep
resistance even at extremely high temperatures close to the melting point of the matrix alloy (Benjamin, 1970; Cairns et al., 1975). Single crystals or coarse and elongated grain structures are commonly
used in order to limit diffusion and grain boundary sliding (Raj and Ashby, 1971; Spingarn and Nix,
1978) as well as void formation at grain boundaries typical for extremely high temperature applications (deMestral et al., 1996; Stevens and Nix, 1985). However, ODS materials could not make a breakthrough into large scale manufacture and use until now due to high costs associated with processing
technology, most notably mechanical alloying and powder consolidation as well as costly heat treatments required to achieve conventional coarse and elongated grain structures. The motivation for
the research work presented in this paper is the observation that many applications in engines, aircraft
and stationary gas turbines would not be characterized by extremely high temperatures. Rather, common stationary operating conditions are temperatures in the intermediate range around 800 C along
with relatively high static and dynamic mechanical loads. A prominent example would be gas turbine
disks which are presently made of c0 -strengthened powder metallurgical (PM) Ni-base superalloys. Under such conditions dislocation slip is assumed to be the primary deformation process whereas diffusional creep processes (Stickforth, 1986) are considered to be secondary phenomena (Frost and Ashby,
1982; Arzt, 1991). Therefore, the superimposed contribution from c0 precipitation hardening and oxide
dispersion strengthening in PM Ni-base superalloys might provide improved strength and fatigue properties. Such a combination, however, may neither be advantageous at extremely high temperatures due
to the thermal instability of c0 particles (Biermann, 1999; Nganbe, 2002) nor at low temperatures
where solely c0 strengthened alloys provide sufcient performance. Furthermore, the cost intensive
manufacturing of single crystalline alloys or elongated and strongly textured grain structures by advanced heat treatments with appropriate thermal gradients may be superuous for applications at
intermediate temperatures. Isotropic grain structures may provide equivalent or superior strength
and fatigue life; therefore, they are again gaining increasing interest from both research community
and industry. The proposed in-depth investigation of the inuence of additional c0 strengthening and
grain size variation on high temperature strength and fatigue performance may, thus, yield criteria
for an optimal microstructural design of c0 precipitation strengthened ODS PM Ni-base superalloys.
2. Materials and experimental
PM 3030 is a Ni-base superalloy manufactured by Plansee GmbH, Lechbruck, Germany, with a
nominal chemical composition in weight percentage of 17Cr-2Mo-3.5W-2Ta-6.6Al-1.1Y2O3-bal. nick-

824

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

el. Appropriate heat treatment and the relatively high aluminum content result in the precipitation of
L12 ordered Ni3Al intermetallic c0 particles within the disordered face centered cubic (fcc) Ni matrix.
Very good corrosion resistance is provided by relatively high chromium and aluminum contents at
intermediate and high temperatures, respectively. Incorporation of thermally stable Y2O3 dispersoids
is realized by mechanical alloying (MA) (Benjamin, 1970). A secondary recrystallized, fully heat treated (1 h/1230 C/air cooled + 2 h/955 C/air cooled + 24 h/845 C/air cooled) batch of PM 3030 was obtained from Plansee GmbH as a round bar with 34 mm diameter (designated R34 as follows) and
compared with PM 1000, a solely solid solution and oxide dispersion hardened alloy with similar grain
structure and chemical composition, except for the much lower Al content which therefore does not
lead to c0 precipitates (Estrin et al., 1999). The comparison allows evaluating the contribution of c0 precipitation strengthening to the overall strength of PM 3030. Furthermore, data of a purely solid solution strengthened alloy, Nimonic 75, were obtained from literature (Heilmaier, 1992; Heilmaier and
Reppich, 1996). Comparing the strength of Nimonic 75 with that of PM 1000 allows the assessment
of the contribution of dispersion hardening to strength. A non-recrystallized variant with ne and
(crystallographically) isotropic grains (R90, round bar with diameter 90 mm after hot extrusion)
was also received from Plansee. A part of the R90 bar has been used to develop a coarser grained additional batch of PM 3030 in the lab using an isothermal heat treatment at 1315 C for 3 h (designated
R901315). Since this high temperature heat treatment had to be carried out above the c0 solvus, reprecipitation by isothermal heat treatment at 850 C has been done producing c0 precipitate parameters similar to those of R34.
Grain structure and damage mechanisms are investigated by optical microscopy (OM). Transmission electron microscopy (TEM, JEOL JSM 2000FX) is used for the characterization of oxide particles
as well as for the investigation of dislocation structures and deformation mechanisms. Scanning electron microscopy (SEM, JEOL JSM6400) is used for c0 particle characterization and fractography investigation of fatigue samples.
Mechanical tests have been performed on a computer controlled electromechanical testing device
INSTRON 8562. In contrast to determining ultimate tensile strength (UTS) and elongation to failure
that requires tensile testing of relatively large samples, the 0.2% yield strength r0.2 can well be determined in compression on small, cost efcient samples with good accuracy. Therefore, compression
tests have been performed in air at a constant (true) strain rate of 104 s1 between room temperature
and 1250 C using cylindrical samples with 6 mm height and 4 mm diameter. Furthermore, fatigue
tests have been performed on samples prepared according to ASTM standard E606. The total sample
length was 100 mm and the sample diameter at the xtures was 10 mm; the actual test section at the
centre of the samples had initial gage length and diameter of 20 mm and 6 mm, respectively. A triangular wave shape and a frequency of 1 Hz were used. Two temperatures, representative for the intermediate temperature range, have been chosen for fatigue testing: 600 C slightly below the
equicohesion temperature (TE) and 850 C above TE. Tensile tests have then been performed at the
technologically relevant temperature of 850 C using ASTM E606 samples as well in order to assess
the ductility of the R34 and R901315 batches of PM 3030 with the highest application potential.

3. Results and discussion


3.1. Microstructure and microstructure changes
Similar to PM 1000 (Estrin et al., 1999), R34 has large and elongated grains (in the millimeter
range) with a grain aspect ratio (GAR) value above 10. In contrast, R90 has isotropic grains with a
mean grain size of 1 lm. The heat treatment performed in the laboratory produced R901315 with
an isotropic mean grain size of 17 lm. In contrast to R34 and R90, R901315 shows a strong concentration of precipitates along grain boundaries, predominantly c0 precipitates and aluminumyttrium-oxide particles. The c0 particle structure of R34 is shown in Fig. 1. The precipitates are
homogeneously distributed in the solid solution c matrix. They are cuboidal in shape and exhibit
an edge length of around 555 nm (SEM value), which is much larger compared to alloys of other manufacturers such as MA 6000 (Heilmaier and Reppich, 1996) and can potentially result in lower strength

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

825

Fig. 1. SEM micrograph of the c0 precipitation structure in PM 3030 R34; the c0 particle size and volume fraction are about
555 nm and 55%, respectively. The c0 particle structure in R901315 is similar.

as discussed in the following Section 3.2. The volume fraction of c0 precipitates has been quantied as
55% (SEM value) using the procedure outlined in (Heilmaier et al., 2001), which approaches typical
values of single crystal alloys ranging from 60 to 75% (Nabarro and Villiers, 1995). The edges of the
c0 cubes are aligned along h1 0 0i crystallographic directions. The c0 precipitate structure in R901315
grains is essentially identical to that of R34. Because of the small grain size (around 1 lm) the microstructure of R90 shows features of a duplex structure of c and c0 grains lying adjacent to each other
rather than those of a precipitation hardened alloy (Fig. 2), but an analysis of their crystallographic
relation was not undertaken here. The dispersoids are found to be homogeneously distributed in
the c matrix as well as in the c0 precipitates. In R34 they are spherical in shape and have a mean diameter and a volume fraction of about 30 nm and 2.1%, respectively (Fig. 3). Those values are outside the
range for optimum creep rupture properties according to literature (Benjamin and Bomford, 1974) as
can be seen in Fig. 4, which again indicates that there is potential for further optimization of the dispersoid structure in R34. By contrast, the dispersoid parameters in R90 are within the optimal range;
the oxide particle size is about 18 nm and the volume fraction is about 1.7%. The additional heat treatment used to create R901315 led to signicant coarsening of dispersoids to a mean particle size of
about 42 nm. The oxide volume fraction in R901315 (about 1.6%) remains about equal to that of
R90 meaning that large particles predominantly coarsen at the expense of smaller particles without
any noticeable reaction with other elements from the matrix; this oxide coarsening process obviously
takes place during the grain growth heat treatment at the temperature of 1315 C. Very coarse oxide

Fig. 2. TEM micrograph of the c/c0 duplex grain structure in PM 3030 R90; c0 particles build own grains: c grains appear
brighter and c0 grains appear darker.

826

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

Fig. 3. TEM micrograph of the yttrium-oxide dispersoid structure in PM 3030 R34; oxide structures in R90 and R901315 are
similar. The dispersoid size and volume fraction are about 30 nm and 2.1% (R34), 18 nm and 1.7% (R90), 42 nm and 1.6%
(R901315), respectively.

R34 (30nm,2.1%)

R90 (18nm,1.7%)

R901315 (42nm,1.6%)

fOx / %

89,6 MPa
68,9 MPa

103,4 MPa

3
110,3 MPa

110,3 MPa

2
103,4 MPa

113,8 MPa

89,6 MPa
68,9 MP a

20

30

40

50

60

d Ox / nm
Fig. 4. Impact of Y2O3 particle size (dox) and volume fraction (fox) on the creep life of oxide dispersion and c0 strengthened Ni
Cr-superalloys at 1038 C (Benjamin and Bomford, 1974). Data points highlighting the dispersoid parameters of PM 3030
batches as well as the optimal ranges (hatched) according to Benjamin and Bomford (1974) are shown. The solid curves show
creep stresses for 100 h rupture time.

particles are also observed at the grain boundaries of R901315; it is believed here and supported by
observations from (Heilmaier, 1992; Hotzler and Glasgow, 1982) that they are created by the reaction
of the original Y2O3 particles with aluminum from the matrix to form mixed yttriumaluminum-oxides. In summary, further optimizations of the microstructure of PM 3030 batches are expected to provide properties better than those presented in this paper: the c0 particle structure shall be
substantially rened; oxide dispersoid size and volume fraction in R34 and R901315 shall be reduced
(Fig. 4).
Two types of investigations are carried out on small sample discs of R34 to assess the thermal stability of both c0 precipitates and oxide dispersoids. First, particle dissolution is evaluated using differential scanning calorimetry (DSC) at a constant heating rate of 20 K/min until melting of the samples;
the result is shown in Fig. 5. It is observed that c0 dissolution starts at about 800 C or less and ends
above 1160 C. Neither melting nor any other structural change of the dispersoids can be observed during DSC tests. Second, isothermal changes of particles are investigated by heating the samples at the
same rate of 20 K/min to 1000 C and then holding. The discs are then quenched in air after different
isothermal holding durations and investigated in SEM for c0 precipitate and in TEM for oxide dispersoid
parameters. Practically no change in oxide particle size and volume fraction can be observed even after

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

827

R34; 20 K/min

Q/ w.E.

Endothermic

Melting

200

Dissolution of
' precipitates

400

600

800

1000

1200

1400

T / C
Fig. 5. Differential scanning calorimetry (DSC) curve showing endothermic peaks for dissolution of c0 precipitates and melting
of the c matrix; heating rate 20 K/min.

durations above 1000 h; the measured variations remain within the measurement error (Fig. 6). In
contrast, the average c0 precipitate size increased from 555 nm to 677 nm (SEM values), and the c0 volume fraction decreased from 55% to its thermodynamic equilibrium value of only 32% (SEM values, see
Fig. 7). The combination of larger size and lower volume fraction of c0 precipitates ultimately leads to a
substantial increase in their particle spacing and, therefore, a pronounced decay of the c0 strengthening
contribution can be expected. Two facts shall be mentioned here: First, the selected temperature
1000 C is within the c0 dissolution range as can be seen in the DSC trace in Fig. 5. Second, c0 particle
size and volume fraction have been calculated based on surface analysis; SEM values are therefore
used for all estimate and discussion regarding c0 precipitates in this paper because they are more accurate since SEM pictures show actual surface features of the samples. In contrast, TEM shows the projection of volume features onto the micrograph picture; therefore the simplied calculation of surface
and volume fractions based on TEM pictures as shown in Fig. 7 are overestimated.
3.2. Temperature dependence of the quasi-static strength of PM 3030
The temperature dependence of the compressive yield strength r0.2 of all three PM 3030 batches as
well as that of PM 1000 is illustrated in Fig. 8. At temperatures up to 600 C, about half the melting

40

1000C

1000C

fOx /%

dOx / nm

2.5
35
2

30
1.5

25
1

1
100

t/h

10000

100

t/h

Fig. 6. Time dependent change of oxide dispersoid size (a) and volume fraction (b) at 1000 C.

10000

828

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

1200

1000C

50

800

f / %

600

40

d / nm

1000C

60

1000

400
200

70

30
20

SEM

0
1

10000

100

SEM

10

TEM

TEM

100

10000

t/h

t/h

Fig. 7. Time dependent coarsening (a) and dissolution (b) of c0 precipitates at 1000 C. SEM data are used for surface analysis
and evaluation.

Yield strength / MPa

equicohesion range

R90

1200

.
= 10 -4s -1

R901315
R34

800

MA 6000

PM 1000
400

400

800

1200

T / C
Fig. 8. Temperature dependence of the yield strength of investigated alloys of PM 3030 and PM 1000. The equicohesion range is
shown and data of MA 6000 are also plotted.

point of the metallic matrix (Tm = 1350 C), plastic deformation occurs predominantly by dislocation
glide on well dened {1 1 1}h1 1 0i slip systems of the fcc Ni-base matrix. Dislocation motion can therefore be efciently impaired at matrixdispersoid interfaces according to the Orowan process (Kocks,
1977; Khraishi et al., 2004; Morris et al., 2008) on the one hand, and at c/c0 interfaces due to the necessity of creating antiphase boundaries and maintaining them throughout the c0 precipitate cutting process (Copley and Kear, 1967a; Yashiro et al., 2006). As a result, the combined oxide dispersion and c0
particle strengthening gives the coarse grain variant of PM 3030 (R34) a superior quasi-static strength
being about twice as high as that of the solely oxide dispersion strengthened alloy PM 1000. This
advantage is maintained almost constant from room temperature up to about 600 C, but decreases
progressively towards higher temperatures. This behavior can be explained by the following experimental observations: rst, the increasing inuence of diffusional creep processes above 600 C leads
to predominant dislocation climb processes in the softer c channels. Consequently, a high density
of dislocations in c channels can be observed in samples deformed at high temperatures (Fig. 9:
850 C). In contrast, the relatively homogeneous distribution of dislocations in both c matrix and c0
precipitates at low temperatures (Fig. 10: 600 C) is an evidence for glide-controlled dislocation cutting of precipitates. Second, c0 particles start dissolving at about 800 C as described in Section 3.1,
which leads to a time and temperature dependent reduction of their volume fraction (Fig. 7), to an increase in c0 interparticle spacing and, thus, to an additional reduction in their strengthening effect. The

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

829

Fig. 9. TEM micrograph of PM 3030 R34 sample following compressive constant strain rate test at 850 C: Dislocations
circumventing c0 precipitates by moving along c channels.

Fig. 10. TEM micrograph of PM 3030 R34 sample following compressive constant strain rate test at 600 C: similar dislocation
density in both c and c0 phases suggests that precipitate cutting process takes place.

dissolution of c0 precipitates is nished at about 1160 C; consequently, above this temperature the
strength of the coarse and elongated PM 3030 batch (R34) becomes identical to that of PM 1000. In
other words, PM 3030 behaves like a purely solid solution and oxide dispersion strengthened superalloy above 1160 C. Another observation to highlight in Fig. 8 is the fact that the strength of PM 3030
is lower than that of MA 6000 at temperatures between 800 and 1100 C (Nganbe, 2002; Heilmaier,
1992) due to the coarser c0 precipitates as discussed in Section 3.1. This demonstrates that an optimization of the c0 precipitate structure could further improve the mechanical properties of PM 3030, particularly for the higher intermediate temperature range.
Compared to R34, the strength of the ne grain variant R90 is about twice as high at low temperatures below 450 C. Grain boundary strengthening appears to be very strong in R90 because of its duplex character. However, a signicant drop in strength is already observed at 450 C due to enhanced
diffusion caused by the high density of grain boundaries in this variant. The strength of R90 diminishes
above 1100 C with yield strength values around only 13 MPa. The mechanical strength of R901315 is

830

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

a compromise between R34 and R90. In the low temperature regime, the strength of R901315 is lower
than that of R90, but still slightly higher than that of R34 which can be rationalized by additional
strengthening by grain size reduction (Hall, 1951; Petch, 1953). The r0.2 T curves of all PM 3030 variants cross over in a narrow temperature range between 650 and 680 C, which is called the equicohesion temperature (TE) (Dieter, 1986). In this temperature range, the strength of PM 3030 is
independent of grain size, which means a balance prevails between the strengthening effect of grain
boundaries as obstacles for dislocation motion and their softening effect as locations for accelerated
diffusion and grain boundary sliding (Raj and Ashby, 1971; Spingarn and Nix, 1978). Above the equicohesion temperature, diffusion along grain boundaries overcompensates the dislocation motion
obstruction effect of grain boundaries; the strength of R901315 therefore falls below that of R34 above
850 C and diminishes above 1200 C with yield strength values below 17 MPa. This suggests that the
technological applicability of such isotropic grain materials is limited at intermediate temperatures.
However, a further optimization of grain size to values of up to 100 lm might provide additional
improvement of the high temperature strength of R901315.
3.3. Model-based understanding of oxide dispersion and c0 precipitation strengthening contributions in PM
3030
Comparing PM 1000 with Nimonic 75, Nganbe and Heilmaier (2004) demonstrated that oxide dispersion strengthening contribution to the yield strength below 0.5 Tm can well be described using the
actual Orowan formula as derived by Kocks (1977):

rPM1000 rmatrix DrOR rNimonic75 0:9M

ln8rs =b3=2
lnL=b1=2

"

#
K edge
bL  2r s

Ab
with K edge 4pG1
mA  being the pre-logarithmic coefcient of the line tension of an edge dislocation

segment in the critical by-passing conguration and rs p4 r ox being the mean planar particle radius
in the glide plane.
In Eq. (1) the matrix strength of PM 1000 is set equal to the strength of the solely solid solution
strengthened alloy Nimonic 75 and Dror is the oxide dispersion strengthening contribution according
to the Orowan process. Instead of the common value for polycrystalline materials a Taylor factor
M = 2.45 is used here because of the strong crystallographic texture and the coarse grained microstructure (Estrin et al., 1999); b is the Burgers vector; L and rox are the mean oxide interparticle spacing and radius, respectively; GA and mA are the anisotropic shear modulus and Poisson ratio for the
{1 1 1}h1 1 0i slip systems that are relevant for the fcc matrix of all investigated materials. In lack of
appropriate data, single crystal constants determined on the cast c0 precipitation strengthened Ni-base
superalloy IN738 LC (Bayerlein, 1991) with chemical composition similar to PM 3030 are used to calculate GA and mA; specic values are shown in Table 1. Above 0.5 Tm, climb-controlled dislocation
movement prevails and the dispersoid detachment model lastly rened by Reppich (1998) can be used
to accurately describe oxide dispersion strengthening:

R3=2

rPM1000 rmatrix Drocr rNimonic75 0:9M p

2 2 R3=2

GA b=L

where Drocr is the oxide dispersion strengthening contribution to the yield strength. With an R value
of 0.77 as expected by Arzt and Ashby (1982) for local p
climb,
Eq.
(2) describes test data accurately
2
above 800 C (Nganbe and Heilmaier, 2004). Using R 2 1  k (Reppich, 1998), R = 0.77 also corresponds to a relaxation factor k = 0.923 for dislocation relaxation within the particle/matrix interface
(Arzt and Wilkinson, 1986; Arzt and Rsler, 1988). Because this value of k is smaller than 0.94, which
is known to be sufciently low to cause interfacial pinning of dislocations at the departure side of dispersoids (Arzt and Wilkinson, 1986; Heilmaier and Reppich, 1999), both processes local climb and
interfacial pinning are considered to concurrently control dislocation/dispersoid interaction and consequently dispersion strengthening in PM 3030 in the present approach. The (rather weak) temperature dependence of oxide dispersion strengthening in any of both regimes is accounted for by the
variation of GA and mA with temperature (Table 1).

831

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

Table 1
Anisotropic shear modulus (GA) and Poisson ratio (mA) calculated using single crystal constants of IN738 LC (Bayerlein, 1991) and
the pre-logarithmic coefcient of the line tension of an edge dislocation segment in the critical conguration during by-passing of
dispersoids (Kedge)
T/C

GA/GPa

mA

Kedge/109N

25
220
420
615
800
850
992
1092
1200
1250

75.332
71.072
67.358
62.671
57.660
55.825
49.231
43.114
34.610
30.610

0.417
0.423
0.406
0.422
0.436
0.436
0.447
0.464
0.480
0.491

0.663
0.633
0.582
0.557
0.525
0.508
0.457
0.413
0.342
0.309

Assuming linear superposition of dispersion and c0 precipitation strengthening, Nganbe and Heilmaier (2004) modied the original Copley and Kear (1967b) equation to describe the c0 strengthening
contribution to the yield strength of PM 3030 for temperatures up to 1000 C:

rPM3030 rmatrix Drox Drc0 rPM1000 M

 c
1 0
Cm
K sc sc0 APB 
2
2b
brc0


3

where Drox and Drc0 are the oxide dispersion and the c0 precipitation strengthening contributions
respectively; K0 = 0.32 is a t factor and rc0 the mean equivalent radius of c0 precipitates. sc is set equal
to the yield shear strength of PM 1000. The yield shear strength values sc0 of Ni86Al14 (Nabarro and
Villiers, 1995) are used after addition of an oxide dispersion hardening contribution set equal to that
in the c matrix. Because of the complex dislocation conguration during cutting of c0 precipitates the
line tension Cm of mixed dislocations according to the equation below was used:

Cm
edge

Cedge Cscrew

screw

where C
and C
are the line tension of edge and screw dislocations, respectively. In lack of
appropriate data, single crystal constants determined on the cast c0 precipitation strengthened Ni-base
superalloy IN 760 (Bayerlein, 1991) are used to calculate Cedge, Cscrew and Cm; values are shown in
Table 2. The temperature dependence of sc, sc0 and Tm has been considered for modeling. Antiphase
boundary energy (cAPB) values are taken from IN 760 as well; they are essentially constant up to about
750 C (Copley and Kear, 1967b); cAPB was therefore considered constant and equal to 0.17 J/m2 for
modeling.
Eq. (3) describes test data accurately up to 850 C and even reproduces the slight stress anomaly
peak at about 650 C (Nganbe and Heilmaier, 2004). This phenomenon, where strength increases with
increasing temperature, has been frequently observed in the bulk intermetallic compound Ni3Al and in
Ni-base superalloys containing high volume fractions of c0 . It has been explained as thermally activated cross-slip of {1 1 1}h1 1 0i dislocation pairs (superdislocations) (Nabarro and Villiers, 1995; Nabarro and Villiers, 1996; Behr et al., 1997) onto {0 0 1} planes, thereby creating additional antiphase

Table 2
Line tension of edge (Cedge), screw (Cscrew) and mixed (Cm) dislocations calculated using single crystal constants of IN760 LC
(Bayerlein, 1991)
T/C

Cedge/109N

Cscrew/109N

Cm/109N

22
400
600
681
1000

0.993
0.990
0.977
0.973
5.191

5.649
5.632
5.560
5.536
5.191

2.368
2.362
2.331
2.321
5.191

832

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

boundary segments that can act as KearWilsdorf locks (Kear and Wilsdorf, 1962; Nabarro and Villiers, 1995). This results in strongly impairing the overall dislocation motion. The strength of PM
3030 is overestimated above 850 C due to the change from relatively difcult low temperature cutting of c0 precipitates that is assumption for the used model to relatively easy circumvention of c0 precipitates at higher temperatures as discussed in the previous Section 3.2.
3.4. HCF lifetime
Due to the poor quasi-static strength of R90 at intermediate and high temperatures, fatigue investigations have been limited to R34 and R901315. The HCF behavior of R34 is presented and compared
to that of PM 1000 in Fig. 11 for 600 C (a) and 850 C (b), respectively. It can be seen that the presence
of c0 precipitates in PM 3030 provides a substantial improvement in fatigue life at both temperatures
due to the higher strength level that retards crack initiation (Albrecht, 1999). The HCF life at a given
stress amplitude is increased in average by about a factor 102103 at both temperatures. Two effects
can be observed by comparing both SN diagrams in Fig. 11. First, the advantage of PM 3030 (R34) is
highest at the lower temperature and decreases as the test temperature increases. Second, the slope of
the SN-curve of PM 3030 (R34) is less steep than that of PM 1000 at the lower temperature of 600 C
(cf. the strength exponents b = 0.05 for R34 and b = 0.093 for PM 1000; Table 3). This means fatigue
life improvement due to additional c0 hardening as compared to solely dispersion strengthened alloys
is more pronounced at low stress amplitudes in the case of low temperature. In contrast, the slope of
the SN-curve of R34 (b = 0.16) is steeper than that of PM 1000 (b = 0.072) at the higher temperature of 850 C (Table 3), which indicates that the increase in fatigue life due to additional c0 hardening

700

700
850C ; 1 Hz

a / MPa

600C ; 1 Hz

a / MPa

R34

R34

R901315

R901315

PM1000
200

103

104

PM1000
10 5

106

107

100

10 3

10 4

10 5

10 6

10 7

2N f / Cycles

2N f / Cycles

Fig. 11. SN diagram of PM 3030 (R34 and R901315) and PM 1000 at 600 C (a) and 850 C (b); R = 1; triangular wave shape.
Table 3
Stress exponent (b) and stress coefcient (rf ) values of PM 3030 batches and PM 1000 at 600 and 850 C
Material

PM 3030 R34
PM 3030 R901315
PM 1000

600 C

850 C

rf =MPa

rf =MPa

0.05
0.088
0.093

1097
1636
1096

0.16
0.231
0.072

2981
3641
518

The values are determined using Fig. 11.

833

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

is most accentuated at high stress amplitudes and decreases as the stress amplitude decreases in the
case of high temperatures. In agreement with the observations in Sections 3.13.3, the impact of temperature and stress amplitude described in this section reects the thermal instability of c0 particles
and the larger impact of time dependent creep in c0 strengthened alloys (Bernstein, 1982) as compared
to solely oxide dispersion strengthened materials.
Fig. 11 also shows the HCF behavior of the isotropic intermediate grain size batch of PM 3030
(R901315). Similar to its quasi-static behavior (Fig. 8), the HCF performance of R901315 is equivalent
to that of the coarse and elongated batch (R34) at 600 C. At 850 C however, grain boundary diffusion
prevails and R901315 performs poorly compared to R34 and even performs worse than PM 1000 at
low stress amplitudes (Fig. 11b). For this reason and due to the additional detrimental effect of accelerated corrosion at grain boundaries, the absolute value of the stress exponent (b = 0.2.31) of
R901315 is the highest of all materials investigated at 850 C.
Fig. 12 shows a different representation of the same HCF results, in which at the ordinate the stress
amplitude (ra) is normalized by the yield strength (r0.2) and the ultimate compressive strength (UCS),
respectively. In such a representation, fatigue lives of both batches of PM 3030 and of PM 1000 are
roughly at the same level at lower temperatures (600 C). This indicates that the quasi-static strength
advantage is nearly proportionally translated into an improvement of fatigue performance due to substantial retardation of crack initiation. At 850 C however, the HCF resistance of both PM 3030 batches
is inferior to that of PM 1000 for a given ra-to-r0.2 or ra-to-UCS ratio due to three main reasons: First,
the strength increase in c0 strengthened alloys is generally accompanied by a decrease in fracture
toughness due to the localization of plastic deformation in few slip bands (Kubin et al., 1995); consequently, although crack initiation is retarded due to higher strength, cracks propagate quickly in PM
3030 during cyclic deformation as soon as they reach the critical size. The second and most important
reason for the weaker fatigue performance of c0 strengthening at higher temperature for a given stress
amplitude-to-quasi-static strength ratio is the more advanced dissolution of c0 particles (see Fig. 7)
during the longer high temperature exposure duration associated to fatigue tests as compared to relatively rapid constant strain rate tests. Also for this reason, the impact of creep damage processes on
fatigue resistance (Rees and Dyson, 1986) is stronger for PM 3030 than for PM 1000. Therefore, the
gain in strength due to the c0 particles is higher under quasi-static loading conditions than the corre-

850C; 1Hz

600C; 1Hz
1

a / Strength

a / Strength

/ 0.2 a / UCS

0.2 a / UCS

a
R3
R901315
PM1000

0.1

0.1
3

10

10

10

2Nf / Cycles

10

10

10

a/
R34
R901315
PM1000
3

104

105

106

107

2Nf / Cycles

Fig. 12. Stress amplitude-to-strength (r0.2 or UCS) ratio life cycles diagram of PM 3030 (R34 and R901315) and PM 1000 at
600 C (a) and 850 C (b); R = 1; triangular wave shape.

834

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

sponding gain in fatigue life. This discrepancy is more accentuated in the case of the ner grain batch
R901315, especially at the higher temperature of 850 C due to premature crack initiation at coarse
precipitates as well as enhanced grain boundary damage processes such as pore formation (Kassner
and Hayes, 2002). This third reason for the particularly poor HCF performance of R901315 will be discussed in the following Section 3.5 using SEM fractography investigations.
3.5. Fracture and fracture resistance
Figs. 1315 compare SEM fractography micrographs of the two technologically most relevant
batches of PM 3030 (R34 and R901315) following HCF tests at 850 C. The dashed line in Figs. 13
and 14 shows the limit between stable fatigue crack propagation and unstable nal fracture, and
the arrow shows the crack initiation location. As can be seen by comparing both micrographs, a relatively small stable crack propagation region is observed on the R34 fracture surface, which is an evidence for the lower fracture toughness discussed in Section 3.4. In contrast, the stable crack

Fig. 13. SEM fractography of PM 3030 R34 after HCF test at 850 C and 1 Hz; crack initiation at sample surface and relatively
small stable crack propagation area; arrow shows crack initiation site.

Fig. 14. SEM fractography of PM 3030 R901315 after HCF test at 850 C and 1 Hz; crack initiation at coarse particle in sample
interior and relatively large stable crack propagation area; arrow shows crack initiation site; area in the rectangular frame is
magnied in Fig. 15.

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

835

Fig. 15. SEM fractography of PM 3030 R901315 after HCF test at 850 C and 1 Hz; the premature crack initiating coarse
yttrium-oxide particle is magnied.

propagation area is relatively large on R901315 fracture surfaces, indicating that the crack tolerance of
this isotropic ner grain batch is higher than that of R34. Therefore, fracture toughness and ductility of
R901315 could be expected to be higher as well.
To closer evaluate this potential, tensile tests have been performed at 850 C to determine the elongation to failure; however, the test results could not meet expectations. Instead, R34 and R901315
show similar elongations to failure under tensile loading, giving no evidence for superior ductility
of R901315 as compared to R34. This result can be explained by the relatively easy crack initiation
at grain boundary defects in this batch. EDX analysis in SEM shows that the most critical defects
are coarse yttriumaluminum-oxide particles (Fig. 15) that are obviously formed during the additional
grain coarsening heat treatment performed in the laboratory on R901315. Furthermore, the strong
concentration of defects and precipitates along grain boundaries of R901315 leads to partially easy
intergranular crack propagation. It can therefore be concluded that eliminating those defects through
a better heat treatment process is expected to potentially lead to a substantial improvement of the
ductility, crack tolerance and consequently fatigue performance of R901315.
4. Conclusions
The quasi-static strength and high cycle fatigue life of the c0 and ODS nickel-base superalloy PM
3030 are described as a function of temperature and assessed to the properties of a solely oxide dispersion strengthened nickel-base superalloy PM 1000 with comparable chemical composition and
grain structure. It is shown that additional c0 precipitation strengthening can substantially improve
both the strength and the fatigue life of ODS superalloys. This advantage is greater for lower temperature, higher stress amplitudes and shorter test durations, for which the impact of the thermal instability of c0 precipitates on mechanical properties is small due to short load and high temperature
exposure durations. At higher temperatures however, the quasi-static strength advantage of PM
3030 is only partially translated into fatigue life improvement due to the thermal instability of c0 precipitates and the decrease in ductility that accompanies c0 strengthening. Hence, it can be concluded
that the technological advantage of additional c0 hardening is limited for temperatures up to around
1160 C, above which complete dissolution of c0 precipitates occurs. This can be assessed in more detail by dividing the relevant temperature range between 600 and 1160 C into two sub-ranges, each
requiring an appropriate design of grain structure. First, in the upper sub-range between 850 C and
1160 C, coarse and elongated, hence cost intensive, grain structures have proven very benecial in
terms of creep rupture response and are widely established (Arzt, 1991). However, the impact of c0
precipitates on improving fatigue life is only moderate, if at all, in this range. Second, in the lower
intermediate sub-range between 600 C and 850 C, combining oxide dispersion and c0 precipitation

836

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

strengthening in nickel-base superalloys provides the greatest strength advantage. In this temperature
range isotropic microstructures with intermediate grain sizes between 10 and 100 lm have been
demonstrated to represent a cost saving alternative with similar mechanical properties as their counterparts with coarse and elongated grains. Furthermore, ner grained materials have the potential to
provide higher ductility and crack tolerance and consequently better fatigue properties. To fully exploit this potential, however, an improvement of the manufacturing process is necessary in order to
further optimize grain, precipitate and dispersoid sizes, as well as eliminate premature crack initiating
coarse yttrium-oxide particles.
Acknowledgements
The fruitful collaboration with Plansee GmbH, particularly with Dr. Frank E.H. Mueller; and with
the Leibniz Institute for Solid State and Materials Research, particularly with Professor Ludwig
Schultz, is greatly appreciated.
References
Albrecht, J., 1999. Comparing fatigue behaviour of titanium and nickel-base alloys. Mater. Sci. Eng. A 263, 176186.
Arzt, E., 1991. Creep of dispersion strengthened materials: a critical assessment. Res. Mech. 31, 399453.
Arzt, E., Ashby, M.F., 1982. Threshold stresses in materials containing dispersed particles. Scripta Mater. 16, 12851290.
Arzt, E., Rsler, J., 1988. The kinetics of dislocation climb over hard particles. 2. Effect of an attractive particle dislocation
interaction. Acta Metall. 36, 10531060.
Arzt, E., Wilkinson, D.S., 1986. Threshold stresses for dislocation climb over hard particles: the effect of an attractive interaction.
Acta Metall. 34, 18931898.
Bayerlein, U., 1991. Zur Ermittlung der Textur- und Gefgeabhngigkeit der elastischen Eigenschaften sowie der
Einkristallkonstanten von Superlegierungen bei hheren Temperaturen. VDI Fortschrifttsberichte, Reihe 5, Nr. 236.
Behr, R., Mayer, J., Arzt, E., 1997. TEM-study of the interaction between superdislocations and dispersoids in Ni3Al alloys. Scripta
Metall. 36, 341345.
Benjamin, J.S., 1970. Dispersion strengthened superalloys by mechanical alloying. Metall. Trans. 1, 29432951.
Benjamin, J.S., Bomford, M.J., 1974. Effect of yttrium oxide volume fraction and particle size on elevated temperature strength of
dispersion strengthened superalloys. Metall. Trans. 5, 615621.
Bernstein, H.L., 1982. An evaluation of four creep-fatigue models for a nickel-base superalloy. In: Low-Cycle Fatigue and Life
Prediction. ASM, Philadelphia, pp. 105134.
Biermann, H., 1999. Ursachen und Auswirkungen der Gerichteten Vergrberung (Flossbildung) in Einkristallinen NickelbasisSuperlegierungen. Habilitation Thesis, University of Erlangen.
Cairns, R.L., Curwick, L.R., Benjamin, J.S., 1975. Grain growth in dispersion strengthened superalloys by moving zone heat
treatments. Metall. Trans. 6A, 179188.
Copley, S.M., Kear, B.H., 1967a. Temperature and orientation dependence of the ow stress in off-stoichiometric Ni3Al (phase).
Trans. Metall. Soc. AIME 239, 977984.
Copley, S.M., Kear, B.H., 1967b. A dynamic theory of coherent precipitation hardening with application to nickel-base
superaloys. Trans. Metall. Soc. AIME 239, 984992.
deMestral, B., Eggeler, G., Klam, H.-J., 1996. On the inuence of grain morphology on the creep deformation and damage
mechanisms in directionally solidied and oxide dispersion strengthened superalloys. Metall. Mater. Trans. 27A, 879890.
Dieter, G.E., 1986. Mechanical Metallurgy. SI Metric ed. McGraw-Hill Book Company, London.
Estrin, Y., Heilmaier, M., Drew, G., 1999. Creep properties of an oxide dispersion strengthened nickel-base alloy: the effect of
grain orientation and grain aspect ratio. Mater. Sci. Eng. A 272 (1), 163173.
Frost, H.J., Ashby, M.F., 1982. Deformation-Mechanism-Maps. The Plasticity and Creep of Metals and Ceramics. Pergamon Press,
Oxford.
Hall, E.O., 1951. The deformation of mild steel: III discussion of results. In: Proceedings of the Physical Society of London, vol.
643. pp. 747753.
Heilmaier, M., 1992. Modellkompatible Berechnung des Kriech- und Zeitstandverhaltens oxid-dispersionsgehrter
Nickelbasisisuperlegierungen. Dissertation Thesis, University of Erlangen.
Heilmaier, M., Reppich, B., 1996. Creep lifetime prediction of oxide dispersion strengthened nickel-base superalloys: a
micromechanically based approach. Metall. Mater. Trans. 27A, 38613870.
Heilmaier, M., Reppich, B., 1999. Particle threshold stresses in high temperatures yielding and creep: a critical review. In: The
Minerals, Metals and Materials Society: Creep Behavior of Advanced Materials for the 21th Century. TMS, Warrendale, pp.
267281.
Heilmaier, M., Leetz, U., Reppich, B., 2001. Order strengthening in the cast nickel-based superalloy IN 100 at room temperature.
Mater. Sci. Eng., 375378.
Hotzler, R.K., Glasgow, T.K., 1982. The Inuence of c0 on the recrystallization of an oxide dispersion strengthened superalloy
MA6000E. Metal. Trans. 13A, 16651674.
Kassner, M.E., Hayes, T.A., 2002. Creep cavitation in metals. Int. J. Plast. 20, 17151748.
Kear, B.H., Wilsdorf, H.G.F., 1962. Dislocation conguration in plastically deformed polycrystalline Cu3Au alloys. Trans. AIME
224, 382386.

M. Nganbe, M. Heilmaier / International Journal of Plasticity 25 (2009) 822837

837

Khraishi, T.A., Yan, L., Shen, Y.-L., 2004. Dynamic simulations of the interaction between dislocations and dilute particle
concentrations in metalmatrix composites (MMCs). Int. J. Plast. 20, 10391057.
Kocks, U.F., 1977. The theory of an obstacle-controlled yield strength. Mater. Sci. Eng. 27, 291298.
Kubin, L.P., Lisiecki, B., Garon, P., 1995. Octahedral slip instabilities in c/c0 superalloys single crystals CMSX and AM3. Philos.
Mag. A 71, 9911009.
Morris, D.G., Gutierrez-Urrutia, I., Muoz-Morris, M.A., 2008. High temperature creep behaviour of an FeAl intermetallic
strengthened by nanoscale oxide particles. Int. J. Plast. 24, 12051223.
Nabarro, F.R.N., Villiers, H.L., 1995. The Physics of Creep: Creep and Creep-Resistant Alloys. Taylor and Francis, London.
Nabarro, F.R.N., Villiers, H.L., 1996. Dislocations in solids. L12 Ordered Alloys, vol. 10. North-Holland Publishing company,
Amsterdam.
Nganbe, M., 2002. Untersuchung und Optimierung der c0 - und Oxiddispersionsgehrteten (ODS) Nickelbasissuperlegierung PM
3030. Dissertation Thesis, University of Dresden.
Nganbe, M., Heilmaier, M., 2004. Modelling of particle strengthening in the c0 and oxide dispersion strengthened nickel-base
superalloy PM3030. Mater. Sci. Eng., 609612.
Petch, N.J., 1953. The cleavage strength of polycrystals. J. Iron Steel Inst. 174, 2528.
Raj, R., Ashby, M.F., 1971. On the grain boundary sliding and diffusional creep. Metall. Trans. 2, 11131127.
Rees, D.W.A., Dyson, B.F., 1986. Deformation and fracture behaviour under combined creep and fatigue. Int. J. Plast. 2, 119.
Reppich, B., 1998. On the attractive particle-dislocation interaction in dispersion-strengthened materials. Acta Mater. 46, 6167.
Spingarn, J.R., Nix, W.D., 1978. Diffusional creep and diffusionally accommodated grain rearrangement. Acta Metall. 26, 1389
1398.
Stevens, J.J., Nix, W.D., 1985. The effect of grain morphology on longitudinal creep properties of INCONEL MA 754 at elevated
temperature. Metall. Mater. Trans. 16A, 13071324.
Stickforth, J., 1986. On stress relaxation, creep, and plastic ow. Int. J. Plast. 2, 347357.
Yashiro, K., Kurose, F., Nakashima, Y., Kubo, K., Tomita, Y., Zbib, H.M., 2006. Discrete dislocation dynamics simulation of cutting
of c0 precipitate and interfacial dislocation network in Ni-based superalloys. Int. J. Plast. 22, 713723.

You might also like