Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/4670931

Yielding and deformation behavior of the single crystal superalloy PWA 1480

Article in Metallurgical Transactions A · February 1987


DOI: 10.1007/BF02646225 · Source: NTRS

CITATIONS READS
150 694

2 authors:

Walter Milligan Stephen Antolovich


Michigan Technological University Georgia Institute of Technology
64 PUBLICATIONS 3,166 CITATIONS 166 PUBLICATIONS 4,080 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Stephen Antolovich on 23 October 2018.

The user has requested enhancement of the downloaded file.


Yielding and Deformation Behavior of
the Single Crystal Superalloy PWA 1480
WALTER W. MILLIGAN and STEPHEN D. ANTOLOVICH

Interrupted tensile tests were conducted to fixed plastic strain levels on (001) oriented single crystals
of the nickel-base superalloy PWA 1480. Testing was done in the range from 20 to 1093 ~ at strain
rates of 0.5 and 50 pct/min. The yield strength was constant from 20 to 760 ~ above which the
strength dropped rapidly and became a strong function of strain rate. The data could be represented
very well by an Arrhenius-type equation, which resulted in three distinct temperature regimes. The
deformation substructures could also be grouped in the same three regimes, indicating that there was
a fundamental relationship between the deformation mechanisms and the activation energies. At low
temperatures, the activation energy for yielding was zero, and the deformation was dominated by y'
shearing by pairs of {lll}a/2(l10) dislocations. At high temperatures, the true activation energy for
yielding was calculated to be 500 kJ/mol, which is indicative of a diffusion-controlled process, and
deformation was dominated by y' by-pass. Intermediate temperatures exhibited transitional behavior.
No currently available precipitation hardening model could adequately describe the behavior observed
in the low temperature regime, due to the observation that penetration into the precipitate was not
rate-limiting at all temperatures. In the high temperature regime, the functional form of the Brown-
Ham by-pass model fit the data fairly well. The results of this study also demonstrated that the initial
deformation mechanism was frequently different from that which would be inferred by examination
of specimens which had been tested to failure.

I. INTRODUCTION Table I. Alloy Composition


NICKEL-base superalloys are being used in single crystal Element Weight Percent
form due to improvements in creep, rupture, and fatigue A1 4.8
resistance over conventionally cast alloys.l PWA 1480 is a Ti 1.3
modem single crystal alloy which is being used in gas tur- Ta 11.9
bine engines as a turbine blade material, 1 and is being con- Cr 10.4
sidered for use in the space shuttle main engine (SSME) as Co 5.3
the turbopump blade material. 2 In this study, mechanisms of W 4.1
monotonic deformation in PWA 1480 were investigated as C (42 ppm)
a function of temperature, strain level, and strain rate. This Ni bal.
information provides a basis for the next phase of the
project, which is the development of monotonic and cyclic
treatment at 1080 ~ for 4 hours followed by rapid cooling,
constitutive models.
and a final aging treatment at 870 ~ for 32 hours followed
In addition to the specific goal of characterizing the mate-
by air cooling.
rial as a part of the SSME program, a further goal was to add
Test specimens were designed and machined in accor-
to the fundamental understanding of deformation processes
dance with ASTM Specification E-8. Two specimens were
in high 3/' volume fraction superalloys.
machined from each bar, and specimens had a 6 mm di-
ameter and a 25 mm gage length.
II. EXPERIMENTAL PROCEDURES B. Mechanical Testing
A. Material Interrupted tensile tests were conducted in order to study
The composition of the alloy is given in Table I. Single deformation mechanisms which operate during yielding.
crystal bars measuring 2.5 cm in diameter and 15 cm in Tests were conducted at 20,705,760, 815,871,927, 982,
length were produced by TRW, Inc. The crystal orientations and 1093 ~ and at strain rates of 8.33 x 10 -5 s -1
were determined by the Laue back reflection X-ray tech- (0.5 pct/min) and 8.33 x 10-3 s -l (50 pct/min). Testing
nique. All bars whose tensile axes were within 10 deg of was done under total strain control using a SATEC comput-
(001) were accepted. The bars were subjected to a heat erized test system which was equipped with a high tem-
treatment procedure consisting of a solutionizing treatment perature extensometer. Specimens were induction heated,
at 1285 ~ for 4 hours followed by rapid cooling, an aging and a temperature gradient not exceeding • ~ was main-
tained over the gage length. After the preprogrammed plas-
tic strain level was reached (nominally 0.1 or 0.3 pct),
WALTER W. MILLIGAN, Graduate Research Assistant, and the specimens were immediately returned to zero load and
STEPHEN D. ANTOLOVICH, Director and Professor, are with the
Fracture and Fatigue Research Lab, School of Materials Engineering, cooled to room temperature.
Georgia Institute of Technology, Atlanta, GA 30332-0245. Samples of the same alloy which were tested to failure
Manuscript submitted February 10, 1986. were supplied by Pratt&Whiteney Aircraft. Although

METALLURGICALTRANSACTIONSA VOLUME 18A, JANUARY 1987--85


the P & W material has a slightly higher A1 + Ti content
than the TRW material, the heat treatments, 7' sizes, and
Y' volume fractions were almost identical. These tests
were conducted at 760, 871,982, and 1093 ~ and a strain
rate of 0.5 pct/min. In addition, two samples which had
been tested under creep conditions were supplied by
Pratt & Whitney. One sample was tested at 871 ~ and
414 MPa, and the other was tested at 1093 ~ and 117 MPa.

C. Metallography
The microporosity, 7' structure, and 7/3" eutectic pools
were documented by optical and scanning electron micros-
copy (SEM), and volume fractions were measured by stan-
dard stereological techniques) Residual dendritic structure
and variations in the 7' size across dendrite boundaries were
also characterized. Thin foils for transmission electron mi-
croscopy (TEM) were prepared by twin-jet electropolishing. (a)
The heat-treated microstructure and the deformation micro-
structures were characterized by using a JEOL 100C TEM
operated at 100 kV. At least four foils were studied at both
strain rates at 20,760,870, 892, and 1093 ~ while at least
two foils were studied at each of the remaining conditions.

IlL RESULTS AND DISCUSSION


A. Heat Treated Microstructure
The microstructure of the alloy is similar to that of other
high Y' volume fraction alloys typified by MAR-M* 200. 4
*MAR-M is a trademark of Martin Marietta Company.

The alloy contains a fine dispersion of ordered, cuboidal 7'


particles in a disordered 7 matrix (Figure l(a)). The 3" size
was fairly uniform, ranging from 0.25 to 1.0/xm, with an
average size of 0.5/xm. There were isolated areas in the
interdendritic regions which contained larger primary y' (b)
particles. In these regions, the average y ' size was about Fig. 1 - - Initial microstructure. (a) TEM micrograph of the y / y ' structure,
1 /zm, and particles up to 1.5/~m were sometimes ob- g = (200). (b) Optical micrograph showing residual dendritic structure,
served. No "hyperfine" 7' was found) The primary y' eutectic areas (A), and micropores (B).
volume fraction was measured to be 55 to 60 pct.
As shown in Figure l(b), the alloy contained a residual
dendritic structure with micropores and large interdendritic In a previous study, Bowman 6 measured the lattice pa-
eutectic pools. The volume fraction of microporosity was
rameter mismatch using extracted 7'. The unconstrained y '
measured to be 0.15 to 0.20 pct. Most pores were spherical,
lattice parameter was found to be 0.28 pct larger than the
with a diameter of less than 50/xm, but several elongated matrix lattice parameter.
pores with a major axis length of up to 250/zm were
present. The volume fraction of eutectic was measured to be
B. Mechanical Behavior
4.5 to 5.0 pct. The maximum eutectic pool size was about
250/zm. 1. Temperature effects
Due to a low carbon content (42 ppm), few carbides were As shown in Table II and Figure 2, the yield strength at
present. The carbides were only occasionally observed 0.05 pct offset was the same at 20 and 705 ~ Above
via TEM in thin foils, and the maximum observed carbide 760 ~ the strength began to drop rapidly, and the strain
size was 0.5/xm. No phases other than y, 7', and carbides rate began to have a strong effect on the strength. The
were found. decrease in strength as a function of temperature was linear
As shown in Figure l(a), the initial dislocation density at 50 pct/min over the entire temperature range, and linear
was extremely low, with most dislocations residing in the at 0.5 pct/min up to 927 ~ Such behavior is typical of
matrix and in the 7 / y ' interfaces. The average dislocation high volume fraction superalloys, and has been documented
density was measured to be 108/cm 2 in the as-heat treated for several similar systems) '7-1~The results are also in rea-
material, and very few dislocations were observed within sonable agreement with those obtained for PWA 1480 in
the y'. another study. H

86--VOLUME 18A, JANUARY 1987 METALLURGICAL TRANSACTIONS A


Table II. Test Data

Strain Elastic Plastic


Sample Temperature Rate Degrees Modulus Strain Offset Yield Strength (MPa)
Number (of) (Pct/min) from (001> (GPa) (Pct) 0.05 Pct 0.1 Pct 0.2 Pct
61-2 20 0.5 6 137 0.14 1010 1015 --
37-2 4 120 0.24 1006 1020 1034
70-1 705 0.5 3 105 0.08 970 -- --
35-2 6 105 0.29 1000 1034 1120
70-2 50 3 105 0.14 970 1000 --
35-1 6 103 0.35 950 970 1034
60-1 760 0.5 5 100 0.11 844 893 --
44-2 10 103 0.26 908 956 1048
60-2 50 5 96 0.14 950 1000 --
44-1 10 105 0.35 949 978 1062
1-1 815 0.5 6 94 0.10 675 696 --
83-1 5 94 0.25 640 683 774
1-2 50 6 98 0.19 942 1006 --
83-2 5 98 0.34 886 957 1083
5-1 871 0.5 4 94 0.09 535 -- --
40-2 8 92 0.24 506 549 612
5-2 50 4 95 0.15 788 823 --
40-1 8 102 0.32 816 865 932
42-1 927 0.5 4 86 0.09 373 -- --
39-1 6 87 0.24 394 422 457
42-2 50 4 90 0.18 640 682 --
39-2 6 94 0.35 675 710 767
7-1 982 0.5 6 91 0.10 302 323 --
63-1 5 81 0.25 299 309 351
7-2 50 6 88 0.19 510 538 --
63-2 5 80 0.34 492 527 583
2-1 1093 0.5 5 71 0.09 168 -- --
37-1 4 71 0.22 172 183 221
2-2 50 5 72 0.20 274 295 310
61-1 6 77 0.38 302 316 330

1200 the strength began to fall at 760 ~ while at the higher strain
rate the strength did not begin to drop until above 815 ~
At constant temperature, the strength was significantly
lower for the slower strain rate. Similar trends have been
[] reported in studies 8'9 of the strain rate dependence of yield-
800 s ing in M A R - M 200 and U D I M E T 115.*
*UDIMET is a trademark of Special Metals Corporation.
V

B ~ 3. Data correlation
~>~ 4 0 0 As indicated by the strain rate and temperature de-
Io B
pendence of the yield strength, yielding at elevated tem-
[] = 0.5%/MIN peratures is a thermally activated process. It is therefore
A = 50%/MIN [] appealing to present the data in the form o f an Arrhenius-
type relationship. Rate-controlling mechanisms may then be
0 I I deduced as a function of temperature. Figure 3 is a plot of
0 400 800 1200 the modulus-normalized yield strength vs inverse tem-

T (~ / perature, which represents an equation of the form


o/E = A[exp(Q'/RT)], [1]
Fig. 2--Yield strength (0.05 pct offset) vs temperature.
where or = yield strength,
E = elastic modulus at temperature, T,
2. Strain rate effects Q' = apparent activation energy,
At 705 ~ and below, there was no effect of strain rate in A = a constant.
the range tested (0.5 to 50 pct/min). At 760 ~ and above, As illustrated by Figure 3, there are three distinct tem-
strain rate became very important. At the lower strain rate, perature regimes when the data are represented in this way:

METALLURGICAL TRANSACTIONS A VOLUME 18A, JANUARY 1987--87


-4
I I
High I Med I Low
I 2, z~ ~ -
I .... --I
~ -5 ~ i & . / "_ -~ i
IJJ ~ I! Ii
\
>,
D
"W -6 - ' :
._1
[] = 0.5~/MIN
z~= 50%/MIN
-7 I I I
7 8 9 10 11
(a)
1/T (K-')
Fig. 3--Arrhenius representation of the 0.05 pct yield strength data.
Slopes were determined by least squares analysis.

(a) At low temperatures (below 760 ~ Q' was equal to


zero, so thermal activation was not a factor in the range of
strain rates tested.
(b) At high temperatures (above 927 ~ at 50 pct/min and
above 815 ~ at 0.5 pct/min), Q' was a constant equal to
50 kJ/mol, and was independent of strain rate. The true
activation energy is calculated and discussed later in
the paper.
(c) At intermediate temperatures, a transition from the low
to the high temperature behavior occurred. It is evident from
the shape of the curves that the transition region boundaries
and functional forms were a strong function of strain rate.

(b)
C. Deformation Substructures
Fig. 4--Typical low temperature deformation structures. (a) #61-2,
Analysis of the deformation substructures resulted in the 20 ~ (001) zone axis multibeam condition. (b) #70-2, 705 ~ (001)
same three temperature regimes as the Arrhenius analysis, zone axis multibeam condition.
and the boundaries of the three regimes were the same. As
discussed below, low temperature deformation at yield was Through systematic analysis of the dislocation Burgers
dominated by y ' shearing, high temperature deformation at vectors and line directions, it was determined that the vast
yield was dominated by y ' by-pass, and intermediate tem- majority of dislocations observed after deformation at low
peratures exhibited a transition from shearing to by-pass. temperatures were of the type {111}a/2(110). Dislocations
1. Low temperatures traveled through the y' as closely-spaced pairs in order to
Deformation substructures at 20, 705, and 760 ~ (high minimize the anti-phase boundary (APB) area created by
strain rate only) were qualitatively similar. The dominant the a/2(110) displacement of the superlal tice. 12This is dem-
deformation mechanism was shearing of the y' by pairs of onstrated by Figure 5, in which those portions of the
a/2(l10) dislocations which were confined to octahedral dislocations within the precipitate are constricted due to
planes. At 20 ~ relatively few dislocations were present at the high APBE, while those portions of the same disloca-
yield, and the spacing between dislocations was large tions which had exited the precipitates are split due to the
(Figure 4(a)). However, the dislocation density at yield was elastic repulsion. Figure 6 demonstrates that the y ' was
significantly higher than it was in the as-heat treated mate- sheared during deformation, as the precipitate exhibits a
rial. There was also evidence of y ' shearing, including the shear offset which is parallel to the projection of the disloca-
presence of dislocation pairs and residual loops within the tions Burgers vector.
y ' . At 705 and 760 ~ the dislocation density at yield was 2. High temperature
very high, and the structure consisted of intense slip bands The boundary for the high temperature region was a func-
which contained closely-spaced dislocations (Figure 4(b)). tion of strain rate. At 0.5 pct/min, high temperature behav-
It is possible that this type of structure may have been ior was dominant at 815 ~ and above, while at50 pct/min,
formed but not observed at 20 ~ due to the possibility of high temperature behavior did not manifest itself until
extremely localized, nonhomogeneous slip. 927 ~ and above.

88--VOLUME 18A, JANUARY 1987 METALLURGICAL TRANSACTIONS A


In contrast to the shearing which was observed at low
temperatures, deformation during initial yielding at high
temperatures occurred primarily by dislocations moving be-
tween and around the precipitates. Figure 7 shows a typical
substructure consisting of dislocation loops left in the y / y '
interface and in the 3' matrix. The stereo pair in Figure 8
clearly shows groups of dislocations weaving between and
wrapping around the y ' precipitates. Dislocations were only
infrequently observed within the y' after interrupted tests
conducted at 0.5 pct/min above 815 ~ Although a few
slip bands were observed to have cut the y ' even at 1093 ~
(at 50 pct/min), the dominant mechanism at low strain
levels at 927 ~ and above was particle by-pass at both
strain rates.
While dislocations were not observed within the y ' after
interrupted tests at the low strain rate, the specimens which
Fig. 5--Micrograph showing splitting of a/2(l10) pairs upon emerging were tested to failure at 871 and 982 ~ contained a high
from the precipitate. #70-1,705 ~ g = (200). density of dislocations within the y ' (Figure 9). This indi-
cates that the first step in deformation was by-pass of the y ' ,
which was followed by shearing of the y ' later in the test.
Shearing occurred only after large increases in the matrix
dislocation density and significant strain hardening had
occurred.
The dislocations which became trapped in the y / y ' inter-
faces were frequently observed to be pure edge dislocations
lying on {011} planes. This phenomenon was observed at
every temperature, and was very common at 982 and
1093 ~ An example is shown in Figure 10(a), where the
dislocations marked by the arrows were interfacial, with an
a/2[101] Burgers vector and a near-[010] line direction.
This characterizes the dislocations as nearly pure edge, lying
on the (101) plane. By following the dislocations into the
matrix, it was determined that they had cross-slipped from
{111} planes to {011} planes at the interface. Pure edge
dislocations lying on {011} planes in the y / y ' interfaces
have been reported in a similar alloy after creep testing at
850 ~ and it was proposed that thermally activated glide
(a) had occurred on {011} planes in the matrix. 13 Our results
suggest that such a mechanism did not occur in this study,
as the dislocations which lay on {011} in the interface were
observed to lie on {111} in the matrix.

(b)
Fig. 6--Sheafing of the 3'' by {lll}a/2(l10) dislocations in #70-1.
(a) Bright-field micrograph, g = [111]. Dark area at left is a carbide.
(b) Dark-field micrograph using the [030] superlattice spot. Burgers vector
of the dislocations is a/2[T01], inclined at 45 deg to the micrograph.
Dislocations are invisible in (b) because g 9b = 0 invisibility criterion Fig. 7--Typical y ' by-pass microstructure after high temperature defor-
is satisfied. mation. # 6 3 - 1 , 9 8 2 ~ g = (200).

METALLURGICAL TRANSACTIONS A VOLUME 18A, JANUARY 1987--89


Fig. 8--Stereo pair showing dislocations wrapping around the 3" precipitates. # 7 - 1 , 9 8 2 ~ g = (liD.

Pure edge dislocations create an elastic strain field with a


dilatational component, 14 so the total energy of the y / y '
interface can be reduced if an edge dislocation of the appro-
priate sign lies in the interface and accommodates some of
the lattice mismatch strain. It should also be noted that {011}
planes can easily accommodate pure edge dislocations with
low index line directions, (100). This provides a driving
force for cross-slip onto {011} planes, which is not normally
observed in fcc alloys.
At 1093 ~ the interfacial dislocations coalesced to form
a homogeneous, regular array after about 0.2 pct plastic
strain (Figure 10(b)). The networks were hexagonal in na-
ture, and consisted of primarily two types of dislocations:
pure edge dislocations lying on {011} planes with (100) line
directions, and mixed dislocations lying on {111} planes
with (110) line directions. This network was extremely sta-
ble. The specimen which was tested to failure (30 pct elon- (a)
gation) exhibited the same type of deformation substructure

(b)
Fig. 1 0 - - Inteffacial arrays developed at very high temperature. (a) At low
strains, segments marked by arrows cross slipped to {011}. #2-1, 1093 ~
Fig. 9 - - H i g h dislocation density after testing to failure at 871 ~ #JA36, ep = 0.09pcI, g = (200). (b) At higher strains, hexagonal arrays devel-
g = (111). oped from structure in (a). #37-1, 1093 ~ ep = 0.22 pct, g = (200).
90--VOLUME 18A, JANUARY 1987 METALLURGICAL TRANSACTIONS A
(a) (a)

(b) (b)
Fig. l l - - D e f o r m a t i o n structures at failure at 1093 ~ (a) Tensile test, Fig. 1 2 - - 7' shearing by the {111} (112) system at 760 ~ and 0.5 pct/min.
#JA38, g = (200). Note by-pass dominated structure and inteffacial net- (a) #44-2, ep = 0.26 pct, g = ( l i D , w (deviation parameter) = 0.1.
works. (b) Creep test, #JA48, g = (220). Note significant y' coarsening (b) #JA34, tested to failure (14 pct elongation), g = (200), w = 0.
and inteffacial networks.

after 0.3 pct strain, after tensile failure, and after creep
as the interrupted test specimen (0.2 pct plastic strain) failure were slight refinements of the interfacial networks
(Figure ll(a)). The failed specimen contained finer net- and coarsening of the y ' .
works and a few dislocations within the 7', but the-sub-
structure was essentially the same as that of the interrupted 3. Intermediate temperatures
test. In addition to a slight refinement of the interfacial Not surprisingly, a transition from shearing to by-pass
networks, the y' did coarsen slightly in localized regions. was observed in the range from 760 to 927 ~ (depending on
The specimens tested under creep conditions at stresses strain rate). Slip bands were observed only at the high strain
equal to 70 to 80 pct of the low strain rate yield strength rate, and the slip band density decreased as the temperature
exhibited the same type of deformation substructures as increased.
those which developed during yielding. Figure 1l(b) shows Similar to observations 15'16in other alloy systems during
the substructure after creep testing at 1093 ~ and 117 MPa. creep at 760 ~ y ' shearing by the {111}(112) slip system
The only difference between the tensile and creep defor- was observed after slow strain rate testing at 760 ~ Ini-
mation was the increased 7' coarsening during creep tially, the only operative deformation mechanism was slip of
(Figure ll(b)). The specimen which was creep tested at a/2(l10) dislocations in the matrix. At about 0.25 pct plas-
871 ~ and 414 MPa developed the same type of interfacial tic strain, however, deformation also began to occur by slip
arrays, but the 7' did not coarsen significantly. of (112) type partials through the 7 ', resulting in the creation
It is evident not only that the deformation during creep of stacking faults (Figure 12). The partial dislocations seen
and yielding was similar, but also that the steady-state defor- in Figure 12(a) were found to have a Burgers vector direc-
mation substructures developed at low plastic strain levels. tion of [112], and the stacking faults were found to be
The only difference between the substructures at 1093 ~ intrinsic, lying on (]-11). A superlattice-intrinsic stacking

METALLURGICALTRANSACTIONS A VOLUME 18A, JANUARY 1987--91


fault (S-ISF) can be formed in Ni3AI by the glide of implies that both the leading and trailing dislocations are in
{ 111}a/3{ 112) partials after the reaction the same particle during shearing, occurs in PWA 1480
(Figure 5). The equation for the increase in the CRSS due
a/2[011] = a/3[112] + a / 6 1 2 1 1 ] [2] to precipitates when "strong" pair coupling occurs is:
occurs. 17 Also, extrinsic/intrinsic "fault pairs" can be
A r c = 0.86{(Tfl/2w)/bd}{(1.28dyo/wT) - 1}la, [31
formed in some alloys by the glide of a/3{112) partial
dislocations. 15a6 In this study, fault pairs were not found, where T = dislocation line tension,
and the faults were intrinsic in nature. It appears, therefore, f = volume fraction,
that the dislocations creating the stacking faults were w = term based on elastic interaction between the
(-[11)a/3[1]2] partials, and Eq. [2] represents the reaction leading and trailing dislocation; magnitude is on
which occurred in the y / y ' interface. the order of 1,
b = dislocation Burgers vector,
D. Discussion of Yielding Behavior d = particle diameter,
Y0 = APBE.
It is desirable to develop micro-mechanical models of the
yielding process, both as an aid in understanding and as a At very large particle sizes, the stress necessary to cause the
guide to alloy development and heat-treatment. From the first dislocation to penetrate the particle can become larger
preceding discussion, it is evident that two distinct models than the CRSS predicted by Eq. [3], and can therefore domi-
of the yielding process in PWA 1480 must be developed: a nate the CRSS. The penetration stress is given by
low temperature model based on y' shearing, and a high A r c = (yo/b) - (T/bro), [4]
temperature model based on ~,' by-pass. The intermediate
temperature regime exhibited complex transitional behav- where r0 = particle radius in slip plane. By using the con-
ior which was very dependent on strain rate, and would stant line tension approximation,
be difficult to model with the presently available experi- T = Gb2/2, [5]
mental results.
where G = shear modulus. Equation [4] becomes
1. Low temperatures
At temperatures below 760 ~ yielding was controlled Arc = (y0/b) - (Gb/2ro). [6]
by "},' shearing. However, the deformation microstructures
at the yield point were not completely independent of tem- By substituting reasonable values for PWA 1480, it is found
perature. At room temperature, the dislocation density was that the penetration stress, Eq. [6], is much larger than the
low, and there were relatively few dislocations within the shearing stress, Eq. [3] (see Appendix A). Therefore, the
y'. At 705 ~ the dislocation density was much higher, and precipitates cannot be treated as weak obstacles, and Eq. [6]
there was a large number of dislocations within the y'. This should predict the CRSS according to this theory.
suggests that the rate-controlling step in the shearing process Copley and Kear 22 have developed a similar expression
changed with temperature. As shown in the following dis- based on observations of MAR-M 200. They showed via
cussion, no currently available model of yielding can fully TEM that penetration of the first dislocation into the particle
explain this behavior. was rate-controlling. A static force balance yielded the fol-
Theories of precipitation hardening based on shearing are lowing equation for the CRSS:
well developed for low volume fraction, underaged super- 1
alloys.~8'19 These models are based on the critical assump- rc = (yo/2b) - (T/bro) + - ~ ( r m + re), [7]
tions that the particles are small enough to be treated as point
obstacles, and that deformation occurs by the glide of where rm = CRSS of the matrix,
"loosely coupled" dislocation pairs. In PWA 1480, neither rp = CRSS of the precipitate.
of these assumptions is valid. Two models which were de-
veloped specifically for high volume fraction systems are Since J/z(r,, + %) may be considered an average Peierls
available. One was developed by Huther and Reppich, 2~ stress when V/ is about 0.5, Eq. [7] may be written as
and the other was developed by Copley and Kear. = Both follows (after substituting the constant line tension ap-
models neglect the contribution of misfit strengthening. In proximation):
PWA 1480 such an assumption is fairly reasonable, due to Arc = (yo/2O) - (GO/2ro). [8]
the relatively low lattice parameter mismatch (Section
III-A). When applied to PWA 1480, both models predict a The only difference between Eqs. [6] and [8] is the factor of
similar form for the critical resolved shear stress (CRSS). ~/2in the APB term, which arises from the fact that Copley
The major reason that most precipitation hardening mod- and Kear considered both the leading and trailing dis-
els are not applicable at high volume fractions is that they locations in their force balance, and it appears that Huther
are based on solid solution strengthening models which were and Reppich considered only the leading dislocation in their
developed for dilute arrays of point obstacles. Huther and analysis. As shown in Appendix A, the Copley-Kear pene-
Reppich 2~ have developed a model which attempts to tration stress predicts the CRSS for PWA 1480 more accu-
overcome this. They derived their model from the solid rately than the Huther-Reppich penetration stress.
solution strengthening model of Schwarz and Labusch, z3 Equation [7] has been used to model the temperature de-
which treats both dilute arrays and concentrated arrays of pendence of the CRSS in MAR-M 200. 7,22It was postulated
obstacles. An additional strong point of the Huther-Reppich that the CRSS was independent of temperature up to 760 ~
model is that it allows for a transition from "weak" dis- because (i) y0 was independent of temperature, (ii) Gb/2ro
location pair coupling to "strong" dislocation pair coupling changed negligibly from 20 to 760 ~ (iii) A(r,, +
above a critical particle size. "Strong" pair coupling, which %) = 0 from 20 to 760 ~ and (iv) the rate controlling
92--VOLUME 18A, JANUARY 1987 METALLURGICAL TRANSACTIONS A
mechanism was penetration at all temperatures. The major interface and the climbing dislocation could reduce the
problem with using this approach to model the temperature climb rate, thus increasing the CRSS. The Brown-Ham
dependence of the yield strength of PWA 1480 is the as- model was derived for mismatch-free precipitates, and this
sumption that penetration into the particle is rate controlling type of interaction is not considered in their model.
at all temperatures. Figures 4 through 6 show that at inter-
mediate temperatures, about 8 to 10 dislocations were able
IV. SUMMARY AND CONCLUSIONS
to penetrate the y ' particle before the first dislocation was
able to completely shear it. This implies that the drag stress A. Mechanical Behavior
of the particle was probably more important than the pene-
tration stress, and therefore a new type of temperature- The yield strength trends of PWA 1480 as a function of
dependent model must be developed. The present authors temperature and strain rate were similar to other high vol-
are investigating this possibility. ume fraction superalloys. The strength was constant and
independent of strain rate up to 760 ~ above which the
2. High temperatures strength dropped rapidly and became a strong function of
At high temperatures, plastic flow occurred during yield- strain rate.
ing by y' by-pass. Under these conditions, the rate-limiting The yield strength vs temperature was correlated very
step in many systems is diffusion-controlled climb. ~8'z4-28 well by an Arrhenius-type relationship. This resulted in
The true activation energy for yielding at high temperature three temperature regimes, and the boundaries were a func-
in PWA 1480, which is calculated to be 500 kJ/mol in tion of strain rate. At low temperatures, the yield strength
Appendix B, indicates that self-diffusion was probably the was independent of temperature. At high temperatures, the
rate limiting step in the yielding process (see discussion in slope of the Arrhenius curve was independent of strain rate,
Appendix B). Thus, both the activation energy analysis and and the activation energy for yielding was indicative of a
the microstructural evidence for by-pass provide a strong self-diffusion controlled process. At intermediate tem-
argument for dislocation climb as the rate-limiting step dur- peratures, the shape of the Arrhenius curve was a function
ing yielding, and a model of yielding based on y' by-pass of strain rate, and represented a transition from low to high
which is controlled by climb is appropriate. temperature behavior.
Brown and Ham TM have proposed such a model for a
system of cuboidal precipitates. The model is derived from B. Dominant Deformation Mechanisms
an analysis of vacancy and interstitial diffusion in the vicin- The deformation substructures at yield can be divided into
ity of a climbing edge dislocation, under the assumption that the same three temperature regimes. At low temperatures,
diffusion controls the climb rate. Although the model does deformation at yield occurred by shearing of the y ' by pairs
not quantitatively predict the actual CRSS except for low of a/2(l10) dislocations on {111} planes. At high tem-
volume fraction alloys, its functional form with respect to peratures, deformation was dominated at yield by particle
temperature does agree with the results of the present study. by-pass by single a/2(l10) dislocations. At intermediate
The CRSS as a function of temperature is given by temperatures, a transition from shearing to by-pass oc-
Zc = Zo + k, Qo - k2r[ln(k3k)[, [9] curred. Since classification of the deformation substructures
resulted in the same three temperature regimes as the
where r0 = threshold stress which depends on the type of Arrhenius analysis, it is clear that there was a fundamental
microstructure, relationship between deformation mechanisms and activa-
Qo = activation energy of the rate controlling dif- tion energies.
fusion process,
k's = constants which include microstructural C. Detailed Substructures
parameters. As the temperature was increased from 20 to 760 ~ the
Equation [9] predicts a linear decrease in CRSS with tem- dislocation density at yield increased, and many dislocations
perature, which is observed except above 982 ~ at the slow were within the y ' when the test was interrupted at 705 and
strain rate (Figure 2). However, it must be noted that 760 ~ At 760 ~ and a strain rate of 0.5 pct/min, intrinsic
Eq. [9] is not applicable at the high strain rate until above stacking faults were produced in the y' by shearing by
927 ~ even though the linear relationship starts as low as a/3(112) partial dislocations. At temperatures above about
815 ~ This is clear because the transition region at the 870 ~ many dislocations cross-slipped from {111} to {011}
higher strain rate did not end until 927 ~ and y ' sheafing planes in the )'/3" interface. Most {011} plane dislocations
was important in the transition regime. were found to be nearly pure edge in character, which allows
It is not clear why the yield strength vs temperature re- the dislocations to accommodate the lattice mismatch strain
lationship deviates from lineafity above 982 ~ at the slow and thereby reduce the interfacial energy. At 982 and
strain rate. One possibility is that the threshold stress is 1093 ~ this leads to the formation of stable hexagonal
being approached and the curve is reaching an asymptotic networks in the interface. These interfacial networks were
limit. Brown and Ham argue that there is always a finite also developed during creep at 871 and 1093 ~ In this
yield stress, which is due to the force necessary tO create temperature regime, there was very little difference between
additional dislocation line length during the climb process. the deformation substructures at yield, after tensile failure,
Another possible reason for the deviation from lineafity is and after creep failure. At 815 to 982 ~ however, the
related to the deformation mechanisms: the temperature substructures developed during tensile testing were a strong
where the deviation from linearity begins to occur is the function of strain level: by-pass was observed at yield, while
same temperature at which the {011} interfacial dislocations shearing was evident later in the test. y' coarsening oc-
became common and the interfacial networks began to de- curred during slow strain rate tensile testing at 1093 ~ and
velop. It is feasible that the attractive force between the creep testing at 871 and 1093 ~
METALLURGICAL TRANSACTIONS A VOLUME 18A, JANUARY 1987--93
D. Modeling o f Yielding The actual CRSS is about 400 MPa. This value is
predicted more accurately by the Copley-Kear penetra-
To model the yielding behavior, the same three tem- tion stress, which resulted22 in a CRSS of 420 MPa for
perature regimes must be considered. At low temperatures,
MAR-M 200.
a model based on y' shearing is needed, and two applicable
models are available. When applied to PWA 1480, both
APPENDIX B
models predict the CRSS based on the premise that pene-
tration into the particle is the rate-limiting step. However, Calculation of the true activation energy for yielding
our results show that penetration was not the rate-limiting The apparent activation energy for yielding at high tem-
step in PWA 1480 at 705 and 760 ~ so a new model must peratures was calculated to be 50 kJ/mol by Eq. [1]. The
be developed. At high temperatures, a model based on y' true activation energy, which should be physically meaning-
by-pass is needed. The deformation substructures and calcu- ful, may be derived in the following way: since the defor-
lated activation energies imply that the theory of Brown and mation mechanisms which operated during tensile yielding
Ham was valid, and the functional form of the model fit and steady-state creep were the same, the yielding data may
the data well. However, their model must be modified at be correlated by a steady-state creep equation. However,
very high temperatures in order to be fully applicable to this approach does involve some fundamental assumptions:
PWA 1480. The intermediate temperature regime exhibited
complex transitional behavior which was very dependent on (i) The plastic strain rate is constant during yielding.
strain rate, and would be difficult to model with the pres- (ii) At high stresses and very high temperatures, primary
ently available data. creep (during yielding) may be ignored. Or
(iii) At high stresses and very high temperatures, the rate-
limiting step during primary and steady-state creep is
APPENDIX A the same.
Calculation of the critical stresses Although assumptions (ii) and (iii) are speculative, some
in the Huther-Reppich model evidence in their support is available. Several studies z8-3'
have shown that the primary creep strain produced in (001)
According to the Huther-Reppich model, the CRSS for a
oriented single crystal or D.S. superalloys is almost negli-
superalloy with overaged, spherical particles which are cut
gible at temperatures above about 850 ~ Also, in a study
by strongly coupled dislocation pairs is given by Eq. [3]:
of IN738" which was creep tested at various stress levels at
A~-c = { 0 . 8 6 ( T f m w ) / ( b d ) } { 1 . 2 8 ( d y o / w T ) - 1}'/2, *IN is a trademark of the INCO family of companies.
[A1] 850 ~ the primary creep strain was reduced from 1 pct at
where the terms are defined in the text. The stress to cause 200 MPa to almost zero at 400 MPa. 28 This indicates that
the first dislocation to penetrate the precipitate is given by primary creep may be negligible at the high stresses which
Eq. [6]: develop during yielding. In support of assumption (iii),
several studies have shown that the deformation mechanism
A z c = ( y o / b ) - (Gb/2ro). [A2] at high temperatures in alloys similar to PWA 1480 is the
Since by-pass did not occur at low temperatures, and since same during primary and steady-state creep. 32'33Perhaps the
the two mechanisms are essentially in series, the largest of strongest support for this argument lies in the similarity of
the two stresses predicted by Eqs. [A1] and [A2] will be the the deformation substructures which developed during
CRSS. The following values were measured for PWA 1480: steady-state creep and yielding. On the other hand, creep
data for PWA 1480 indicate that primary creep may not be
f= 0.6 negligible, 34 and several studies have indicated that the
amount of primary creep strain is increased as the stress is
b = 2.5 x 10- s c m
increased. 34'35 Therefore, it appears that assumption (iii) is
d = 5 x 10-5 cm the strongest.
The steady-state creep rate is represented by: 36
The value of G can be estimated as GNi parallel to
{111}(110), which is 57 GPa at 20 ~ 22 The values of ~s = A ( ~ n exp(-Qc/RT), [B1]
Y0 and w have been estimated for similar alloys, and will where trc = applied stress,
be assumed to be close to the values for PWA 1480: the A, n = material dependent constants,
value of T0 for MAR-M 200 has been estimated2z to be Qc = activation energy for creep.
160 dynes/cm, and the value of w has been estimated to be
2.8 for NIMONIC 105" by Reppich et al. 2, Under the assumptions outlined above, the activation energy
for creep and yielding should be the same, and the yield
*NIMONIC is a trademark of the INCO family of companies. stress can be substituted for the creep stress. The resulting
Using the constant line tension approximation, the value equation for the yield stress is:
of the sheafing stress, Eq. [A1], is 230 MPa, while the ln(or/E ) = (ay/nn)(I/T) q- { l n ( k s / A ) } / n . [B2]
value for the penetration stress, Eq. [A2], is 620 MPa. Even
when G is varied between 50 and 80 GPa, and Y0 is varied From Eqs. [1] and [B2], the activation energy for yielding
between 100 and 200 dynes/cm, the sheafing stress never can be calculated from the apparent activation energy for
exceeds 75 pct of the penetration stress. Thus, the pene- yielding:
tration stress is always larger than the shearing stress, and Qy = n " a ' , [B3]
the penetration stress controls the CRSS in the Huther-
Reppich model. where n is the stress exponent in Eq. [B 1]. From published
94--VOLUME 18A, JANUARY 1987 METALLURGICAL TRANSACTIONS A
creep data on PWA 1480 at 871 ~ 37 the stress exponent Bricknell, W. B. Kent, and J. E Radavich, eds., AIME, Warrendale,
has a value of about 10, which is within the range of values PA, 1984, pp. 105-14.
12. P.A. Flinn: Trans. TMS-AIME, 1960, vol. 218, pp. 145-54.
for similar alloys. 26'28'33'38'39By Eq. [B3], the activation en- 13. C. Carry and J. L. Strudel: Scripta Metall., 1975, vol. 9, pp. 731-36.
ergy for yielding is calculated to be 500 kJ/mol, which is 14. J. Weertman and J.R. Weertman: Elementary Dislocation Theory,
well within the range of activation energies for steady-state MacMillan, New York, NY, 1964, pp. 32-35.
creep in similar alloys. 28'32'33'38'4~ 15. G.R. Leverant and B.H. Kear: Metall. Trans., 1970, vol. 1,
By using the effective stress approach, 44 it has been dem- pp. 491-98.
16. B.H. Kear, J.M. Oblak, and A. E Giamei: Metall. Trans., 1970,
onstrated that the true activation energy for creep in these vol. 1, pp. 2477-86.
alloys is very close to the activation energy for self- 17. P. Veysierre, J. Douin, and P. Beauchamp: Phil. Mag. A, 1985,
diffusion. 26'3s'42'43This indicates that the rate-limiting step in vol. 51, pp. 469-78.
creep deformation in these alloys is diffusion-controlled. 18. L.M. Brown and R. K. Ham: in Strengthening Methods in Crystals,
A. Kelly and R.B. Nicholson, eds., Wiley, New York, NY, 1971,
Since the activation energy for yielding in PWA 1480 was pp. 9-134.
in the range of creep activation energies for similar alloys, 19. D. Raynor and J. M. Silcock: Metal Sci. J., 1970, vol. 4, pp. 121-30.
it appears that the yielding process in PWA 1480 at high 20. B. Reppich: Acta MetaU., 1982, vol. 30, pp. 87-94.
temperatures was diffusion controlled as well. 21. B. Reppich, P. Schepp, and G. Wehner: Acta Metall., 1982, vol. 30,
pp. 95-104.
22. S.M. Copley and B.H. Kear: Trans. TMS-AIME, 1967, vol. 239,
ACKNOWLEDGMENTS pp. 984-92.
23. R.B. Schwarz and R. Labusch: J. Appl. Physics, 1978, vol. 49,
This work was supported by NASA-Lewis Research pp. 5174-87.
Center under Grant NAG3-503. The authors wish to thank 24. G.S. Ansell and J. Weertman: Trans. TMS-A1ME, 1959, vol. 215,
pp. 838-43.
the grant monitors, Drs. Robert C. Bill and Rebecca A. 25. J.H. Hausselt and W.D. Nix: Acta Metatl., 1977, vol. 25,
MacKay, for their interest and helpful discussions. We pp. 1491-1502.
would also like to thank Gustav Swanson and Dr. David 26. C. Carry and J.L. Strudel: Acta Metall., 1978, vol. 26, pp. 859-70.
Duhl of Pratt & Whitney Aircraft for supplying the failed 27. P.J. Henderson and M. McClean: Acta Metall., 1983, vol. 31,
creep and tensile specimens and for useful comments during pp. 1203-19.
28. r . A . Stevens and P. E.J. Flewitt: Acta Metall., 1981, vol. 29,
the course of the work. pp. 867-82.
29. B.H. Kear and B.J. Piearcey: Trans. TMS-AIME, 1967, vol. 239,
pp. 1209-15.
REFERENCES
30. C. Carry and J. L. Strudel: Acta Metall., 1977, vol. 25, pp. 767-77.
1. M. (3ell, D. N. Duhl, and A. E (3iamei: in Superalloys 1980, Proc. 4th 31. A. E Giamei: Air Force Office of Scientific Research Report FR-
Intl. Symposium on Superalloys, J.K. Tien, S.T. Wlodek, H. 12637, 1979.
Morrow, M. (3ell, and (3. Manrer, eds., ASM, Metals Park, OH, 32. G.R. Leverant, B.H. Kear, and J. M. Oblak: Metall. Trans., 1973,
1980, pp. 205-14. vol. 4, pp. 355-62.
2. W.T. Chandler: NASA CR-174729, 1983. 33. G. Jianting, D. Ranucci, E. Picco, and P. M. Strocchi: Metall. Trans.
3. E.E. Underwood: Quantitative Stereology, Addison-Wesley, A, 1983, vol. 14A, pp. 2329-35.
Reading, MA, 1970, pp. 23-44. 34. D.N. Duhl: private communication, Pratt and Whitney Aircraft, E.
4. B.J. Piearcey, B.H. Kear, and R.W. Smashey: Trans. ASM, 1967, Hartford, CT, 1985.
vol. 60, pp. 634-45. 35. G.A.. Webster and B.J. Piearcey: Metal Sci. J., 1967, vol. 1,
5. P. Beardmore, R. (3. Davies, and T. L. Johnston: Trans. TMS-AIME, pp. 97-104.
1969, vol. 245, pp. 1537-45. 36. A.K. Mukherjee, J.E. Bird, and J.E. Dorn: Trans. ASM, 1969,
6. R.R. Bowman: M.S. Thesis, Georgia Institute of Technology, Atlanta, vol. 62, pp. 155-79.
(3A, 1986. 37. G.A. Swanson, I. Linask, D. Nissley, T. P. Norris, T. G. Meyer, and
7. S.M. Copley, B. H. Kear, and (3. M. Rowe: Mater. Sci. Eng., 1972, K.P. Walker: NASA CR-174952, 1985.
vol. 10, pp. 87-92. 38. S. Purushothaman and J.K. Tien: Acta Metall., 1978, vol. 26,
8. G.R. Leverant, M. Gell, and S. W. Hopkins: Mater. Sci. Eng., 1971, pp. 519-28.
vol. 8, pp. 125-33. 39. T.E. Howson, D. A. Mervyn, and J. K. Tien: MetaU. Trans. A, 1980,
9. R.R. Jensen, T.E. Howson, and J. K Tien: in Superalloys 1980, vol. llA, pp. 1609-16.
Proc. 4th Intl. Symposium on Superalloys, J. K. Tien, S.T. Wlodek, 40. M.V. Nathal and L.J. Ebert: Metall. Trans. A, 1985, vol. 16A,
H. Morrow, M. Gell, and G. Maurer, eds., ASM, Metals Park, OH, pp. 427-39.
1980, pp. 679-88. 41. W.J. Evans and G. E Harrison: MetalSci., 1976, vol. 9, pp. 307-13.
10. M.V. Nathal, R.D. Maier, and L.J. Ebert: Metall. Trans. A, 1982, 42. M. McClean: Proc. R. Soc. London A, 1980, vol. 371, pp. 279-94.
vol. 13A, pp. 1767-74. 43. M. McClean: Proc. R. Soc. London A, 1980, vol. 373, pp. 93-109.
11. D.M. Shah and D.N. Duhl: in SuperaUoys 1984, Proc. 5th Intl. 44. C.N. Ahlquist and W.D. Nix: Acta Metall., 1971, vol. 19,
Symposium on Superalloys, M. Gell, C.S. Kortovich, R.H. pp. 373-85.

METALLURGICALTRANSACTIONS A VOLUME 18A, JANUARY 1987--95

View publication stats

You might also like