You are on page 1of 16

Strengthening Mechanisms in Polycrystalline Multimodal

Nickel-Base Superalloys
R.W. KOZAR, A. SUZUKI, W.W. MILLIGAN, J.J. SCHIRRA, M.F. SAVAGE,
and T.M. POLLOCK

Polycrystalline c-c¢ superalloys with varying grain sizes and unimodal, bimodal, or trimodal
distributions of precipitates have been studied. To assess the contributions of specific features of
the microstructure to the overall strength of the material, a model that considers solid-solution
strengthening, Hall–Petch effects, precipitate shearing in the strong and weak pair-coupled
modes, and dislocation bowing between precipitates has been developed and assessed. Cross-
slip-induced hardening of the Ni3Al phase and precipitate size distributions in multimodal
microstructures are also considered. New experimental observations on the contribution of
precipitate shearing to the peak in flow stress at elevated temperatures are presented. Various
alloys having comparable yield strengths were investigated and were found to derive their
strength from different combinations of microconstituents (mechanisms). In all variants of the
microstructure, there is a strong effect of antiphase boundary (APB) energy on strength.
Materials subjected to heat treatments below the c¢ solvus temperature benefit from a strong
Hall–Petch contribution, while supersolvus heat-treated materials gain the majority of their
strength from their resistance to precipitate shearing.

DOI: 10.1007/s11661-009-9858-5
 The Minerals, Metals & Materials Society and ASM International 2009

I. INTRODUCTION then required to determine the extent of the improve-


ment. Although this process has resulted in many useful
POLYCRYSTALLINE nickel-base superalloys are alloys, it is expensive and time consuming; the design
capable of maintaining strength in excess of 1 GPa at process could benefit from a greater contribution of
temperatures up to 800 C. Because of this, these process and property modeling.
superalloys are widely used in aircraft engines and Commercial polycrystalline superalloys that are used
power generation turbines as turbine blade and disk for turbine disk alloys possess complex microstructures
materials. The ever-present drive to improve perfor- that are optimized for a range of mechanical proper-
mance and efficiency by operating engines at increas- ties.[1–4] Typical premium wrought nickel-base superal-
ingly higher temperatures has led to the continual loys consist of a c matrix that contains up to three
development of new alloys that retain their strength at different size distributions (approximately 0.01, 0.1, and
higher temperatures. 1.0 lm in diameter) of L12 (ordered fcc) c¢ precipitates.
Traditionally, in the design of new superalloys, the The smaller two populations of precipitates are coherent
compositions of existing alloys and the specific features with the c matrix, while high-angle boundaries exist
of processing are incrementally changed to achieve between the micron-sized c¢ precipitates and the c
improved properties. Extensive mechanical testing is matrix. Due to the presence of ordered Ni3Al precipi-
tates, the strength of these alloys often increases between
R.W. KOZAR, formerly Graduate Student, with the Department of room temperature and 800 C. These nickel-base super-
Materials Science and Engineering, University of Michigan, Ann Arbor, alloys also have excellent creep properties.[5,6]
MI 48109, is Senior Engineer, with Materials Technology, Bettis Atomic Given the presence of ordered precipitates and the
Power Laboratory, West Mifflin, PA 15122. Contact e-mail: rwkozar88 exceptional strength of this class of materials, superal-
@yahoo.com A. SUZUKI, formerly Post Doctoral Research Fellow,
with the Department of Materials Science and Engineering, University
loys have been the subject of a range of analytical
of Michigan, Ann Arbor, MI 48109, is Metallurgist, with Ceramic and models for strengthening.[7–13] A feature of virtually all
Metallurgy Technologies, GE Global Research, Niskayuna, NY 12309. of these models is that they consider simple microstruc-
W.W. MILLIGAN, Professor, is with the Department of Materials tures in which only one or two strengthening mecha-
Science and Engineering, Michigan Technological University, Hought- nisms are active. These models are not directly
on, MI 49931. J.J. SCHIRRA, Manager, is with the Aircraft Group,
Pratt & Whitney, East Hartford, CT 06108. M.F. SAVAGE, formerly applicable to the complex microstructures present in
Staff Engineer, with Materials and Processes Engineering, Pratt & the commercial forms of these premium materials, in
Whitney, East Hartford, CT 06108, is Manager, Group Strategy and which many strengthening mechanisms are active con-
Development, Pratt & Whitney, East Hartford, CT 06108. T.M. currently. The goal of this research was to quantitatively
POLLOCK, L.H. and F.E. Van Vlack Professor, is with Materials
Science and Engineering, University of Michigan, Ann Arbor, MI
assess the influence of a range of microstructural
48109. features on strength across a range of temperatures.
Manuscript submitted September 17, 2004. This was accomplished by the development of a model
Article published online May 28, 2009

1588—VOLUME 40A, JULY 2009 METALLURGICAL AND MATERIALS TRANSACTIONS A


that considers all relevant strengthening mechanisms for
a given microstructure, integrating and building on
previous analyses of strengthening in nickel-base alloys.
Transmission electron microscopy (TEM) was used
to investigate the operation of specific mechanisms and
to discern transitions in deformation modes unique to
these multimodal microstructures. Finally, the relation-
ship between specific features of the microstructure and
their overall contribution to strength is assessed with the
model and is compared to experimental data.

II. EXPERIMENTAL MATERIALS


AND PROCEDURES
A. Experimental Materials
The microstructural variation induced by commer-
cially relevant regimes of processing and heat treatment of
the nickel-base alloy IN100 motivated the experiments
Fig. 1—Typical processing for superalloy disk material.
and modeling. Experiments on IN100 and an alloy with a
modified chemistry but a microstructure similar to IN100
were used for validation of the yield strength model. The 2 hours with a 1-hour stabilization at 982 C, followed
materials were chemically and microstructurally charac- by gas fan cooling and then an 8-hour aging step at
terized in detail by Wusatowska-Sarnek et al.[14] using 732 C followed by air cooling. Heat treatments that
transmission electron microscopy (TEM), scanning elec- vary from this will be presented as appropriate.
tron microscopy (SEM), and chemical analysis tech-
niques. For the IN100 alloy, the phase compositions were
measured by electrochemical extraction with inductively B. Microstructure
coupled plasma analysis. The volume fractions of the Figures 2(a) through (c), (d) and (e), and (f) and (g)
phases were measured by SEM and TEM analysis and this show the microstructures at different scales of IN100
information was incorporated into a thermodynamic given the standard subsolvus heat treatment, the IN100
database that was subsequently used to calculate the given a supersolvus heat treatment, and the supersol-
phase compositions in the modified chemistry alloy. The vus alloy with a modified chemistry, respectively.
compositions of these two alloys along with the compo- Figures 2(a) and (b) are courtesy of Wusatowska-
sitions of the c and c¢ phases for IN100[14] are summarized Sarnek et al.[14] Figures 2(d) through (g) were prepared
in Table I. For the modified chemistry alloy, changes in using an etchant consisting of 50 vol pct lactic acid,
the c and c¢ compositions were calculated on the basis of 30 vol pct HCl, and 20 vol pct HF. Figure 2(c) was
expected (small) differences in partitioning of the elements acquired using a PHILIPS* STEM CM12 transmission
to the two phases.
All materials were subjected to standard powder
metallurgy hot compaction, extrusion, and isothermal *PHILIPS is a trademark of Philips Electronic Instruments Corp.,
forging procedures. Following this, the materials were Mahwah, NJ.
subjected to several heat treatments, in order to obtain a
variety of microstructures. The typical heat treatment, electron microscope. Foils for the TEM were ground to
shown schematically in Figure 1 for IN100, consisted of a thickness of 0.12 mm in three stages, using 320-, 600-,
a subsolvus or supersolvus heat treatment (for IN100, and 1200-grit SiC papers sequentially. The samples were
the c¢ solvus temperature is 1185 C), followed by then twin-jet electropolished with a solution of 340 mL
stabilization and aging heat treatments. These three heat methanol, 50 mL perchloric acid, 45 mL distilled water,
treatments establish the final microstructure of the alloy. and 65 mL butyl cellusolve at –40 C with a voltage of
Most of the samples in this study were subjected to a 20 to 25 V.
standard subsolvus heat treatment at 1143 C for

Table I. Chemical Composition of Alloys and Their Microconstituents (Weight Percent)

Alloy/Phase Designation Ni Al Cr Co Mo Ti V Fe C Zr B Y
IN100 56.0 4.9 12.3 18.3 3.3 4.3 0.70 0.10 0.06 0.02 0.02 —
c phase, IN100 38.7 2.25 24.5 27.8 5.73 0.93 0.05 — — — — —
c¢ phase, IN100 71.8 7.06 2.59 8.94 1.42 6.97 1.23 — — — — —
Modified alloy 52.9 4.5 14.5 20.0 3.6 3.6 0.70 0.10 0.08 0.02 0.02 0.07
c phase, modified alloy 36.0 1.83 27.2 28.9 5.31 0.66 0.05 — — — — —
c¢ phase, modified alloy 69.7 7.16 3.37 10.7 1.64 6.14 1.34 — — — — —

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 40A, JULY 2009—1589


have an average size of 1.43 lm. The primary c¢
precipitates are not isolated and high-angle grain
boundaries often exist between adjacent primary c¢
precipitates.
In Figure 2(b), smaller secondary c¢ precipitates can
be resolved within the c matrix. Secondary c¢ precipi-
tates nucleate and grow during cooling following the
subsolvus heat treatment and in the subsequent stabil-
ization cycle. The sample in Figures 2(a) and (b) had an
average secondary c¢ diameter of 170 nm. Tertiary c¢
precipitates form in the regions between the secondary c¢
precipitates during the cooling from the stabilization
cycle and final lower temperature aging step; they can be
seen in Figure 2(c). Tertiary c¢ precipitates may have an
average size as large as 20 nm, but for the standard heat
treatment shown in Figure 2(c), the average size is
12 nm.
The supersolvus heat treatment of IN100 dissolves the
primary c¢, resulting in a bimodal distribution of
precipitates for a series of heat treatments similar to
those described earlier. Figures 2(d) and (e) shows the
microstructure of IN100 after a supersolvus heat treat-
ment at 1205 C for 2 hours and subsequent stabiliza-
tion and aging at 982 C for 1 hour and 732 C for
8 hours, respectively. As with subsolvus IN100, the
microstructure of supersolvus IN100 consists of 40 pct c
matrix and 60 pct c¢ precipitates by volume, with the c
matrix containing two distributions of c¢ precipitates.
The average grain size of the supersolvus IN100
shown in Figure 2(d) is 34.4 lm, which is an order of
magnitude larger than that of the subsolvus IN100.
Figure 2(d) also illustrates the absence of primary c¢
precipitates in supersolvus IN100. In Figure 2(e), sec-
ondary c¢ precipitates can be resolved within the c
matrix, because they have an average size of 340 nm and
a higher volume fraction than in subsolvus IN100.
Tertiary c¢ also exists in supersolvus IN100; it has a
similar average size but a higher volume fraction in
comparison to subsolvus IN100 (Figure 2(e)).
The microstructure of the modified chemistry alloy is
similar to supersolvus IN100, except that it has a lower
volume fraction of c¢ precipitates (53 pct). The modified
chemistry alloy, shown in Figure 2(g), has an average c
Fig. 2—Microstructure of (a) through (c) subsolvus IN100, (d) and grain size of 26.8 lm, which is smaller than the
(e) supersolvus IN100, and (f) and (g) supersolvus modified chemis-
try alloy. (a) and (b) Courtesy of Wusatowska-Sarnek et al.[14] (c) supersolvus IN100. Figure 2(g) shows that the second-
through (g) Etched according to Ref. 14. ary c¢ precipitates are only 110 nm, considerably smaller
than in the supersolvus IN100 alloy, due in part to the
lower solution heat-treatment temperature.
The subsolvus heat treatment for IN100 results in a Three additional unimodal and bimodal microstruc-
trimodal distribution of precipitates. The microstructure tures established by the supersolvus heat treatment of
of subsolvus IN100, shown in Figures 2(a) through (c), unmodified IN100 were studied. These were as follows: a
consists of 40 to 45 pct c matrix and 55 to 60 pct c¢ unimodal microstructure of secondary c¢ (mean preci-
precipitates by volume. The c matrix, appearing as light pitate diameter of 180 nm) was achieved by supersolvus
gray in Figure 2(a), contains two distributions of heat treatment at 1205 C for 2 hours, followed by
smaller c¢ precipitates; for this sample, the grain size is cooling at 25 C/min, then stabilization heat treatment
3.1 lm. Also appearing in Figure 2(a) as the darker at 990 C for 1 hour; a bimodal microstructure of
constituent are large c¢ precipitates, referred to as secondary c¢ (mean precipitate diameter of 180 nm) and
primary c¢. This population of Ni3Al is large in size, tertiary c¢ was achieved by supersolvus heat treatment at
due to the fact that it is not solutionized and coarsens 1205 C for 2 hours, followed by cooling at 25 C/min,
during the subsolvus thermomechanical processing and then stabilization and aging heat treatments at 990 C
final heat treatment. In this sample, these primary c¢ for 1 hour and at 732 C for 8 hours, respectively;
precipitates comprise 20 pct of the microstructure and and a bimodal microstructure of secondary c¢ (mean

1590—VOLUME 40A, JULY 2009 METALLURGICAL AND MATERIALS TRANSACTIONS A


precipitate diameter of 564 nm) and tertiary c¢ was According to the Hall–Petch relation,[16] the extent of
achieved by supersolvus heat treatment at 1205 C for grain size hardening can be described by
2 hours, followed by cooling at 10 C/min, then stabil-
ization and aging heat treatments at 990 C for 1 hour r ¼ ro þ ky d1=2
m ½1
and at 732 C for 8 hours, respectively.
where r is the yield strength, ro is the friction stress, ky is
the Hall–Petch coefficient, and dm is the average grain
C. Mechanical Testing size. When considering measurements of the Hall–Petch
Tensile testing was performed to measure the yield constant ky for pure Ni and a range of superalloys,[17–19]
strength and strain-rate sensitivity of experimental a value of 750 MPa lm½ was used in all analyses
materials at temperatures ranging from 22 C to reported here.
760 C. Testing was conducted on a screw-type Instron The form of the equation used in the model was
5582 (Instron, Norwood, MA). Strain measurements modified slightly to account for the fact that the
were taken with a linear variable displacement trans- mechanism does not provide strength to the entire
ducer (LVDT). The test temperature was regulated to microstructure, but only to the volume fraction f
±3 C in a resistance-heated furnace. Cylindrical sam- occupied by the grains of interest. Also, ro was omitted,
ples with a 4.09-mm-diameter gage section were strained because it is accounted for by other mechanisms
to within the range of 0.2 to 0.5 pct past yielding at a considered separately in the model. Including these
rate of 5 9 105/s; the strain rate was then increased to modifications, the Hall–Petch contribution is
5 9 104/s. Strain-rate sensitivity was calculated using Dr ¼ f ky d1=2 ½2a
m
the method of Venkadesan et al.[15]
where Dr is the strengthening increment due to grain
size hardening and f is the fraction of the alloy occupied
by either the c-c¢ grains or large, primary c¢ separated by
III. STRENGTHENING MECHANISMS IN IN100 high-angle boundaries.
When viewing the microstructures in Figure 2, it is In Ni-base superalloys, only a fraction of the primary
apparent that multiple strengthening modes may be c¢ precipitates are in contact with other primary c¢
active in these materials. Possible operative mechanisms precipitates; therefore, the Hall–Petch contribution
include solid-solution strengthening of the c matrix and attributable to primary c¢ is less than Eq. [2a] predicts.
c¢ precipitates, grain size hardening of both the c matrix To account for this, several hundred primary c¢ precip-
and the primary c¢ precipitates, and shearing of each itates were inspected and the fraction of primary c¢ with
population of precipitates. In addition, depending on adjacent primary c¢ precipitates was measured. It was
the size of the c¢ precipitates, the mechanism limiting the found that 2/3 of primary c¢ had adjacent primary c¢
dislocation motion may be weak pair coupling shearing, precipitates. To account for this, Eq. [2a] was modified to
strong pair coupling shearing, or cross-slip-induced 2
hardening at elevated temperatures. Each of these Dr ¼ f kyc0 d1=2
m ½2b
3
mechanisms will be discussed in more detail in the
following sections. Bowing between precipitates was for primary c¢ precipitates. This is clearly an approxi-
considered but, due the high volume fraction and small mate estimate needed to capture the contribution of the
size of the c¢, bowing was not generally operative. varying primary c¢ size on the overall strength. More
Coherency strengthening was also considered but, in the detailed modeling of deformation in this type of duplex
alloys being studied, the precipitate-matrix misfit was system would be useful, to establish a sound analytical
near zero, so misfit strengthening was neglected. basis for this aspect of strengthening. In the present
In the following sections, analytical expressions that analyses, this contribution to strengthening is small, due
account for the strengthening due to each of these to the low volume fraction of primary c¢.
individual mechanisms are introduced, the exact forms
of the equations used in the model will be presented and
B. Solid-Solution Strengthening of c Matrix
the mechanism(s) responsible for the strengthening due
to each microconstituent will be discussed. The theory for solid-solution hardening of the nickel
matrix is based on a model proposed by Gypen and
Deruyttere,[20] which assumes a superposition of
A. Grain Size Hardening strengthening of individual solutes that individually
The most commonly encountered hardening mecha- have differing potencies.[21–24] Because solute spacing is
nism is that associated with variations in grain size. proportional to the square root of the concentration, the
Strengthening is well known to be inversely proportional resultant strengthening is given by
to the grain size; this phenomenon has been attributed X  dr pffiffiffiffiffi
to the existence of grain boundaries as sources of Dr ¼ pffiffiffiffiffiffiffi Ci ½3
dislocations and barriers to deformation. In IN100, i
dCi
high-angle grain boundaries serving as sources for ffiffiffiffiffi is a strengthening coefficient that reflects the
where pdr
dislocations are present both between c-c¢ grains and dCi
between the large primary c¢ precipitates. strengthening potency of each alloying element.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 40A, JULY 2009—1591


Table II. Solid-Solution-Strengthening Constants
for Alloying Additions to the c Matrix (dr/dC1/2
i )

Al Cr Co Mo Ti V
1/2
Constant (MPa/pct ) 225 337 39.4 1015 775 408

By studying the effect of ternary and higher-order


additions to the c nickel matrix, Roth[25] developed
strengthening coefficients dr/dC1/2 i that describe the
contribution of various elements in solution with nickel.
The constants pertinent to strengthening the c matrix in
IN100 are shown in Table II.

C. Strengthening Due to Dislocation Interactions


The c¢ precipitates in IN100 have the L12 (ordered fcc)
crystal structure. When these ordered precipitates are
sheared, ordinary dislocations must travel in pairs so
that order is not destroyed. When a pair of dislocations
shears a c¢ precipitate, there are several mechanisms that
can dictate the interaction, depending upon the size of
the precipitates and the (antiphase boundary) APB
energy (and, thus, the equilibrium spacing of paired
dislocations). For small precipitates (less than approx-
imately 20 nm), the pair of dislocations may not lie
within an individual precipitate (Figure 3(a)). This case Fig. 3—(a) Configuration of dislocations and precipitates during
is referred to as weak pair coupling. When the precip- weak pair coupling, (b) configuration of dislocations and precipitates
during strong pair coupling, and (c) relative strength due to each
itates are larger, the trailing dislocation enters a precip- mechanism vs precipitate size. Duplicated from Huther et al.[12]
itate before the leading dislocation exits (Figure 3(b)).
This situation is referred to as strong pair coupling. In
both of these situations, the strengthening increment is spacing between particles Ls. For infinitely small spher-
due to the formation of an APB within the precipitate. ical particles, Ardell[36] proposed that
Figure 3(c) schematically compares the stress required sffiffiffiffiffiffiffi
to push a pair of dislocations though a precipitate for 8
each mechanism as a function of precipitate size. The L¼ ds ½5a
3pf
active mechanism changes from weak pair coupling to
strong pair coupling as the c¢ size increases. where f is the volume fraction and ds is the mean pla-
The phenomenon of increasing strength as a function nar diameter. To account for the finite size of the pre-
of temperature is well documented for Ni3Al in its bulk cipitates, the average planar diameter was subtracted
crystalline form.[26,27] While detailed models continue to from L, resulting in the spacing between precipitates:
be developed to describe the associated dislocation sffiffiffiffiffiffiffi
mechanisms, the basis of this hardening mechanism 8
involves thermally activated cross-slip pinning of mobile Ls ¼ ds  ds ½5b
3pf
dislocations.[28–35] As will be discussed in Section 3, this
cross-slip-induced hardening is apparently active in the When dislocations shear ordered, coherent particles,
primary c¢ and large secondary c¢ precipitates (>300 nm). an APB forms on the slip plane. By multiplying the APB
energy cAPB by the length of the precipitate-opposing
1. Weak pair coupling dislocation motion, the mean planar diameter ds, F can
According to Brown and Ham,[10] the stress required be replaced by cAPBds. Substituting this expression into
to move a dislocation though a random array of Eq. [4] gives
obstacles is
3=2
cAPB 3=2 ds
F3=2 s¼ pffiffiffiffiffiffiffiffiffi ½6
s ¼ pffiffiffiffiffiffiffiffiffi ½4 bLs 2TL
bL 2TL
where s is the shear stress, F is the obstacle strength, b is Equation [6] represents the shear stress required to
the Burgers vector, L is the particle spacing, and TL is push the leading dislocation into a precipitate. While the
the dislocation line tension. leading dislocation is creating disorder within the parti-
Equation [4] assumes that precipitates are infinitely cle, the trailing dislocation is restoring order and, as a
small. To account for the finite size of precipitates, the result, the stress required to move the second dislocation
center-to-center particle spacing L was replaced with the through a precipitate is not equal to Eq. [6]. To find the

1592—VOLUME 40A, JULY 2009 METALLURGICAL AND MATERIALS TRANSACTIONS A


stress on the second dislocation, an interaction force sinter
must be accounted for, due to the elastic repulsion
between the dislocations. The following set of equations
describes the force balance on each dislocation:
leading dislocation: sbL1 þ sinter bL1  cAPB d1 ¼ 0
½7
trailing dislocation: sbL2  sinter bL2 þ cAPB d2 ¼ 0
where s is the shear stress required to move the dislo-
cation pair and d is the mean length of dislocation ly-
ing inside a precipitate. The subscripts 1 and 2 refer to
the leading and trailing dislocation, respectively. The
interaction force can be eliminated from Eq. [7] by
solving the two equations simultaneously, giving
 
c d1 d2
s ¼ APB  ½8
2b L1 L2
Fig. 4—TEM bright-field image of subsolvus IN100 after 1 pct
In order to comply with Eq. [6], plastic deformation at 621 C, showing pairs of dislocations between
secondary c¢ precipitates. Inset is the TEM dark-field image of
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi subsolvus IN100 before deformation, showing tertiary c¢ particles
d1 cAPB ds ds between secondary c¢ precipitates.
¼
L1 2TL Ls
precipitate; the maximum stress required to push the
The second dislocation passes through the material dislocations through the precipitate occurs just as the
while remaining essentially straight (Figure 3(a)). second partial dislocation reaches the precipitate. A
Because of this, L2 is equal to L (Eq. [5a]) and d2 is detailed derivation of the strong pair coupling equation
the average length of a random line intersecting a is presented in Reference 12 and will not be repeated
particle of diameter ds, which is pd2s /4L2. Substituting here. The analysis by Huther and Reppich resulted in
the expressions for d1/L1 and d2/L2 into Eq. [8] produces the following expression for the strengthening increment
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi due to shearing in the strong pair mode:
  !    1=2
cAPB cAPB ds ds p ds 2 1 Gb 1=2 pds cAPB
s¼  ½9 Ds ¼ f 0:72w  1 ½10
2b 2TL Ls 4 L 2 ds wGb2
where G is the shear modulus, ds is the mean planar
Equation [9] is the weak pair-coupling equation used diameter, and w is a parameter reflecting the elastic
in the present model; it is essentially the same equation interaction between paired dislocations.
as that presented in Gleiter and Hornbogen[8] and Several modifications to the model of Eq. [10] were
Brown and Ham,[10] except for the use of interparticle incorporated, to capture the strengthening mechanisms
spacing Ls. Equation [9] indicates that weak pair in the more complex alloys of this study, as follows.
coupling is approximately a function of d1/2 s and c3/2
APB.
Weak pair coupling is the active deformation mecha- (a) In the model of Huther and Reppich, the yield
nism for precipitates with an average planar diameter less strength was attributed to a single deformation
than approximately 10 to 20 nm, depending on the APB mechanism. Because of this, the empirical parameter
energy. Because of this, weak pair coupling is generally w was apparently required to account for the effects
only operative when tertiary c¢ precipitates are present. of other strengthening mechanisms, such as solid-
Figure 4 shows a pair of dislocations gliding between solution strengthening. Because the goal of this pres-
two larger secondary c¢ precipitates at 1 pct plastic ent study was to create a model without any fitting
deformation at 621 C. The spacing between leading parameters, w was removed from their equation.
and trailing dislocations is 20 to 30 nm, as indicated by (b) Huther and Reppich assumed that all dislocations
the white arrows. The inserted TEM dark-field image were of a screw character and used a constant line
shows tertiary c¢ particles between secondary particles. tension in their derivation, with TL = Gb2/2 incor-
Because tertiary particles are approximately 10 nm in porated into Eq. [10]. In the present study, it was
diameter, there should be some tertiary particles assumed that dislocations were 60 deg, mixed in
between the leading and trailing dislocations. character. Accounting for the character of the
dislocation:[10]
   
2. Strong pair coupling Gb2 1  mcos2 h R
When ordered precipitates reach a critical size, their TL ¼ ln ½11
4p 1m ro
resistance to shearing begins to decrease. Huther and
Reppich[12] conducted a theoretical analysis of this where m is the Poisson’s ratio; h is the angle between
phenomenon. They found that, when a precipitate is the dislocation line and its Burgers vector; R is the
large enough, both dislocations can be within the same outer cutoff distance, which is approximately the

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 40A, JULY 2009—1593


distance to the closest parallel dislocation of the in a manner similar to bulk Ni3Al. Considering that the c¢
opposite sign; and ro is the inner cutoff radius, taken precipitates are initially free of dislocations, the operation
to be b. According to Brown and Ham,[10] R is of such hardening processes would result in the storage of
approximately equal to the interparticle spacing. dislocations within the Ni3Al. If these processes do not
Due to the high volume fraction of precipitates in operate, it is expected that the dislocation pairs would
the alloys being studied, the interparticle spacing is enter the precipitate, glide uninhibited across the precip-
generally between 5 and 50 nm. As the precipitates itate, and exit. Figure 5 shows dislocations accumulating
become larger, a substructure in the adjacent matrix in a primary c¢ particle after 1 pct strain at 621 C. This
develops (Figure 4). To account for the combined and other similar observations suggest that thermally
effect of precipitate spacing and the presence of activated cross-slip pinning is an additional strengthening
dislocation networks in the matrix, R is taken to be mechanism active in large c¢ precipitates, including
a constant and is assigned a value of 10 nm. primary and large secondary c¢ precipitates.
(c) In the model of Huther and Reppich, the particle It is a challenge to account for cross-slip-induced
spacing L was taken to be (p/f)1/2 ds/2. As previously hardening in secondary c¢ precipitates, which range in
discussed, the present study has adopted Eq. [5a] to size from 100 to 900 nm. To study this in more detail,
describe the center-to-center precipitate spacing. In two unimodal microstructures with mean precipitate
the case of strong pair coupling, interparticle spac- diameters of 180 and 564 nm were established by the
ing is not described by Eq. [5b] because, at the criti- supersolvus heat treatment of unmodified IN100.
cal stress, the length of the dislocation within a Figure 6 shows TEM images of these samples after
precipitate is not equal to ds. Huther and Reppich[12] 1 pct total deformation at 621 C. In the sample with
have defined d(x) to be the length of the dislocation smaller precipitates, pairs of dislocations are observed
within a precipitate at the critical stress, where mainly in the c matrix (Figure 6(a)). On the other hand,
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi in the sample with larger precipitates, the accumula-
4T dkAPB p tion of dislocations within precipitates is observed
dðxÞ ¼ 1 ½12
kAPB p 2T (Figure 6(b)). The fact that the majority of the disloca-
tions observed in Figure 6(a) exist as pairs located in the
matrix indicates the shearing of precipitates without
Substituting d(x) for ds, Eq. [5b] become r dislocation trapping by thermally activated cross-slip
sffiffiffiffiffiffiffi pinning. Considering the high stresses required to push
8 dislocations into smaller precipitates along with the
Lx ¼ ds  dðxÞ ½13 smaller area swept out by the dislocations as they glide
3pf
through the particle, the opportunity for a thermally
activated cross-slip pinning event would diminish as the
Incorporating these changes, the form of the strong precipitate size decreases.
pair coupling equation used in the model is Quantitative analyses of the incidence of dislocation
 1=2 accumulation in the precipitates as a function of
2TL pds cAPB precipitate size were conducted on a large number of
Ds ¼ 1 ½14 micrographs. A similar volume of precipitate material
pbLx 2TL
was examined at each precipitate size. Figures 7(a) and
(c) show the distribution of the precipitate size, mea-
Evaluation of Eq. [14] established that strong pair sured as a mean value of the precipitate widths along
coupling is approximately a function of d1/2
s and c1/2
APB, two h100i directions. The size ranges measured were 70
for the alloys of interest. to 300 nm and 250 to 900 nm, respectively. The number
Strong pair coupling is the active deformation mech- of dislocation segments per particle was counted and
anism for c¢ precipitates with an average planar diameter plotted against precipitate size, as shown in Figures 7(b)
greater than approximately 10 to 20 nm, again depend- and (d). In the sample with small precipitates, the
ing again on the APB energy. This means that strong value is less than 1, regardless of the precipitate size
pair coupling is generally the active shearing mechanism (Figure 7(b)). However, in the sample with large pre-
for secondary c¢ precipitates. Figure 4 shows pairs of cipitates, the value is approximately 2 pairs of disloca-
a=2h110i dislocations in the matrix region between tions in precipitates smaller than 550 nm, and more than
secondary c¢ precipitates, labeled ‘‘A,’’ indicating that twice that when the size exceeds 700 nm (Figure 7(d)).
dislocations must move in pairs to facilitate the shearing This result indicates that thermally activated cross-slip
of ordered secondary c¢ precipitates. Finally, it is worth pinning is activated much more frequently in precipi-
noting that the effects of plastic constraint[37] were not tates larger than 300 nm and that this is likely to
considered, due to the fact that the precipitates are contribute to the overall strength of the material.
spherical and may eventually be sheared. In addition to investigating the incidence of disloca-
tion accumulation in the samples with a unimodal
3. Cross-slip-induced hardening microstructure with a mean precipitate size of 564 nm,
Similar to single-phase Ni3Al, many two-phase six samples were tensile tested at temperatures ranging
superalloys exhibit a peak in flow stress in the temperature from room temperature to 760 C, to determine the
range 500 C to 750 C.[38–43] This suggests that cross-slip temperature dependence of the flow stress at different
pinning processes operate within the larger c¢ precipitates levels of offset strain. The flow stress at different levels of

1594—VOLUME 40A, JULY 2009 METALLURGICAL AND MATERIALS TRANSACTIONS A


Fig. 5—TEM bright-field image of subsolvus IN100 after 1 pct plastic deformation at 621 C, showing dislocations accumulating in primary c¢.

microstrain was found to increase to a greater degree at approach of Mishima, the cross-slip-induced hardening
temperatures above 400 C, compared to lower temper- constants for individual alloying elements in the super-
atures; this indicates that hardening due to thermally alloys were assessed and are summarized in Table III.
activated cross-slip pinning processes is likely to be The cross-slip-induced hardening equation in the
operative (Figure 8).[37] model is " #
As mentioned previously, there have been a number X  dr 
of studies in Ni3Al that attempt to relate detailed Dr ¼ f rðTÞNi3 Al þ Ci ½15
dislocation-hardening phenomena to macroscopic flow i
dCi
behavior.[28–35] However, there is not yet a comprehen-
sive theory for hardening in these materials. Because where f is the fraction of precipitates subject to cross-
there is an abundance of experimental data pertaining to slip-induced hardening, rðTÞNi3 Al is the strength of pure
cross-slip-induced hardening as a function of tempera- Ni3Al as a function of temperature, and Ci is the
ture, the strength derived from this mechanism was concentration of the ith alloying element. In the model,
quantified with the use of this large experimental data cross-slip-induced hardening occurs in primary c¢ and
set reported in the literature. The relationship between secondary c¢ precipitates larger than 300 nm. An APB
the cross-slip-induced hardening of pure Ni3Al and energy of 200 mJ/m2, an intermediate value among
temperature was adapted from Mishima,[44] Flinn,[45] those reported in the literature,[52–55] was used in the
and Davies[46] (Figure 9). These three authors experi- majority of the analyses. As will be discussed later, this
mentally determined the strength of pure polycrystalline material property, while not well known, has a signif-
Ni3Al at temperatures ranging from 73 to 1273 K. The icant effect on strength.
results of these studies were averaged and were used as
the strength of pure Ni3Al (Figure 9). 4. Bowing
Mishima[44,47] and others[43,48–50] also found experi- Orowan bowing occurs when the stress required for a
mentally that the change in strength due to an alloying dislocation to bow between precipitates is less than the
addition to Ni3Al is linearly proportional to the atomic stress required for the dislocation to penetrate precip-
fraction of that alloying element. Figure 10 shows the itates. The stress above which bowing will occur was
data of Mishima reproduced in Reference 51, demon- determined by Orowan and is[10]
strating how additions of 4 and 8 pct Ti raise the flow Gb
stress in proportion to the titanium content. From a s¼ ½16
L
series of such experiments, Mishima determined con-
stants for the strength increment provided by a wide Equation [16] considers infinitely small precipitates. To
range of the alloying additions to Ni3Al. Using the account for the finite size of precipitates, the center-to-

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 40A, JULY 2009—1595


Fig. 8—Temperature dependence of the flow stress for unimodal
IN100 with 564-nm c¢ precipitates measured at varying levels of mi-
crostrain. The larger increase in flow stress at high temperatures
after larger amounts of microstrain indicate cross-slip pinning is
occurring.

Fig. 6—TEM bright-field images of IN100 with a unimodal micro-


structure after 1 pct total deformation at 621 C. Mean precipitate
size is (a) 180 and (b) 564 nm.

Fig. 9—Experimental data showing the relationship between cross-


slip-induced hardening of binary Ni3Al and test temperature. Data
of Mishima,[44] Flinn,[45] Davies,[46] and the composite curve are used
in the model.

Fig. 7—Histograms of (a) and (c) c¢ precipitate size distribution and


(b) and (d) number of dislocation segments per precipitate in IN100
with a unimodal microstructure after 1 pct total deformation at
621 C. Mean precipitate size is (a) and (b) 180 and (c) and (d) Fig. 10—Experimental data of Mishima[51] demonstrating the linear
564 nm. dependence of the strength of c¢ on the alloying concentration.

1596—VOLUME 40A, JULY 2009 METALLURGICAL AND MATERIALS TRANSACTIONS A


Table III. Cross-Slip-Induced Hardening Constants
for Alloying Additions to c¢ (dr/dCi)

Al Cr Co Mo Ti V
Constant (MPa/pct) 0 7 0 16.8 25 8.2

center spacing was replaced with the interparticle spac-


ing:
Gb
s¼ ½17
Ls
The stress required for dislocations to bow between
precipitates is inversely proportional to the interparticle
spacing and consequently, bowing typically occurs Fig. 11—Comparison of the yield strength attributable to a popula-
between widely spaced precipitates. tion of tertiary c¢ with a volume fraction of 0.06 as a function of the
average diameter. Without using the LSW size distribution, the pair-
coupling equations would overestimate the yield strength by up to
12 MPa. The LSW distribution shown is for a population with an
IV. MODELING CONSIDERATIONS average diameter of 13.25 nm.

To assemble a comprehensive model, several other


factors had to be taken into account. The weak and dependence on temperature. However, it might be
strong pair-coupling equations overlook the fact that expected that the penetration of dislocations into the
precipitates have a distribution of sizes; this may result precipitates is thermally assisted. The kinetics of ther-
in significant error if the precipitates are in the size range mally activated deformation have been considered in
near a mechanism transition. These equations also detail by Kocks, Argon, and Ashby[61] and the relevant
disregard the characteristic temperature dependence of functional form for temperature dependent changes in
the strengthening phenomena, decreasing their accuracy flow stress is
at the temperatures most critical to industrial applica- 
tions. Additionally, the superimposition of the strength- @lnr  Q
¼ ½18
ening due to each mechanism must be considered. @lnTe_ mkT
where Q is the activation energy, m is the strain-rate
A. Precipitate Size Distributions sensitivity, k is the Boltzmann’s constant, e_ is the strain
The nucleation, growth, and coarsening of precipi- rate, and T is the temperature.
tates typically results in a characteristic precipitate Due to the complexity of the deformation modes, the
distribution.[56–59] Parametric studies have revealed that activation energy and strain-rate sensitivity are not
an error in the strengthening increment associated with easily derived from theoretical considerations only.
neglecting the size distribution occurs when the average Therefore, strain-rate change experiments were used to
precipitate size is close to the transition between weak determine the values of these parameters for subsolvus
pair coupling and strong pair coupling. Therefore, a IN100. According to Reference 61:

Lifshitz, Slyozov, and Wagner (LSW) size distribu- @ln_e 
tion[60] was applied to each of the individual populations m¼ ½19
@lnrT
of c¢ precipitates such that the predicted strengthening
increment is the weighted average of the strength due to
all likely precipitate sizes within the distribution. Using Eq. [19], m was determined by measuring the
In Figure 11, the yield strength increment due to change in flow stress associated with a change in the
strong and weak pair coupling as a function of the mean strain rate at temperatures ranging from room temper-
planar diameter is plotted for a c¢ population with a ature to 1033 K, as described in Section II–D. The
volume fraction of 0.06. For this population, the results of these tests are summarized in Table IV. After
transition between weak and strong pair coupling is at the strain-rate sensitivity was determined at these five
13.25 nm. The bold curve in Figure 11 shows the yield test temperatures, an optimization routine using iSight
strength increment due to this population, when an software (SIMULIA, Providence, RI) was used to
LSW size distribution is taken into account. In this case, establish a functional form for strain-rate sensitivity as
the omission of a size distribution results in an overes- a function of temperature.
timation of the yield strength of approximately 12 MPa. The activation energy was taken to be a constant, but
a discontinuity was found at approximately 400 C.
At this temperature, Q was found to change from
B. Temperature Dependence of Pair Coupling
156 kJ/mol.K below 400 C to 380 kJ/mol.K above
The equations prescribed by Brown and Ham and 400 C. It is interesting that 400 C corresponds to the
Huther and Reppich for pair coupling did not include a minimum in strength for IN100, and is consistent with

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 40A, JULY 2009—1597


Table IV. Experimentally Determined Strain-Rate Sensitivity Table V. Values of k from the Literature
and Yield Strength vs Temperature
References Material k
Temperature (K) m Yield Strength (MPa)
61 theoretical calculations 1
295 360.5 1094 66 SS + SiO2 strengthened Cu 1
533 — 1056 67 SS + Al2O3 strengthened Cu-Zn 1.5
894 146.7 1051 62 theoretical calculations 1.5
922 92.3 1057 62 SS + SiO2 strengthened Cu 1.76
977 51.2 1025 18 Ni superalloys with ordered 1.13 to 1.19
1017 22.4 986 coherent c¢

Fig. 13—Experimental and predicted yield strength curves for sub-


Fig. 12—Experimentally determined temperature dependence on pair solvus IN100, assuming different values of k.
coupling.

the onset of thermally activated cross-slip pinning


processes. In Figure 12, the remaining fraction of the with volume fractions ranging from 0.07 to 0.47. They
room-temperature yield strength predicted by the pair- found that k was 1.13 to 1.19 for underaged and peak-
coupling equations is plotted as a function of temper- aged c¢ and slightly higher for overaged c¢.
ature. This temperature dependence was incorporated In the model, taking k to be significantly different
into the yield strength model by multiplying Eqs. [9] and from 1 resulted in considerable underestimation of the
[14] by the function plotted in Figure 12. In this sense, yield strength. Additionally, varying k essentially acted
the temperature dependence is considered to arise from as a fitting parameter that could be used to alter the
the thermally assisted penetration of dislocations into magnitude of the yield strength vs the temperature
the precipitates. profile. These findings are illustrated in Figure 13, in
which experimental yield strength curves are plotted for
C. Superposition subsolvus IN100 along with predicted yield strength
curves, assuming different values of k. In the absence of
Once the increment of strength due to each mecha- any strong evidence for strong superposition effects, k
nism was assessed, it was necessary to consider how was assigned a value of unity. Taking k to be 1, the
these strengthening increments should be superimposed. contributions to yield strength in the model are then
There are several possible methods of superposition, defined as
most of which assume the general form:
X rys ¼ grain size hardening of c matrix (Eq:½2Þ
rktotal ¼ Drki ½20
þ solid-solution hardening of c matrix (Eq:½3Þ
i
þ grain size hardening of primary c0 (Eq: [2b])
where k can be taken as 1 for linear superposition or 2
for Pythagorean superposition, or any value in between. þ anomalous hardening of primary c0 and
There are many examples from the literature in which large secondary c0 (Eq:½15Þ
k has been assigned various values between 1 and 2. þ (shearing/bowing of secondary c0 -strong pair
Some values of k from the literature were summarized in
Reference 62 and have been included in Table V. The coupling (Eqs: [14] and [17])
study of Nembach et al.[18] was conducted on nickel-base þ shearing/bowing of tertiary c0 -weak pair
superalloys similar to Nimonic 105 that contained a
coupling (Eqs: [9] and [17]) fðTÞ ½21
unimodal distribution of ordered coherent c¢ precipitates

1598—VOLUME 40A, JULY 2009 METALLURGICAL AND MATERIALS TRANSACTIONS A


where Eqs. [9] and [12] are modified to incorporate an greater than 300 nm, the shape of the yield-strength-vs-
LSW size distribution and f(T) indicates the temperature temperature profile does not agree with the experimental
dependence of dislocation shearing. Note that Eq. [21] data.
contains no fitting parameters, although some physical Figure 16 contains a summary showing the error
constants contained in the individual equations have between the predicted and experimental yield strength
uncertainty. for all test samples. This graph illustrates good corre-
A program was written to evaluate Eq. [21] based lation between the predicted and experimental yield
upon input from a user or upon direct input from strength for subsolvus IN100. For the supersolvus
software used to simulate microstructures. The program IN100, modified subsolvus IN100, and modified chem-
accepts microstructural input in terms of both the size istry alloy, the correlation between the predicted and
and volume fraction of each microconstituent and the
chemical composition of the c and c¢ phases. This
program was used to evaluate the model in comparison
to a range of experimental yield strength data.

V. RESULTS AND DISCUSSION


A. Predicting Yield Strength
A total of approximately 36 tensile tests were con-
ducted on a variety of microstructures and chemistries
over the temperature range 22 C to 760 C. Supersol-
vus IN100, subsolvus IN100, and two different heat
treatments of the chemically modified supersolvus alloy
(Table I) were analyzed and evaluated against the
model. The heat treatments and tensile testing were
performed at Pratt & Whitney (East Hartford, CT) and
Fig. 14—Comparison of predicted and experimental yield strengths
the University of Michigan (Ann Arbor, MI). The vs test temperature for subsolvus IN100 and supersolvus IN100.
microstructural information summarized in Table VI
was experimentally characterized by Wusatowska-Sar-
nek et al.[14]
Figure 14 shows a comparison of the experimental
and predicted yield strength vs temperature for the
unmodified subsolvus IN100 and supersolvus IN100.
Figure 14 demonstrates good agreement between pre-
dicted and experimental results for the unmodified
subsolvus IN100. It also displays a good prediction of
the yield-strength-vs-temperature profile for supersolvus
IN100. It is worth noting the lack of primary c¢ in the
supersolvus alloy due to the heat treatment and the
corresponding higher volume fractions of secondary and
tertiary c¢. The model overpredicts the supersolvus yield
strength; the likely reasons for this high degree of error
(~200 MPa) will be discussed shortly.
Figure 15 shows the effect of the cross-slip-induced
hardening of large secondary c¢ particles on the yield Fig. 15—Disregarding the cross-slip-induced hardening of large sec-
strength of supersolvus IN100. Aside from the effect of ondary c¢ precipitates leads to underestimation of yield strength,
cross-slip-induced hardening in secondary c¢ precipitates especially at high temperatures.

Table VI. Microstructural Characterization of Alloys Used to Evaluate the Yield Strength Model

Modified Alloy Modified Alloy


Subsolvus Modified Supersolvus Supersolvus Heat Supersolvus Heat
IN100 Subsolvus IN100 IN100 Treatment 1 Treatment 2

Phase Vf Size (lm) Vf Size (lm) Vf Size (lm) Vf Size (lm) Vf Size (lm)
c matrix 0.40 4.1 0.40 4.90 0.40 34.4 0.47 26.79 0.47 32
Primary c¢ 0.20 1.20 0.25 1.27 — — — — — —
Secondary c¢ 0.34 0.17 0.32 0.15 0.46 0.34 0.42 0.11 0.39 0.22
Tertiary c¢ 0.06 0.008 0.03 0.021 0.14 0.011 0.11 0.007 0.14 0.011

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 40A, JULY 2009—1599


Fig. 16—Comparison of predicted to experimental yield strength for
nine test microstructures at all test temperatures.

experimental yield strength is not as good, with devia-


tions between the predicted and experimental yield
strengths ranging from 50 to 200 MPa. Possible causes
for the error between the experimental and predicted
results include the following.
(a) Material characterization. Slight errors in materials
characterization, particularly characterization of
the volume fraction and average diameter of ter-
tiary c¢ precipitates, can lead to large errors in the
predicted yield strength. The magnitude of error
likely due to characterization will be discussed in
Section IV–D. Fig. 17—Even among similar alloys, the source of strength of a
(b) Model assumptions. The theories pertaining to material is highly dependent upon microstructure. These figures
some strengthening mechanisms are not particu- depict the magnitude of strengthening sorted by (a) mechanism and
larly well developed. While an effort was made to (b) microconstituent. The variation in the sources of the strength of
construct the model using the most logical disloca- an alloy illustrates the impossibility of modeling the yield strength of
these materials using a simple model.
tion-precipitate models, in some cases, the models
may not accurately model the actual dislocation
glide characteristics in these complex microstruc-
Similarly, Figure 17(b) shows that each alloy owes its
tures.
strength to different microconstituents. The c matrix
With regard to the strengthening due to strong and gains its strength from solid-solution strengthening and
weak pair coupling, the magnitude of strengthening grain size hardening. The primary c¢ derives its strength
predicted by the model is similar to that predicted by from cross-slip-induced hardening, order strengthening,
some recent two-dimensional dislocation glide mod- and grain size hardening. The secondary c¢ obtains its
els[63,64] but with somewhat different dependencies on strength from strong pair coupling and cross-slip-
APB energy and precipitate size. induced hardening; the tertiary c¢ owes its strength to
weak and strong pair coupling. The input parameters
assumed for all four test alloys summarized in Figure 17
B. Identification of Dominant Mechanisms were APB energy = 0.20 J/m2, c ky = 750 MPa lm1/2,
In addition to predicting the overall yield strength, it and c¢ ky = 500 MPa lm1/2.
is also useful to be able to evaluate the relative This variation in active strengthening mechanisms
contributions of individual mechanisms and, in turn, prevents the construction of simple models. Conversely,
determine which constituents of the microstructure are mapping the mechanisms (microconstituents) responsi-
responsible for the strength of an alloy. Figure 17(a) ble for strength in a manner outlined here could prove
shows the magnitude of strengthening due to each useful in modifying various features of an alloy. It
mechanism for four test alloys. While these alloys have should be highlighted that the modification of these
comparable yield strengths, they each derive their microstructural features and the experimental measure-
strength from different mechanisms. Recall that the ment of their effects on strength can be a time-
strength due to cross-slip-induced hardening may come consuming process.
from both primary and secondary c¢ precipitates and The model can also be used to highlight the elements
that the strength due to strong pair coupling may come crucial to an accurate determination of yield strength.
from both secondary and tertiary c¢ precipitates. Two model inputs found to be especially important to

1600—VOLUME 40A, JULY 2009 METALLURGICAL AND MATERIALS TRANSACTIONS A


yield strength prediction were the APB energy and the input; this variability is summarized in Table VII for
volume fraction of tertiary c¢. Figure 18 shows the subsolvus IN100. The yield strength was then calculated
predicted yield-strength-vs-temperature profile for sub- thousands of times, while each parameter was allowed
solvus IN100 from Figure 13 in bold. For this curve, the to vary independently but randomly within these limits.
values of the APB energy and volume fraction tertiary c¢ The yield strength was found to vary by 247.6 MPa with
are 0.20 J/m2 and 0.06, respectively. Also plotted are the a standard deviation of 39.3, based upon a single set of
curves that indicate how the yield-strength-vs-tempera- input microstructural and chemical data. Figure 19
ture profile would change if either the APB energy or the shows the variation in the yield strength increment due
volume fraction of tertiary c¢ is altered. to tertiary c¢, as a function of the volume fraction of
Decreasing the APB energy from 0.20 to 0.15 J/m2 tertiary c¢. The volume fraction of tertiary c¢ was
resulted in a reduction in the predicted yield strength of allowed to vary from 0.04 to 0.08 and the average
85 to 100 MPa, depending on the temperature. Increas- diameter of tertiary c¢ was allowed to vary from 6.4 to
ing the volume fraction of tertiary c¢ from 0.06 to 0.11 9.6 nm. The characterization of tertiary c¢ to even this
produced a boost in the predicted yield strength ranging level of accuracy is a very difficult process.[65] This plot
from 115 to 130 MPa, depending on the temperature. indicates that the yield strength increment due to
These are significant deviations in the predicted yield tertiary c¢ ranges from 105 to 302 MPa. Even for a
strength, considering that these parameters are difficult constant volume fraction of tertiary c¢, the yield strength
to determine either experimentally or theoretically.[14,65] increment due to tertiary c¢ has a range of approxi-
From this study, it is clear that detailed measurements mately 100 MPa.
of microstructural parameters are crucial to model
development and that the continued development of
D. Testing the Predictive Capability of the Model
rapid characterization techniques is essential.
Investigation of the strengthening mechanisms active
in multimodal nickel-base superalloys has indicated that
C. Variation in Predicted Yield Strength the presence of tertiary c¢ precipitates have a significant
due to Characterization Error impact on yield strength. To test this theory, supersolvus
Regardless of the methods used in materials charac- IN100 with 180-nm secondary c¢ precipitates with and
terization, there is bound to be some degree of error in without tertiary c¢ was tensile tested to 1 pct total strain
the microstructural quantification and chemical compo- at 621 C. The yield strengths with and without tertiary
sition of an alloy. Typical experimental errors were c¢ were 1092 and 884 MPa, respectively. Figure 20
assigned to each microstructural and chemical model shows the presence of tertiary c¢ precipitates between

Fig. 18—Slight changes in the APB energy or volume fraction of ter- Fig. 19—Variation in the predicted yield strength increment due to
tiary c¢ leads to a large error in yield strength prediction. Both of tertiary c¢, owing to errors in microstructural quantifications as a
these elements of the model are difficult to determine with a high de- function of the measured volume fraction of tertiary c¢. Input param-
gree of accuracy. eters were allowed to vary within the limits outlined in Table VII.

Table VII. Expected Errors in Microstructural and Chemical Composition Input*

Property c Matrix Primary c¢ Secondary c¢ Tertiary c¢


Volume fraction 0.4 ± 2.5 pct 0.2 ± 30 pct 34 ± 10 pct 0.06 ± 33 pct
Size (nm) 4100 ± 25 pct 1200 ± 5 pct 170 ± 10 pct 8 ± 20 pct
APB energy (J/m2) not applicable 0.2 ± 10 pct 0.2 ± 10 pct 0.2 ± 10 pct
*All alloying additions to nickel allowed ±5 pct.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 40A, JULY 2009—1601


ACKNOWLEDGMENTS
The authors express their gratitude to Mike Long,
Engineous Software (SIMULIA, Province, RI), for his
assistance with iSight software, and to Dennis Dimi-
duk and Triplicane Parthasarathy for their valuable
discussions. The financial support of the DARPA/
AIM project under Grant No. F005484, sponsored by
Pratt & Whitney, is gratefully acknowledged.

REFERENCES
1. C. Ducrocq, A. Lasalmonie, and Y. Honnrat: in Superalloys, D.N.
Duhl, G. Maurer, S. Antolovich, C. Lund, and S. Reichman, eds.,
TMS, Warrendale, PA, 1988, pp. 63–72.
2. R.H. Caless and D.F. Paulonis: in Superalloys, D.N. Duhl, G.
Maurer, S. Antolovich, C. Lund, and S. Reichman, eds., TMS,
Warrendale, PA, 1988, pp. 101–10.
3. H. Hattori, M. Takekawa, D. Furrer, and R.J. Noel: in Superal-
loys, R.D. Kissinger, D.J. Deye, D.L. Anton, A.D. Cetel, M.V.
Nathal, T.M. Pollock, and D.A. Woodford, eds., TMS, Warren-
Fig. 20—Microstructure of the sample shown in Figure 6(a) follow- dale, PA, 1996, pp. 705–12.
ing an aging heat treatment at 732 C for 8 hours, during which ter- 4. T.E. Howson, Jr., W.H. Couts, and J.E. Coyne: in Superalloys,
tiary c¢ precipitated. M. Gell, C.S. Kortovich, R.H. Bricknell, W.B. Kent, and J.F.
Radavich, eds., TMS, Warrendale, PA, 1984, pp. 277–86.
5. F.R.N. Nabarro and H.L. deVilliers: The Physics of Creep: Creep
and Creep-Resistant Alloys, Taylor & Francis, London, 1995.
the secondary c¢ following the aging heat treatment. The 6. J. Dennison, P.D. Holmes, and B. Wilshire: Mater. Sci. Eng.,
volume fraction of tertiary c¢ was estimated from TEM 1978, vol. 33, pp. 35–47.
micrographs; the difference in the yield strength before 7. A.J. Foreman and M.J. Makin: Philos. Mag. A, 1966, vol. 14,
pp. 911–24.
and after the precipitation of tertiary c¢ was predicted to 8. H. Gleiter and E. Hornbogen: Mater. Sci. Eng., 1968, vol. 2,
be 170 ± 50 MPa, in good agreement with the experi- pp. 285–302.
mental results. This verifies the fact that tertiary c¢ 9. P. Feltham: J. Phys. D, 1968, vol. 1, pp. 303–08.
precipitates have a significant influence on the yield 10. L.M. Brown and R.K. Ham: in Strengthening Methods in Crystals,
A. Kelly and R.B. Nicholson, eds., Elsevier Publishing Co., Ltd.,
strength of multimodal nickel-base superalloys. Essex, England, 1971, pp. 9–135.
11. J.L. Castagne: J. Phys., 1966, vol. 27, pp. 233–39.
12. W. Huther and B. Reppich: Z. Metallkd., 1978, vol. 69, pp. 628–34.
13. E.J. Lee and A.J. Ardell: in Strength of Metals and Alloys, P.
VI. CONCLUSIONS Haasen, V. Gerold, and G. Kostorz, eds., Pergamon Press, Ltd.,
Oxford, England, 1979, pp. 633–38.
In this study, the contributions to the strength of 14. A.M. Wusatowska-Sarnek, G. Ghosh, G.B. Olson, M.J. Blackburn,
polycrystalline multimodal nickel-base superalloys have and M. Aindow: J. Mater. Res., 2003, vol. 18, pp. 2653–63.
been investigated and captured in a multimechanism- 15. S. Venkadesan, P. Rodriguez, K.A. Padmanabhan, P.V. Sivapra-
yielding model. The following conclusions can be made. sad, and C. Phaniraj: Mater. Sci. Eng., 1992, vol. A154, pp. 69–74.
16. M. Meyers and K. Chawla: Mechanical Behavior of Materials,
1. Modeling that consists of the superposition of Prentice Hall, Upper Saddle River, NJ, 1999.
possible strengthening mechanisms in a polycrystal- 17. A.W. Thompson: Acta Metall., 1977, vol. 25, pp. 63–66.
18. S. Schilnzer and E. Nembach: Acta Metall., 1992, vol. 40, pp. 803–13.
line nickel superalloy with multiple distributions of 19. F. Wallow and E. Nembach: Scripta Metall., 1996, vol. 34,
precipitates is useful for optimization of the micro- pp. 499–505.
structure and heat-treatment approaches. 20. L.A. Gypen and A. Deruyttere: J. Mater. Sci., 1977, vol. 12,
2. While APB energies and volume fractions of pp. 1028–33.
tertiary c¢ < 20 nm are difficult to measure experi- 21. R.L. Fleischer: Acta Metall., 1963, vol. 11, pp. 203–09.
22. N.F. Mott and F.R.N. Nabarro: in 1947 Bristol Conference on
mentally, they have a strong influence on the pre- Strength of Solids, Physical Society of London, 1948, p. 1.
dicted yield strength. 23. J. Friedel: Dislocations, Pergamon Press, Oxford, England, 1964.
3. Dislocation storage in precipitates with radii less 24. U.F. Kocks: Metall. Trans. A, 1985, vol. 16A, pp. 2109–30.
than 300 nm is infrequent; therefore, cross-slip- 25. H.A. Roth, C.L. Davis, and R.C. Thomson: Metall. Mater. Trans.
A, 1997, vol. 28A, pp. 1329–35.
induced hardening contributes significantly to the 26. J.H. Westbrook: Trans. AIME, 1957, vol. 209, pp. 898–904.
yield strength when secondary c¢ precipitates 27. P.A. Flinn: Trans. AIME, 1960, vol. 218, pp. 145–54.
become larger than approximately 300 nm in diam- 28. B.H. Kear and H.G.F. Wilsdorf: Trans. TMS-AIME, 1962,
eter. vol. 224, pp. 362–64.
4. Modeling indicates that materials subjected to heat 29. O. Veyssiere and G. Saada: in Dislocations in Solids, F.R.N.
Nabarro and M.S. Duesbery, eds., North Holland, Amsterdam,
treatments below the c¢ solvus temperature benefit 1996, pp. 253–441.
from a strong Hall–Petch contribution, while super- 30. D. Caillard and J.L. Martin: Thermally Activated Mechanisms in
solvus heat-treated materials gain the majority of Crystal Plasticity, Pergamon, London, 2003.
their strength from their resistance to precipitate 31. K.J. Hemker, M.J. Mills, and W.D. Nix: J. Mater. Res., 1992,
vol. 7, pp. 2059–69.
shearing.

1602—VOLUME 40A, JULY 2009 METALLURGICAL AND MATERIALS TRANSACTIONS A


32. B. Viguier, J.L. Martin, and J. Bonneville, in Dislocations in Solids, 49. J. Lopez and G.F. Hancock: Phys. Status Solidi A, 1970, vol. 2,
F.R.N. Nabarro and M.S. Duesbery, eds., North Holland, pp. 469–74.
Amsterdam, 2002, p. 459. 50. K. Aoki and O. Izumi: Phys. Status Solidi A, 1976, vol. 38,
33. S.S. Ezz and P.B. Hirsch: Philos. Mag. A, 1994, vol. 69, pp. 105–27. pp. 587–94.
34. B. Devincre, P. Veyssiere, and G. Saada: Philos. Mag. A, 1999, 51. C.T. Liu and D.P. Pope: in Intermetallic Compounds, J.H.
vol. 79, pp. 1609–28. Westbrook and R.L. Fleischer, eds., John Wiley & Sons,
35. V. Paidar, D.P. Pope, and V. Vitek: Acta. Metall., 1984, vol. 32, New York, NY, 1995.
pp. 435–48. 52. R.J. Taunt and B. Ralph: Philos. Mag., 1974, vol. 30, pp. 1379–94.
36. A.J. Ardell: Metall. Trans. A, 1985, vol. 16A, pp. 2131–66. 53. M. Vittori and A. Mignone: Mater. Sci. Eng., 1985, vol. 71,
37. P.H. Thornton, R.G. Davies, and T.L. Johnston: Metall. Trans., pp. 29–37.
1970, vol. 1, pp. 207–18. 54. D.M. Dimiduk, A.W. Thompson, and J.C. Williams: Philos. Mag.
38. B.H. Kear and B.J. Piearcey: Trans. TMS-AIME, 1967, vol. 239, A, 1993, vol. 67 (3), pp. 675–98.
pp. 1209–15. 55. T. Kruml, B. Viguier, J. Bonneville, P. Spatig, and J.L. Martin:
39. V.K. Sikka and E.A. Loria: in Superalloys, D.N. Duhl, G. Maurer, MRS Symp. Proc., 1997, vol. 460, pp. 529–34.
S. Antolovich, C. Lund, and S. Reichman, eds., TMS, Warrendale, 56. D. McLean: Met. Sci., 1984, vol. 18, pp. 249–56.
PA, 1988, pp. 203–12. 57. P.W. Voorhees: J. Stat. Phys., 1985, vol. 38, pp. 231–53.
40. C. Monier, C. Bertrand, J.-P. Dallas, M.-F. Trichet, and M. 58. S.C. Hardy and P.W. Voorhees: Metall. Trans. A, 1988, vol. 19A,
Cornet: Mater. Sci. Eng., 1994, vol. A188, pp. 133–39. pp. 2713–21.
41. A. Nitz and E. Nembach: Metall. Mater. Trans. A, 1997, vol. 29A, 59. T. Eguchi, Y. Tomokiyo, and S. Matsumura: Phase Transitions,
pp. 799–807. 1987, vol. 8, pp. 213–26.
42. A. Banik and K.A. Green: in Superalloys, T.M. Pollock, R.D. 60. I.M. Lifshitz and V.V. Slyozov: J. Phys. Chem. Solids., 1961,
Kissinger, and R.R. Bowman, eds., 2000, pp. 69–74. vol. 19, pp. 35–50.
43. M. Dollar and I.M. Bernstein: in Superalloys, D.N. Duhl, G. 61. U.F. Kocks, A.S. Argon, and M.F. Ashby: Thermodynamics and
Maurer, S. Antolovich, C. Lund, and S. Reichman, eds., TMS, Kinetics of Slip, Pergamon Press, New York, NY, 1975.
Warrendale, PA, 1988, pp. 275–84. 62. U. Lagerpusch, V. Mohles, and E. Nembach: Mater. Sci. Eng.,
44. Y. Mishima, S. Ochiai, N. Hamao, M. Yodogawa, and T. Suzuki: 2001, vols. A319–A321, pp. 176–78.
Trans. Jpn. Inst. Met., 1986, vol. 27, pp. 648–55. 63. T.A. Parthasarathy, S.I. Rao, and D.M. Dimiduk: Superalloys,
45. P.A. Flinn: Strengthening Mechanisms in Solids, ASM, Metals TMS, Warrendale, PA, 2004, pp. 887–96.
Park, OH, 1962, p. 17. 64. S.I. Rao, T.A. Parthasarathy, D.M. Dimiduk, and P.M. Hazzle-
46. R.G. Davies and N.S. Stoloff: Trans. TMS-AIME, 1965, vol. 233, dine: Philos. Mag. A, 2006, vol. 86 (4), pp. 215–25.
pp. 714–19. 65. P. Sarosi, G. Viswanathan, D. Whitis, and M. Mills: Superalloys,
47. S. Miura, Y. Mishima, and T. Suzuki: J. Jpn. Inst. Met., 1992, TMS, Warrendale, PA, 2004, pp. 989–96.
vol. 56, pp. 1214–20. 66. R. Ebeling and M.F. Ashby: Philos. Mag., 1966, vol. 13, pp. 805–
48. F.E. Haredia and D.P. Pope: High-Temperature Ordered Inter- 34.
metallic Alloys II, Materials Research Society, Pittsburgh, PA, 67. P.B. Hirsch and F.J. Humphreys: Proc. R. Soc., 1970, vol. A318,
1987, pp. 213–20. pp. 45–72.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 40A, JULY 2009—1603

You might also like