Thermomechanical Fatigue Behavior of A Directionally Solidified Ni-Base Superalloy

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

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

net/publication/249505564

Thermomechanical Fatigue Behavior of a Directionally Solidified Ni-Base


Superalloy

Article  in  Journal of Engineering Materials and Technology · July 2005


DOI: 10.1115/1.1924560

CITATIONS READS
61 431

4 authors, including:

Ali P. Gordon Richard W. Neu


University of Central Florida Georgia Institute of Technology
98 PUBLICATIONS   465 CITATIONS    170 PUBLICATIONS   1,899 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Creep and Creep-Fatigue Crack Growth Behavior in Creep-Brittle Materials View project

Gasket Relaxation View project

All content following this page was uploaded by Ali P. Gordon on 21 November 2016.

The user has requested enhancement of the downloaded file.


Thermomechanical Fatigue
Behavior of a Directionally
M. M. Shenoy
A. P. Gordon
Solidified Ni-Base Superalloy
D. L. McDowell A continuum crystal plasticity model is used to simulate the material behavior of a
directionally solidified Ni-base superalloy, DS GTD-111, in the longitudinal and trans-
R. W. Neu1 verse orientations. Isothermal uniaxial fatigue tests with hold times and creep tests are
e-mail: rick.neu@me.gatech.edu conducted at temperatures ranging from room temperature (RT) to 1038° C to charac-
terize the deformation response. The constitutive model is implemented as a User MATe-
The George W. Woodruff School rial subroutine (UMAT) in ABAQUS (2003, Hibbitt, Karlsson, and Sorensen, Inc., Provi-
of Mechanical Engineering, dence, RI, v6.3) and a parameter estimation scheme is developed to obtain the material
Georgia Institute of Technology, constants. Both in-phase and out-of-phase thermo-mechanical fatigue tests are con-
Atlanta, GA 30332-0405 ducted. A physically based model is developed for correlating crack initiation life based
on the experimental life data and predictions are made using the crack initiation
model. 关DOI: 10.1115/1.1924560兴

Keywords: Constitutive Modeling, Crack Initiation, Directionally Solidified Superalloy,


Fatigue Life Prediction

1 Introduction ture, can lead to one of several possible damage mechanisms.


Consequently, any physically based crack initiation model must
Nickel-base superalloys are technologically significant materi-
capture the degradation mechanisms related to the microstructure
als that are used extensively in applications that require a material
and how they interact with the environment and cyclic inelastic
with high strength, good creep, and fatigue and corrosion resis- behavior.
tance, even at elevated temperatures. Some of the applications
include turbine blades, turbine discs, burner cans, and vanes. The
operating temperatures of these components can range from room
temperature 共RT兲 to very high temperatures 共⬎1000° C兲 and they 2 Material Microstructure and Composition
are exposed to complex mechanical loading superimposed with DS GTD-111 nickel-base superalloy has the chemical compo-
thermal transients; for example, creep deformation interacts with sition as shown in Table 1. The as-cast structure contains coherent,
thermal and/or mechanical fatigue. Specialized processing tech- ordered ␥⬘ intermetallic precipitates dispersed uniformly in the
niques are employed to enhance mechanical properties, such as austenitic ␥ phase as shown in Fig. 1. There is a bi-modal distri-
elimination of grain boundaries along the turbine blade axis by bution of ␥⬘ precipitates, with cuboidal primary precipitates
directional solidification. However, this introduces new challenges 共0.5– 1 ␮m兲 and spheroidal secondary precipitates
in constitutive modeling of the deformation behavior and life 共0.05– 0.2 ␮m兲, with an overall volume fraction of approximately
prediction. 46%. These precipitates are interspersed in the relatively soft ma-
The deformation mechanisms change with varying loads, orien- trix and are separated by the thin matrix channels. The grains are
tations, strain rates, and temperatures. A physically based consti- columnar and are roughly 125 mm long and 2 mm wide, which
tutive relation that can simulate the structural response in order to limits the role of grain boundaries in crack initiation. Primary
facilitate component life prediction should incorporate the active dendrite stems grow parallel to the solidification direction. These
deformation mechanisms as a function of stress level, strain level, stems are accompanied by secondary and tertiary dendrite arms,
and temperature. The crystal plasticity approach provides a good which grow outward along 关100兴 and 关010兴 directions. MC car-
compromise between manageability and modeling capabilities, es- bide inclusions, rich in Ti and Ta, are located in the interdendritic
pecially when modeling components where the size, morphology, regions without any preferential orientation. The carbides are
and crystallographic orientation of the grains control the aniso- 50– 150 ␮m long with a high aspect ratio 共⬃10: 1 or greater兲. In
tropic deformation behavior 关1兴. A physically based constitutive addition to these carbides, M23C6 and ␥ / ␥⬘ eutectic nodules are
model is outlined to capture the homogenized deformation re- located in the interdendritic channels. The ␥ / ␥⬘ eutectics are duc-
sponse in a directionally solidified superalloy, DS GTD-111. The tile and vary in size from 30 to 150 ␮m.
material parameters are determined based on experimental low The observed deformation response in DS GTD-111 is consis-
cycle fatigue 共LCF兲 and creep data in both the longitudinal 共L兲 tent with that of other nickel base superalloys and can be summa-
and transverse 共T兲 orientations. Fatigue tests with short hold times rized as follows 关3兴:
are also conducted to accurately capture the short-term primary
creep behavior. 1. The yield strength increases with increasing temperature up
The combination of cyclic centrifugal forces experienced in to a peak temperature of 750° C, beyond which it decreases.
blades at high temperatures in an aggressive environment, coupled 2. Slip is generally dictated by the resolved shear stresses along
with the heterogeneity and inherent anisotropy of the microstruc- the favorable crystallographic slip planes. However, nickel
base superalloys do not obey Schmid’s law in all orienta-
1
tions and slip on any given plane can also depend on the
Author to whom correspondence should be addressed.
Contributed by the Materials Division for publication in the JOURNAL OF ENGINEER-
resolved stresses on other planes 共e.g., dislocation cross slip
ING MATERIALS AND TECHNOLOGY. Manuscript received October 5, 2004. Final manu- planes兲 due to dislocation core spreading effects in the ␥⬘
script received February 2, 2005. Review conducted by: R. Craig McClung. precipitates.

Journal of Engineering Materials and Technology JULY 2005, Vol. 127 / 325
Copyright © 2005 by ASME

Downloaded 25 Apr 2011 to 132.170.11.38. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Table 1 Nominal composition of DS GTD111

3. There is tension 共T兲-compression 共C兲 asymmetry of the flow


stress. The nature of this asymmetry is orientation and tem-
perature dependent 关4–8兴.
4. The initial yield behavior is nearly temperature path history
independent, i.e., if a sample is first deformed at a high
temperature at which the yield strength is high and then
deformed at a low temperature, the material response is
similar to a virgin material deformed at the lower
temperature 关9兴.
Fig. 2 Elastoplastic decomposition of the deformation gradi-
Any constitutive model should take these points into ent †11,12‡
consideration.

Nslip
3 Crystal Plasticity Formulation
A homogeneous constitutive model is proposed in this paper for Lp = Dp + Wp = 兺

␥˙ ␣共s> ␣ 丢 m
=1
> ␣兲o o 共2兲
which no explicit distinction is made between the matrix and pre-
cipitate phases. A crystal plasticity framework is employed to cap- where the plastic rate of deformation 共Dp兲 and plastic spin 共Wp兲
ture the orientation-dependent material response. As the tempera- are given by the symmetric and anti-symmetric parts of the plastic
ture increases, strain rate effects are increasingly important and a velocity gradient, i.e.,
rate-dependent method is employed. Crystal plasticity approaches 1 1
have been used extensively in modeling single crystal superalloys Dijp = 2 共Lijp + Lijp兲, Wijp = 2 共Lijp − Lijp兲 共3兲
关1,7,10兴 and the general methodology can be briefly described as It should be noted that the intermediate, relaxed configuration in
follows. A multiplicative rule is used for the deformation gradient, Fig. 2 is not unique since any superimposed rigid body rotation on
F, given as 关11,12兴 the deformation gradient will also satisfy Eq. 共1兲. Uniqueness of
F = Fe · Fp 共1兲 the intermediate configuration is obtained by assuming that the
substructure spin is co-rotational with the lattice.
As shown in Fig. 2, the deformation gradient is separated into two
parts: 共i兲 the plastic deformation gradient 共Fp兲 formed due to dis- 3.1 Constitutive Model. A rate-dependent crystal plasticity
location motion, associated with a change in the shape of the grain formulation is used. All slip systems are assumed active above the
but not its crystal lattice, and 共ii兲 the elastic deformation gradient threshold stress, with the shearing rate of the ␣th system related to

冓 冔 再冓 冔冎
共Fe兲, which essentially models reversible elastic stretch and rigid its associated viscous overstress by
body rotation of the lattice. For any given lattice, the close-packed ␶␣v n
␶␣v n+1

planes act as the slip planes with unit normal vector m > ␣o in the ␥˙ ␣ = ␥˙ o⌰共T兲 exp Bo sgn共␶␣ − ␹␣兲 共4兲
D␣ D␣
reference configuration for each of the ␣ slip systems, along
which the dislocations move in the slip direction in the reference where the slip system viscous overstress ␶␣v is given in terms of
configuration with unit vector s> ␣o . The shearing rates ␥˙ ␣ along each the resolved shear stress ␶␣ by
of the active slip systems 共␣兲 depend on the resolved shear stress ␮
on the slip systems. The macroscopic plastic velocity gradient is ␶␣v = 兩␶␣ − ␹␣兩 − ␬␣ 共5兲
obtained by summing over all slip systems in the intermediate ␮o
relaxed configuration according to Here, ␹␣ is the backstress and ␬␣ is the threshold stress on each
slip system. Increase in the threshold stress leads to an increase in
size of the viscoplastic flow potential, while the back stress results
in a shift of the potential surface. The threshold stress may be
viewed as the resistance to plastic flow arising from statistical
strengthening mechanisms associated with an increase of disloca-
tion density, solid solution strengthening, etc. The back stress dis-
plays a directional dependence and is associated with several fea-
tures of the heterogeneous material at the microscale, including
internal stresses that develop with deformation due to dislocation
pile-up at obstacles such as precipitate particles, grain or phase
boundaries, differential yielding with strain in hard 共dislocation
walls兲 and soft regions of the microstructure, and statistical inter-
actions of dislocations bypassing barriers. Both back stress, ␹␣,
and the threshold stress, ␬␣, depend on the history of temperature
and viscoplastic deformation.
Equations 共4兲 and 共5兲 contain several other additional param-
eters defined as follows: ␥˙ o is a reference shearing rate, n governs
the power law creep regime, Bo and 共n + 1兲 collectively govern the
power law breakdown behavior at higher strain rates where the
response becomes nearly rate independent, D␣ is the drag stress
Fig. 1 Two-phase microstructure of DS GTD-111 superalloy †2‡ that is weakly dependent on the history of temperature and visco-

326 / Vol. 127, JULY 2005 Transactions of the ASME

Downloaded 25 Apr 2011 to 132.170.11.38. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
plastic deformation when the behavior of the ␥ and ␥⬘ phases are the current set of constitutive relations through a critical threshold
homogenized, ␮ is the shear modulus, with ␮o being the value of stress 共␬␣兲, i.e.,
the shear modulus at absolute zero, and ⌰共T兲 is the diffusivity
parameter given by ␬␣ = ␬␣c + ␬␣e 共9兲

⌰共T兲 = exp − 冉 冊 Qo
RT
for T 艌
Tm
2
;
where
␬␣c = ␬o␣共T兲 + h pe␶␣pe + hse␶se

+ hcb兩␶␣cb兩

再 冋 冉 冊 册冎
共6兲 and
2Qo Tm Tm
⌰共T兲 = exp − ln +1 for T 艋 Nslip Nslip
RTm 2T 2
where Qo is the activation energy for thermally activated disloca-
␬˙ ␣e = ho 兺
␤=1
q␣␤兩␥˙ ␤兩 − h␬s␬␣ 兺

兩␥˙ ␤兩 − h ⌰共T兲具␬␣ − ␬
=1
s th典
rs

tion bypass of obstacles, R is the universal gas constant, and Tm is


The back stress evolution for homogenized ␥-␥⬘ models can be
the absolute melting point temperature.
written in the hardening-dynamic recovery plus static thermal re-
The effective strain rate sensitivity exponent on the stress 共in-
covery format as
verse strain rate sensitivity兲 is given by

m=
⳵ ln ␥˙ ␣ ⳵ ln ␥˙ ␣
⳵ ln ␶ ␣ = ␣ =n+
⳵ ln ␶v
Bo共n + 1兲
冓 冔
␶␣v
␶ˆ o␮/␮o ␶ˆ o␮/␮o
n
共7兲
␹˙ ␣ = h␹兩␥˙ ␣兩sgn共␶␣ − ␹␣兲 − h␹d␹␣兩␥˙ ␣兩 +

− ⍀␹␣
冉 1 ⳵R␹
+
R ␹ ⳵ T h ␹d ⳵ T

1 ⳵ h ␹d ␣
␹ Ṫ

共10兲
The anomalous yield behavior of ␥-␥⬘ nickel base superalloys,
i.e., the yield stress increases with temperature in the intermediate where ⍀␹␣ = h␹s共T兲兩␹␣兩r␹s
is the static thermal recovery term, h␹,
temperature regime, is incorporated in the threshold stress. The h␹d, h␹s, r␹s are material parameters, and R␹ = h␹ / h␹d. The
anomalous behavior has been attributed to mechanisms involving temperature-rate-dependent term is necessary to properly model
兵111其 具110典 octahedral slip within the ␥⬘ precipitates. Several hysteresis behavior under thermomechanical fatigue 共TMF兲 关18兴,
models attempt to explain this mechanism, including the Takeuchi since the model parameters that govern the back stress evolution
and Karamoto model 关13,14兴, the Paidar-Pope-Vitek 共PPV兲 model rates are themselves functions of temperature. Note that for single
关5兴, and the model proposed by Hirsch 关15兴. Though recent ex- phase models, the back stress evolution equation can likely be
perimental evidence 关16兴 seems to support the Hirsch theory, the neglected, but it is essential to describe the overall average behav-
PPV model has perhaps been the most widely used model 关10兴, at ior of the crystallographic grains in homogenized ␥-␥⬘ models. It
least in the anomalous temperature regime 共room temperature to is a very significant fraction of the flow stress 共typically⬎ 30% 兲
750° C兲. for Ni-base superalloys 关19兴. The constitutive equations are sum-
According to the PPV model 关5兴, in the anomalous behavior marized in Table 2.
temperature regime 共room temperature to 750° C兲 具101̄典 screw
dislocations on 共111兲 planes split into two super-Shockley partials
separated by an antiphase boundary 共APB兲. There are two con-
4 Experiments
figurations of the a / 2具101̄典 screw dislocation: a glissile configu-
ration with its core spread in the 共111兲 plane and a sessile con- Test sections were sectioned from two batches of cast plates.
Solid specimens with a uniform gage length of 12.7 mm 共0.5 in.兲
figuration with a nonplanar core spread in the 共11̄1兲 plane. The
and a diameter of 6.35 mm 共0.25 in.兲 were machined from sec-
core transformation is explained in three steps: 共i兲 constriction of
tions oriented in the longitudinal 共L兲 and transverse 共T兲 directions.
the glissile core on the 共111兲 plane, 共ii兲 movement of the con-
The specimen was prepared according to ASTM Standard E606
stricted dislocation along the 共010兲 plane, and 共iii兲 spreading of
关20兴. Tests were performed on a 45 kN 共10 kip兲 axial servo-
the dislocation on the 共11̄1兲 plane. The glissile core has to con- hydraulic MTS testing machine with dual-channel controllers and
strict first for this to happen; hence, any shear stress that aids in TestStar software 共Testware SX 4.0D兲. Load, strain, and tempera-
the constriction of the partial dislocations will assist cross-slip, ture data were recorded digitally. The tests were conducted under
leading to formation of sessile locks which act as obstacles to the mechanical strain control using an extensometer with ceramic tips
glissile dislocations. Slip on the octahedral planes will be influ- in direct contact with the specimen to obtain the total strain. The
enced by the stresses on the primary 共step 共i兲兲 and secondary cross mechanical strain ratio R␧ = ␧min / ␧max was held constant at R␧ =
slip planes 共step 共iii兲兲, along the direction perpendicular to the −1. An Ameriterm 2 kW RF induction heater provided heat to the
partial dislocations, as well as the stresses acting on the cube specimens. Five K-type thermocouples were directly attached to
plane 共step 共ii兲兲. The additional dependence of the activation en- the specimen 共three within the gage section兲 to achieve a near-
thalpy for the glissile to sessile transformation on the shear stress uniform temperature distribution within the gage section, and the
components 共␶ pe, ␶se, and ␶cb, i.e., the shear stresses on the pri- induction coil was iteratively modified to achieve a maximum
mary, secondary, and cube slip systems, respectively兲 leads to a gage section temperature variation of either 1% of Tmax or ±3 K.
non-Schmid effect. Since the enthalpy is also dependent on the For both LCF and TMF tests, the temperature was maintained
sign of the stress components for steps 共i兲 and 共ii兲, it will differ in using a PID temperature controller. In TMF tests, the temperature
tension and compression. This leads to a tension-compression reading was transmitted to the TestStar controller and used to
asymmetry that is orientation dependent. It should be noted that calculate the components of thermal and mechanical strain. By
step 共iii兲 is independent of the sense of loading. following the procedural guidelines in ASTM E2368 关21兴, TMF
To include the non-Schmid effects, the yield criterion for octa- testing can accurately simulate damage caused by the combined
hedral slip systems is modified based on the PPV theory as pro- effects of thermal cycling and mechanical fatigue. Prior to each
posed by Qin and Bassani 关17兴. Each octahedral slip system is test, a relationship for thermal strain 共␧th共T兲兲 is determined using
active only when the Schmid stress and a linear combination of the free response under unloaded thermal cycles. This correlation
the additional stress terms exceed a critical value, i.e., is replayed during TMF cycling. Thus, all TMF tests were con-
␶␣ = h pe␶␣pe + hse␶se

+ hcb兩␶␣cb兩 = ␶cr 共8兲 ducted using closed-loop, temperature-based thermal strain con-
trol. A total of 36 strain-controlled isothermal LCF tests and 5
where ␶ pe, ␶se, and ␶cb are the non-Schmid shear stresses and ␶cr is TMF tests were conducted, as shown in Table 3. Three types of
the critical resolved shear stress. This formulation is embedded in cycles were used: continuous cycling with 共a兲 no hold time 共CC兲,

Journal of Engineering Materials and Technology JULY 2005, Vol. 127 / 327

Downloaded 25 Apr 2011 to 132.170.11.38. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Table 2 Summary of constitutive model

共b兲 with 2 min hold time in tension 共HT兲, and 共c兲 with 2 min hold It is assumed that the flow rule given in Eq. 共4兲 is suitable for
time in compression 共HC兲. A constant strain rate of 0.005 s−1 was modeling the homogenized ␥-␥⬘ phases, but with a much higher
used during the loading and unloading ramps. intrinsic lattice friction necessary to model the higher activation
The cycles to crack initiation in Table 3 is defined as the num- energy for slip compared to modeling only a single phase, such as
ber of cycles at which a 20% drop in the maximum axial load ␥. It is commonly observed that the creep exponent of Ni-base
occurred. In many instances, the cycle with the critical load drop- superalloys is more or less invariant with respect to ␥⬘ volume
off coincided with complete fracture of the gage section. fraction and morphology, suggesting that slip is indeed restricted
in the precipitates, which provide primary elevated temperature
5 Implementation and Determination of Material Pa- strength.
rameters The 12 octahedral 兵111其具110典 slip systems are appropriate for
The constitutive model was implemented as User MATerial FCC single crystals and are active for the entire temperature
共UMAT兲 subroutine in ABAQUS 关22兴, based on the implicit inte- range. The six cube slip systems 兵100其具110典 type 共three planes and
gration scheme of McGinty 关23兴. A hyperelastic formulation was two directions= six systems兲 may be active at high homologous
used, valid for arbitrary finite strain. A time step subincrementa- temperatures and high resolved shear stress 共in ␥⬘ phase兲, but their
tion scheme and a linear search algorithm were employed in the role is less well understood and characterized. A precise mecha-
constitutive subroutine to ensure convergence. nism is still largely conjectural, since the identification of the role

328 / Vol. 127, JULY 2005 Transactions of the ASME

Downloaded 25 Apr 2011 to 132.170.11.38. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Table 3 Test matrix for isothermal LCF and TMF. „a… Test matrix for TMF in longitudinal „L… and
transverse „T… orientations using in-phase „IP… and out-of-phase „OP… thermomechanical load-
ing histories. Cycle type CC—continuous cycling without hold time, Cycle type 2 min
C—continuous cycling with 120 s hold time at peak strain in compression, Cycle type 2 min
T—continuous cycling with 120 s hold time at peak strain in tension. „b… Test matrix for iso-
thermal LCF in longitudinal orientation and „c… test matrix for isothermal LCF in transverse
orientation.

of cube slip is challenging. For homogenized ␥-␥⬘ single crystals, roscopic cube slip is a manifestation of octahedral slip and takes
the role of cube slip is usually assessed by considering the break- place by multiple cross slip events on octahedral planes, produc-
down of the capability of octahedral slip to model the stress- ing zig-zag dislocation lines. This enables the dislocations to
strain-time behavior as a function of temperature, which of course travel large distances along matrix channels, thus effectively
depends on other elements of the constitutive framework. The shearing material along the cube planes. In addition, actual cube
work of Bettge and Osterle 关24兴 on SC16 has revealed that mac- slip has been observed 关16兴 in the ␥⬘ precipitates at higher tem-

Journal of Engineering Materials and Technology JULY 2005, Vol. 127 / 329

Downloaded 25 Apr 2011 to 132.170.11.38. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
peratures and along orientations closer to 关111兴. Experimentally, simulations involving different thermomechanical histories in-
softening occurs as the orientation is rotated from the 关100兴 ori- cluding cyclic deformation with and without hold periods, as well
entation 关25兴. When only octahedral slip is active, hardening is as creep deformation, using a set of parameters suggested by the
predicted as the loading axis is rotated from the 关100兴 direction, optimization code. Then the model response is compared to the
which is inconsistent with experimental observation. Both sources actual experimental response generating a value of an objective
of cube slip, zig-zag cross slip in the ␥ phase and motion of function defined on the basis of the norm of the error between the
dislocation segments on cube planes in the ␥⬘ phase, are lumped experimental and predicted deformation responses. The Epogy
together in the present model, treated via the activation of the code by Synaps 关28兴 is used to minimize the objective function by
cube slip systems. Both the octahedral and cube slip systems are iteratively modifying the material parameters based on minimiza-
assumed to be active in this model at all times. tion of the objective function. Epogy employs four widely used
The material parameters for the octahedral slip systems are de- search methods 共linear simplex, downhill simplex, gradient, and
termined from data obtained from uniaxial cyclic loading of lon- genetic algorithms兲 to assure stability and enhance the probability
gitudinal specimens along the 关100兴 orientation, for which the the that an optimum set of material parameters will be obtained
loading axis is oriented along the grain growth direction. Cube among various local minima. The rate of convergence and effec-
slip is not activated in this orientation for small deformations. The tiveness of the optimization is highly dependent on the objective
material parameters for the cube slip systems are obtained using function, which in turn depends on the type of loading history. For
data from transverse specimens 共loading perpendicular to the uniaxial cyclic loading, the error function is defined as

兺 共␴
grain growth direction兲 due to lack of data in other orientations. It
should be noted that the response of directionally solidified mate- Error = i
exp − ␴pre
i
兲 2W i 共11兲
i
rial in the transverse orientation is an average response compris-
ing several grains. A total of 30 grains with random transverse where ␴iexp is the experimentally measured stress along the load-
orientations are used to obtain this average behavior along the ing direction, ␴pre
i
is the stress predicted by the model at the same
loading axis. The Taylor constraint for intergranular interactions is strain, and Wi is a weighting parameter for the ith strain value.
assumed in this averaging process, such that each grain 共with The weighting parameter is varied from 0.8 to 1.2, depending on
different orientations along the transverse direction兲 is subjected which region of the stress-strain response is being simulated. Less
to the same deformation gradient; this approach is convenient in weight is placed on the elastic portions of the deformation 共Wi
that a single element can be used to obtain the initial estimate of
= 0.8兲 while Wi = 1.2 is used during inelastic flow. During strain
the material parameters, and is sufficiently accurate for slip of
hold periods in which inelastic deformation and stress relaxation
polycrystals with high symmetry cubic structures. A more rigor-
ous finite element analysis is subsequently run with 30 hexagonal occurs, Wi = 1.0.
grains, each containing 96 elements, to optimize the fit of material Since the static thermal recovery terms are influential only for
long hold periods and into the steady state creep regime, these
parameters.
terms can be neglected in the preliminary parameter estimation
Under cyclic loading, lattice rotation may be insignificant, but
exercise that uses only continuous cycling data without hold
in the presence of creep or cyclic ratcheting, lattice rotation may
times. The results of this preliminary exercise provides tighter
build up at sites of stress concentration or near interfaces such as
bounds on those parameters that are less sensitive to rate for pur-
grain boundaries. However, the creep and ratcheting strains are
poses of later optimization exercises that include cyclic data with
small in the cyclic loading experiments conducted to determine
strain holds and creep data intended to refine the estimates of
the material parameters, so these lattice rotations are neglected for
material parameters. The error norm used for estimating param-
the purpose of estimating the material parameters. Several other
eters associated with creep behavior is
simplifications can be made regarding the deformation behavior of
this material at this temperature. In the temperature range under
consideration, the precipitate structures 共size, shape, morphology兲
Errorcreep = 兺 共␧
i
i
exp − ␧pre
i
兲2E2av 共12兲
are stable, so higher-scale microstructural changes such as coars-
ening of precipitates can be neglected in the evolution equations. where ␧iexp − ␧pre
i
is the difference between experimentally mea-
From the cyclic deformation data, it is observed that the initial sured strain and predicted strain at a given time instant. This error
yield strength does not change significantly with cyclic loading. norm employs the average value of the Young’s modulus, Eav, in
This indicates that the threshold and drag stress are nearly con- the specified orientation to scale the errors in terms of pseudo-
stant and that the backstress is the primary internal state variable stresses. For this final optimization step, the total error is the sum
that evolves with deformation. The threshold stress is mainly lim- of the errors for the cyclic and creep deformation data,
ited to the critical value ␬␣c that incorporates non-Schmid terms as Error = Errorcyclic + Errorcreep 共13兲
previously explained, and the drag stress 共D␣兲 is kept constant.
The remaining material parameters must be physically admissible. The material parameters obtained are listed in Table 4. Some of
This calls for a systematic procedure for determining the con- the material parameters are not listed due to unavailability of ex-
stants. First, certain physical parameters such as elastic properties perimental data for fitting parameters. Static thermal recovery ef-
and activation energies of thermally activated diffusional pro- fects in evolution of ␬ are omitted, for example. Accordingly,
cesses are obtained from literature. The activation energy is set to dashed entries in Table 4 signify zero values of the parameters.
Q = 309 kJ/ mol, based on the data from Daleo and Wilson on The most sensitive material parameters for this data set as de-
GTD-111 关26兴. Estimating the remaining material parameters by termined by the optimization scheme are h␹d and h␹s, which con-
using graphical methods in the analysis of deformation data is trol the backstress evolution. It should be noted that backstress
often not possible with viscoplasticity models because there is a plays an important role in Ni-base superalloys. Based on the tests
strong coupling between creep and cyclic plasticity. Approximate by Ferney et al. 关29兴 and Latif et al. 关30兴, it could be to the order
values of the material parameters can often be obtained, but a of 50%–60% of the flow stress. Experimental data used to fit
trial-and-error approach almost always must be employed to refine parameters and correlated simulations based on the constitutive
the parameters to accurately capture the experimental deformation model are compared in Figs. 3–9.
response. A more robust and efficient approach is to integrate the
constitutive model, simulating the actual loading histories,
using an optimization scheme to iteratively evaluate the model 6 Thermomechanical Fatigue (TMF)
parameters 关27兴. For TMF, a deformation gradient F␪ based on the thermal ex-
The general optimization scheme involves running several pansion is included in the multiplicative decomposition of the

330 / Vol. 127, JULY 2005 Transactions of the ASME

Downloaded 25 Apr 2011 to 132.170.11.38. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Table 4 Material parameters: „a… octahedral slip system constants ␹␣„0… = 0, ␬e␣„0… = 0; „b… cube
slip system constants ␹␣„0… = 0, ␬e␣„0… = 0; and „c… common constants

deformation gradient. An intermediate, thermally expanded, hot L␪o = Ḟ␪ · 共F␪兲−1 = ␣␪˙ I 共14兲
unstressed configuration and a hot plastically deformed relaxed
configuration are introduced 关31兴. Thus the total deformation gra- where ␣ is the thermal expansion coefficient 共assumed constant兲
dient is given by F = Fe · F p · F␪, where the linearized elastic defor- and I is the second rank identity tensor. Solving the differential
mation is described by Fe, the plastic deformation by F p, and the equation over a time step gives
thermal expansion/contraction by F␪. Assuming isotropic thermal ␪
Ft+⌬t = exp共Lo␪⌬t兲Ft␪ 共15兲
expansion, the velocity gradient associated with thermal expan-
sion effects is given as Material parameters are assigned temperature dependence using
third-order interpolative polynomial fits of values determined for
isothermal cyclic loading cases. The temperature is updated and

Fig. 3 Stress-strain response: experimental data and corre- Fig. 4 Stress-strain response: experimental data and corre-
lated simulations at 427° C „longitudinal, CC, first cycle… lated simulations at 760° C „longitudinal, CC, first cycle…

Journal of Engineering Materials and Technology JULY 2005, Vol. 127 / 331

Downloaded 25 Apr 2011 to 132.170.11.38. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 8 Stress-strain response: experimental data and corre-
Fig. 5 Stress-strain response: experimental data and corre-
lated simulations at 1038° C „transverse, CC, first cycle…
lated simulations at 871° C „longitudinal, HC, first cycle…

held constant for each Newton Raphson step and the material 7 Fatigue Crack Initiation
parameters relevant to that temperature are used. These TMF The crack initiation life is defined as the number of cycles at
simulations can be considered as true predictions which correlate which the load drops below 20% of the initial maximum stabilized
reasonably well with the experimental results for loading in both tensile load. The constitutive model described above is used to
the longitudinal and transverse orientations, as shown in Figs. predict the cyclic material response under imposed isothermal and
10–12.

Fig. 9 Primary and secondary creep responses: experimental


Fig. 6 Stress-strain response: experimental data and corre- data and correlated simulations at 871° C „longitudinal…
lated simulations at 982° C „longitudinal, CC, first cycle…

Fig. 10 Stress-strain response: comparison of experimental


Fig. 7 Stress-strain response: experimental data and corre- data with model predictions for in-phase „IP… TMF
lated simulations at 982° C „transverse, CC, first cycle… 538° C – 1038° C in the longitudinal orientation „first cycle…

332 / Vol. 127, JULY 2005 Transactions of the ASME

Downloaded 25 Apr 2011 to 132.170.11.38. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
lives of turbine blade materials subjected to high homologous
temperatures 关32–39兴. The approaches that have been developed
are categorized as energy-based, strain-based, and damage based.
Coffin 关32兴 modified the Coffin-Manson law to include environ-
mental effects via a cycle frequency factor. Neu and Sehitoglu
关33兴 extended Coffin-Manson concepts in a strain-based strategy
to address coupled cyclic plastic strain, creep, and oxidation dam-
age mechanisms. An energy-based approach was presented by Os-
tergren 关34兴 and later modified by Zamrik and Renauld 关35兴 and
included the monotonic tensile elongation and ultimate strength
data as functions of temperature. The model introduced by Rémy
et al. 关37兴 is restricted to a microstructurally small volume ele-
ment adjacent to an oxidized crack tip of a CT-type specimen; in
this case, the critical fracture stress of the alloy and the oxidized
alloy are required. Miller et al. 关38兴 extended damage rate equa-
tions discussed by McDowell et al. 关39兴 for small fatigue crack
propagation under thermo mechanical loading to address coupled
Fig. 11 Stress-strain response: comparison of experimental fatigue, creep, and oxidation effects in Ni-base superalloys. Most
data with model predictions for out-of-phase „OP… TMF of these models were developed for isotropic polycrystalline 共PC兲
538° C – 927C in the longitudinal orientation materials in which creep damage mechanisms such as void growth
and grain boundary slip were dominant. In this paper, a crack
initiation model is presented based on the observed damage
non-isothermal fatigue conditions for purposes of predicting fa- mechanisms for DS GTD-111.
tigue crack initiation life. Several isothermal fatigue 共IF兲 and TMF
life prediction models have been developed to predict the service 7.1 Observations. The stress response of DS GTD-111 under
TMF is plotted with respect to the mechanical strain ␧m 共total
strain less thermal strain兲 for several test cases in Figs. 13共a兲 and
13共b兲. Figure 13共c兲 plots the inelastic strain range 共mechanical
strain less elastic strain兲 versus the number of cycles to crack
initiation, ⌬␧in versus Ni, for in-phase 共IP兲 and out-of-phase 共OP兲
TMF cases. The mechanisms for crack initiation change with
loading history; however, no significant change in the initiation
mechanism is observed with change in orientation 共longitudinal
versus transverse兲.
Three primary damage mechanisms are identified: fatigue,
creep-fatigue, and oxidation-fatigue damage. A single stage II ma-
jor crack is observed in the isothermal fatigue tests without hold
times at all temperatures, with very little secondary cracking. At
lower temperatures 共艋650° C兲, cracks initiate at cracked carbide
inclusions near the surface as shown in Fig. 14共a兲. Initiation at
higher temperatures occurs at the specimen surface from an oxide
spike along the interdendritic region, as shown in Fig. 14共b兲. Car-
bide inclusions near the surface act as diffusion short circuits, thus
serving as preferential sites for oxidation. A softer ␥ depleted
layer exists below the oxide layer further from the surface⬘ due to
Fig. 12 Stress-strain response: comparison of experimental
data with model predictions for out-of-phase „OP… TMF diffusion of aluminum. For the isothermal fatigue tests with hold
538° C – 927° C in the transverse orientation

Fig. 13 Initial stress-mechanical strain responses of TMF cycled „a… longitudinal and „b… transverse
specimens. „c… Inelastic strain versus fatigue crack initiation lives for TMF and LCF experiments at
871° C

Journal of Engineering Materials and Technology JULY 2005, Vol. 127 / 333

Downloaded 25 Apr 2011 to 132.170.11.38. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Nifat = C1 f 1共⌬␧in兲C2 共17兲
where C1 and C2 are constants, and f 1 is a factor that accounts for
material orientation with respect to the applied loading direction
共i.e., L or T兲. To characterize this relation, data are needed from
low temperatures, high strain rates, and CC tests. Tests conducted
at and below 871° C are used to correlate Eq. 共17兲. The creep-
fatigue interaction mechanism is found to correlate to

Nicr = C3⌰cr共T,t兲 冉 tc + th
tc + 2th
冊 f2
共⌬␧in兲C4 共18兲

where C3, C4, and f 2 are constants 关33兴. The inelastic strain range,
⌬␧in, is used here since it captures orientation and time-dependent
deformation. The diffusion coefficient, ⌰cr, along with the cycle
Fig. 14 „a… Crack initiation at a subsurface carbide and „b… and dwell times, given by tc and th, respectively, account for time
crack initiation at the surface due to an oxide spike and temperature-dependent degradation mechanisms such as inter-
nal void nucleation and growth that are not captured by Eq. 共17兲.
The diffusion term is given by

冕 冉 冊
tc
times 共2 min tension or compression兲, multiple voids are observed 1 Qcr
around ␥ / ␥⬘ eutectic nodules where crack initiation takes place, ⌰cr共T,t兲 = exp − dt 共19兲
tc 0
RT共t兲
thus, indicating a creep degradation mechanism. From Table 3, the
crack initiation life is shorter in longitudinal oriented specimen in Here Qcr is the activation energy, T共t兲 is the temperature history,
the tests with hold time as compared to those without hold times and R is the universal gas constant. Specimens subjected to dwell
due to the contribution from a creep component in addition to the times experience the most significant stress relaxation induced by
fatigue and oxidation components. This is not true in the trans- viscoplastic deformation. The correlation of Eq. 共18兲 involves HT
verse orientation at higher temperatures 共⬎871° C兲 where the hold and HC tests at or above 871° C.
times in compression give higher crack initiation lives. More ex- The term relating oxidation-fatigue interaction to crack initia-
perimental tests will be conducted and fatigue specimens will be tion is motivated by the oxide spiking mechanism. The interaction
analyzed as a part of future work to understand this anomalous is related to cycle-dependent repeated fracture of the oxide film.
behavior in the transverse orientation at higher temperatures. In This process is controlled by the elastic part of the mechanical
the TMF tests, the material microstructure plays a much more strain range, ⌬␧el, the oxidation growth kinetics, and the phasing
important role in the IP tests as compared to the OP tests. MC between the mechanical and thermal loading captured through a
carbide inclusion cracking is observed near the specimen surface phasing factor, ⌽ox, i.e.,
in the IP tests and some voids are also observed around the ␥ / ␥⬘
eutectic nodules. Oxide spiking is the primary mode of crack ini- Niox = C5关⌽ox⌰ox共tc + th兲兴C6⌬␧el
C7
共20兲
tiation in the OP TMF tests and oxide spallation is also observed Here C5, C6, and C7 are constants 关33兴. The diffusion factor, ⌰ox,
but to a much smaller extent. in analogy to Eq. 共19兲 for creep, captures the oxide film growth
7.2 Creep-Fatigue-Environment Crack Initiation Model. kinetics. The phasing factor is given by
The model in this section is based on application of the TMF ⌽ox = exp关− C8共␧˙ th/␧˙ m + 1兲2兴 共21兲
model developed by Neu and Sehitoglu 关33兴. Assuming a unique
relationship between the damage fraction and cycle fraction with This factor has a maximum value of unity for linear OP TMF, the
respect to cycles to crack initiation for each damage mode, the condition most susceptible to the oxide spiking mechanism. Hys-
total crack initiation life can be represented in terms of the indi- teresis data from isothermal tests above 871° C, along with dwells
vidual damage components 共fatigue, creep-fatigue, and oxidation- and TMF tests, are used to determine the constants for Eqs. 共20兲
fatigue, respectively兲, and can be written as 关33兴 and 共21兲. Constants are summarized in Table 5.
7.3 Model Correlation and Predictions. Some trends pre-
1 1 1 1 dicted by the model are shown in Fig. 15 for continuously cycled
= + + 共16兲 longitudinally oriented DS GTD-111 specimens. For both isother-
Ntot Nfat Ncr Nox
mal 共Fig. 15共a兲兲 and nonisothermal 共Fig. 15共b兲兲 cases, the crack
Relationships for each of these components are determined by initiation life is characterized by the stress-strain response at high
conducting LCF and TMF tests to isolate the damage mecha- mechanical strain ranges since the fatigue term dominates.
nisms. Mathematical formulations are developed to relate each Under isothermal conditions, the change in slope of predicted
mechanism to temperature, grain orientation, hold times, etc. The life below the mechanical strain range of 1.0% relates to a transi-
fatigue damage mechanism is governed by 关33兴 tion from predominantly fatigue to a coupling of oxidation and

Table 5 Constants for the crack initiation model

334 / Vol. 127, JULY 2005 Transactions of the ASME

Downloaded 25 Apr 2011 to 132.170.11.38. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 15 Predicted trends and selected experimental results for „a… isothermal LCF and „b… TMF of longitudinal
DS GTD111 under CC

fatigue mechanisms. As the temperature is decreased below for Structural Calculations Part 1–Model Presentation,” J. Eng. Mater. Tech-
nol., 162共113兲, pp. 162–170.
871° C, there is insufficient thermodynamic driving force for oxi- 关2兴 Trexler M., and Sanders, T., 2002, personal communication, Georgia Institute
dation and the role of environment is limited as a result. Under of Technology.
TMF conditions at lower mechanical strain ranges 共Fig. 15共b兲兲, 关3兴 Westbrook, J. H., 1996, “Superalloys 共Ni base兲 and dislocations—An Intro-
the OP cycle is predicted to be more damaging than the IP cycle. duction,” Dislocat. Solids, 10, pp. 1–25.
关4兴 Lall, C., Chin, S., and Pope, D., 1979, “Orientation and temperature depen-
This prediction is consistent with experiments since the oxide dence of the yield stress of Ni3 共al, nb兲 single crystals,” Metall. Trans. A, 10A,
spiking, which is more detrimental to life than oxide spallation, is pp. 1323.
experienced during OP cycling. During IP cycling with ⌬␧m 关5兴 Paidar, V., Pope, D. P., and Vitek, V., 1984, “A theory of the anamolous yield
⬍ 1.0%, life is dominated by the fatigue mechanism. With the behavior in LI2 ordered alloys,” Acta Metall., 32共3兲, pp. 435–448.
关6兴
increase of ⌬T from 389° C to 500° C, the constitutive model Umakoshi, Y., Pope, D. P., and Vitek, V., 1984, “The asymmetry of the flow
stress in Ni3 共Al, Ta兲 single crystals,” Acta Metall., 32共3兲, pp. 449–456.
predicts ⌬␧in to increase. Tests conducted at same temperature and 关7兴 Osterle, W., Bettge, D., Fedelich, B., and Klingelhoffer, K., 2000, “Modelling
total strain amplitudes on longitudinal and transverse oriented DS the orientation and direction dependence of the critical resolved shear stress of
GTD-111 specimens lead to larger inelastic strain ranges for trans- Nickel-base superalloy single crystals,” Acta Mater., 48, pp. 689–700.
verse specimens since transverse specimens have a larger elastic 关8兴 Jiao, F., Bettge, D., Osterle, W., and Ziebs, J., 1996, “Tension compression
asymmetry of the 共001兲 single crystal superalloy SC16 under cyclic loading at
stiffness 共Figs. 6 and 7兲. Consequently, the lives of longitudinal elevated temperatures,” Acta Metall., 44共10兲, pp. 3933–3942.
specimens typically outlast those of transverse specimens when 关9兴 Vitek, V., Pope, D. P., and Bassani, J. L., 1996, “Superalloys 共Ni base兲 and
tested at the same total strain range. The crystallographic orienta- dislocations—An Introduction,” Dislocat. Solids, 10, pp. 135–186.
tion dependence in the constitutive model captures this phenom- 关10兴 Lee, E. H., 1969, “Elastic-Plastic Deformations at Finite Strains,” ASME J.
Appl. Mech., 36, pp. 1–6.
enon, predicting the trends especially at the higher mechanical 关11兴 Sheh, M. Y., and Stouffer, D. C., 1988, “Anisotropic Constitutive model for
strain ranges. Nickel-base Single Crystal Superalloys,” NASA Report No. CR-182157.
关12兴 Bilby, B., Bullough, R., and Smith, E., 1955, “Continuous Distributions of
8 Summary Dislocations: A New Application of the Methods of Non-Riemannian Geom-
etry,” Proc. R. Soc. London, Ser. A, 231, pp. 263–273.
The deformation and failure mechanisms occurring in a direc- 关13兴 Takeuchi, S., and Karamoto, E., 1971, “Anomalous temperature dependence of
tionally solidified Ni-base superalloy, DS GTD-111, have been the yield stress in Ni3Ga,” J. Phys. Soc. Jpn., 31, pp. 1282.
关14兴 Takeuchi, S., and Kuramoto, E., 1973, “Temperature and Orientation Depen-
characterized as a function of temperature and loading history. A dence of the Yield Stress in Ni3Ga Single Crystals,” Acta Metall., 21, pp.
continuum crystal plasticity model is developed to characterize 415–425.
the material behavior in the longitudinal and transverse orienta- 关15兴 Hirsch, P. B., 1992, “A new theory of the anomalous yield stress in LI2 alloys,”
tions. Isothermal and thermomechanical uniaxial fatigue tests with Philos. Mag. A, 62共3兲, pp. 569–612.
关16兴 Sun, Y. Q., and Hazzledine, P. M., 1996, “Geometry of dislocation glide in L12
hold times and creep tests are conducted at temperatures ranging
␥⬘-phase: TEM observations,” Dislocat. Solids, 10, pp. 27–68.
from room temperature 共RT兲 to 1038° C to characterize the defor- 关17兴 Qin, Q., and Bassani, J. L., 1992, “Non-Schmid yield behavior in single crys-
mation response. Using isothermal test data to fit constants, IP and tals,” J. Mech. Phys. Solids, 40共4兲, pp. 813–833.
OP TMF responses are adequately predicted. The fatigue speci- 关18兴 McDowell, D. L., 1992, “A nonlinear kinematic hardening theory for cyclic
mens are studied to identify the damage mechanisms and a physi- thermoplasticity and thermoviscoplasticity,” Int. J. Plast., 8, pp. 695–728.
关19兴 Castelli, M. G., and Ellis, J. R., 1993, “Improved techniques for thermome-
cally based model for predicting crack initiation life is developed chanical testing in support of deformation modeling,” Symposium on Thermo-
based on the experimental stress-strain data. Three primary dam- mechanical Fatigue Behavior of Materials 共STP 1186兲, Sehitoglu., J. 共Ed.兲, San
age mechanisms are identified: fatigue, creep-fatigue, and Diego, CA, ASTM, pp. 195–211.
oxidation-fatigue induced damage. Crack initiation life predic- 关20兴 ASTM E-606-92, 2001, “Standard Practice for Strain-Controlled Fatigue Test-
ing,” ASTM, West Conshohocken, PA.
tions, based on the life models for each damage mechanism, fol- 关21兴 ASTM E2368, 2004, “Standard Practice for Strain-Controlled Thermome-
low the experimental trends. chanical Fatigue Testing,” ASTM International.
关22兴 ABAQUS, 2003, Hibbitt, Karlsson, and Sorensen, Inc., Providence, RI, v6.3.
关23兴 McGinty, R. D., 2001, “Multiscale Representation of Polycrystalline Inelastic-
Acknowledgment ity,” Ph.D. thesis, Georgia Institute of Technology.
The author acknowledges the help from Matt Trexler, Garry 关24兴 Bettge, D., and Osterle, W., 1999, “Cube Slip in Near-关111兴 Oriented Speci-
ments of a Single-Crystal Nickel-Base Superalloy,” Scr. Mater., 40共4兲, pp.
Donaghy, and Tam Nguyen. 389–395.
关25兴 Mielek, J., Novak, V., Zarubova, N., and Gemberle, A., 1997, “Orientation
Dependence of Plastic Deformation in NiAl Single Crystals,” Mater. Sci. Eng.,
References A, 234–236, pp. 410–413.
关1兴 Meric, L., Poubanne, P., and Cailletaud, G., 1991, “Single Crystal Modeling 关26兴 Daleo, J. A., and Wilson, J. R., 1998, “GTD-111 Alloy Material Study,” J.

Journal of Engineering Materials and Technology JULY 2005, Vol. 127 / 335

Downloaded 25 Apr 2011 to 132.170.11.38. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
engineering for gas turbines and power, 120, pp. 375–382. Low-Cycle Fatigue,” J. Test. Eval., 4共5兲, pp. 327–339.
关27兴 Tanner, A. B., McGinty, R. D., and McDowell, D. L., 1999, “Modeling Tem- 关35兴 Zamrik, S. Y., and Renauld, M. L., 2000, “Thermo-Mechanical Out-of-Phase
perature and Strain Rate Sequence Effects on OFHC Copper,” Int. J. Plast., Fatigue Life of Overlay Coated IN-738LC Gas Turbine Material,” Thermome-
15, pp. 575–603. chanical Fatigue Behavior of Materials: Third Volume, ASTM 1371, H. Sehi-
关28兴 Epogy, Synaps Inc., 2004, Atlanta, GA, USA, v2004A. toglu, and H. J. Maier, Eds., American Society for Testing and Materials,
关29兴 Ferney, V., Hautefeuille, M., and Clavel, M., 1991, “Multiaxial Cyclic Behav- Philadelphia, pp. 119–137.
ior in Two Precipitates Strengthened Alloys: Influence of the Loading Path and 关36兴 Bernstein, H. L., Grant, T. S., McClung, R. C., and Allen, J. M., 1993, “Pre-
Microstructure,” Mem. Etud. Sci. Rev. Metall., 88, pp. 441–451. diction of Thermal Mechanical Fatigue Life for Gas Turbine Blades in Electric
关30兴 Latif, A., Clavel, M., Ferney, V., and Saanouni, A., 1994, “On the Modeling of Power Generation,” Thermomechanical Fatigue Behavior of Materials, ASTM
Non-Proportional Cyclic Plasticity of Waspaloy,” ASME J. Eng. Mater. Tech- STP 1186, H. Sehitoglu, Ed., ASTM, West Conshohocken, PA, pp. 212–238.
nol., 116, pp. 35–44. 关37兴 Rémy, L., Bernard, H., Malpertu, J. L., and Rezai-Aria, F., 1993, “Fatigue Life
关31兴 Shrikanth, A., and Zabaras, N., 1999, “A Computational Model for the Finite Prediction Under Thermal-Mechanical Loading in a Nickel-Base Superalloy,”
Element Analysis of Thermoplasticty Coupled with Ductile Damage at Finite in Thermomechanical Fatigue Behavior of Materials, ASTM STP 1186, H.
Strains,” Int. J. Numer. Methods Eng., 45, pp. 1569–1605. Sehitoglu, Ed., ASTM, West Conshohocken, PA, pp. 3–16.
关32兴 Coffin, L. F., 1977, Fatigue at High Temperature, Fourth International Con- 关38兴 Miller, M. P., McDowell, D. L., and Oehmke, R. L. T., 1992, “A Creep-
ference on Fracture, Pergamon Press, Waterloo, Ont., Canada. Fatigue-Oxidation Microcrack Propagation Model for Thermomechanical Fa-
关33兴 Neu, R. W., and Sehitoglu, H., 1989, “Thermomechanical Fatigue, Oxidation tigue,” ASME J. Eng. Mater. Technol., 114共3兲, pp. 282–288.
and Creep: Part II Life Prediction,” Metall. Trans. A, 20A, pp. 1769–1783. 关39兴 McDowell, D. L., Antolovich, S. D., and Oehmke, R. L. T., 1992, “Mechanis-
关34兴 Ostergren, W. J., 1992, “A Damage Function and Associated Failure Equations tic Considerations for TMF Life Prediction of Nickel-Base Superalloys,” Nucl.
for Predicting Hold Time and Frequency Effects in Elevated Temperature, Eng. Des., 133, pp. 383–399.

336 / Vol. 127, JULY 2005 Transactions of the ASME

Downloaded 25 Apr 2011 to 132.170.11.38. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
View publication stats

You might also like