You are on page 1of 16

Materials Performance and Characterization

doi:10.1520/MPC20200167 / Vol. 10 / No. 2 / 2021 / available online at www.astm.org

Shreya Mukherjee,1 Kaustav Barat,2 Soumitra Tarafder,3 S. Sivaprasad,3 and


Sujoy Kumar Kar4

Creep-Fatigue Behavior of a Newly


Developed Ultra-Supercritical Steam
Turbine Grade Nickel-Based Superalloy,
HAYNES 282

Manuscript received October 28, Reference


2020; accepted for publication S. Mukherjee, K. Barat, S. Tarafder, S. Sivaprasad, and S. K. Kar, “Creep-Fatigue Behavior of a
February 1, 2021; published online
Newly Developed Ultra-Supercritical Steam Turbine Grade Nickel-Based Superalloy, HAYNES
April 22, 2021. Issue published
June 24, 2021. 282,” Materials Performance and Characterization 10, no. 2 (2021): 161–176. https://doi.org/
10.1520/MPC20200167
1
Department of Metallurgical and
Materials Engineering, Indian
Institute of Technology
Kharagpur, Kharagpur 721302,
ABSTRACT
India, https://orcid.org/0000- Most engineering components used in gas/steam turbines are exposed to a range of complex
0003-0898-9821 loading conditions resulting from startup and shutdown procedures. These loading conditions
2
Material Science Division, Council involve superimposition of time-dependent creep on cyclic fatigue and can be simulated by
of Scientific and Industrial properly designed high-temperature creep-fatigue tests. Creep-fatigue interaction is a func-
Research – National Aerospace tion of duration and position of dwell in the loading waveform, and the material microstructure.
Laboratories, HAL Airport Rd.,
Bangalore 560017, India,
The objective of this work is to investigate the creep-fatigue interaction response of a newly
https://orcid.org/0000-0003- developed γ 0 -strengthened wrought nickel-based superalloy (HAYNES 282), which has a po-
2959-0369 tential application in advanced ultra-supercritical steam turbines. Creep-fatigue tests are con-
3
Material Science and Technology ducted at 760°C with strain dwell either at tensile peak or compressive peak or at both tensile
Division, Council of Scientific and and compressive peak positions for different dwell times of 100 and 1,000 s. The test results are
Industrial Research – National analyzed with respect to evolutions of peak stress, stress amplitude, stress relaxation, hyste-
Metallurgical Laboratory, Burma
Mines, Jamshedpur 831007, India,
resis loop, inelastic strain energy density, and degree of softening. Degree of softening is found
https://orcid.org/0000-0001- to increase with dwell position at tensile, compressive, and both peaks in that order. Tests with
7721-0514 (S.T.) dwell at both tensile and compressive peak positions are found to be the most damaging,
4
Department of Metallurgical and showing the least life. Between tensile dwell and compressive dwell tests, interestingly, those
Materials Engineering, Indian with compressive dwell show a significantly reduced life. Increasing dwell time aggravates the
Institute of Technology damaging effect manifold. The mechanism of fracture at the end of life is illustrated with fracto-
Kharagpur, Kharagpur 721302,
India (Corresponding author),
graphic characterization.
e-mail: sujoy.kar@metal.iitkgp.ac.
in, https://orcid.org/0000-
0001-8470-7406

Copyright © 2021 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959 161
ASTM International is not responsible, as a body, for the statements and opinions expressed in this paper. ASTM International does not endorse
any &RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
products represented in this paper.
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
162 MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282

Keywords
creep-fatigue interactions, HAYNES 282, nickel-based superalloy, time-dependent fatigue, strain peak dwell, dwell time,
dwell position, fracture mode

Introduction
Nickel-based superalloys are used in components of gas and steam turbines (STs) because of its high temper-
ature strength, oxidation resistance, and excellent creep resistance. Recently, most of the thermally operated
power plants have a target to reduce CO2 emission.1 This is normally attempted by improving the plant effi-
ciency by enhancing steam temperature and pressure, which calls for materials that can withstand such en-
hanced operating conditions. HAYNES 282, a newly developed nickel-based superalloy is one such promising
material for use in hot gas path components of advanced ultra-supercritical (A-USC) STs.2 It is a γ 0 ((Ni3(Al,
Ti))-L12) strengthened wrought superalloy, which achieves an outstanding balance of properties at high tem-
perature (650°C–760°C).3
During the operations of turbomachinery, cyclic stress that is due to startup/shutdown cycle is superimposed
on the normal operating stress at high temperature.4,5 This superimposition results in a time period in every
loading-unloading cycle during which either the strain or stress remains constant. At high temperature, under
this type of cyclic loading with dwell time, the primary mode of damage in the material is through creep-fatigue
(C-F) interactions. In many instances, failure of several boiler components has been attributed to damage because
of C-F interaction.6 Therefore, understanding C-F damage as a function of various factors in a potential candidate
material for A-USC ST application is very important.
Introducing dwell time in the strain-controlled fatigue waveform is the most popular among different wave-
forms that can simulate the C-F interaction.7 Detailed study of effects of various factors that include temperature,
strain range, mean stress, frequency, dwell time, dwell position, etc. on C-F damage in various candidate materials
(i.e., nickel-based superalloys) have been made by various research groups.4–21 The time and position of dwell are
expected to have a significant implication in the hysteresis characteristics of the cyclic stress-strain response, and
hence would have impact of the life. Numerous researchers have studied the effect of dwell time on C-F response
of nickel-based superalloys.22–33 However, most of the studies were done with tensile dwell only,34–43 assuming it
to be the most damaging. Conversely, studies have shown that other dwell positions may be more damaging
depending on the material and test conditions.
The C-F interaction is a crossbreed type of deformation, in which time-dependent and time-independent
types of deformations are intertwined.30,39 A deconvoluting methodology is needed to separate the contribution
of each deformation process. Various attempts to model the C-F interaction include the following. In one of the
earlier works, Ellyin and Asada33 developed a model based on rate-dependent and rate-independent strain en-
ergies. Subsequently, various other models were developed to predict life under C-F conditions. Chen et al.4
assessed various life prediction methodologies, namely, linear damage summation method, strain range parti-
tioning method, and strain energy partitioning method. Wang et al.38 extended the energy-based model based on
strain energy exhaustion and accumulated damage using cycle by cycle concept. Halford, Hirschberg, and
Manson44 suggested that the concept of hybrid damage models stems from the energy and strain partitioning
approach to low cycle fatigue damage.
A few generalized understandings of the C-F damage summarized based on the literature survey are as
follows. C-F damage is more damaging than either pure creep or pure fatigue. This is due to the fact that during
strain dwell, stress relaxation causes increased inelastic strain, leading to lowering of life.25 Usually life reduces
with increasing dwell time. However, beyond a certain increase in dwell time, creep ductility sets in and life may
be improved at higher dwell time.8,22,42 In general, higher creep dominance to the C-F interaction causes a higher
tendency to the intergranular failure.
In regard to work on HAYNES 282, although a few studies on high-temperature low cycle fatigue as well as
fatigue crack growth were reported,45,46 literature on C-F interaction in this alloy is very limited.18 Brommesson,

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282 163

Ekh, and Persson18 studied the C-F interaction for this alloy with tensile only hold on waveforms with strain ratio
of R∈ = 0. They primarily focused on modeling the C-F data using Golos-Ellyin strain energy density fatigue
criterion to predict C-F life. Rozman, Kruzic, and Hawk46 worked primarily on the fatigue crack growth behavior
in this material in the temperature range of 550°C–750°C. Hence, a systematic experimental study of dwell-fatigue
in HAYNES 282 alloy was much required to understand the deformation mode as a function of both dwell time
and dwell position.
The present study aimed at systematically studying the effects of time and position of strain dwell on the
deformation behavior of HAYNES 282 alloy at 760°C. Tests were carried out at a strain rate (ε̇) of 1×10−4 s−1 with
a strain amplitude ðΔεt =2Þ of 0.4 %. Strain dwells of 100 s and 1,000 s were applied either at tensile peak or at
compressive peak or at both the peak strain positions. Strain fluctuations in actual operations are well below the
chosen ðΔεt =2Þ, and dwell durations were selected based on the practicality of testing in laboratory conditions.
Effects of various combinations of dwell time and position on evolutions of hysteresis loops, peak stresses, stress
amplitude, inelastic strain amplitude, stress relaxation during dwell, etc. were studied. Finally, the damage mode
of the present alloy under dwell fatigue conditions was investigated as a function of dwell time and position
through fractographic analysis.

Materials and Methods


MATERIAL DETAILS
A cast and forged block of ∼40 kg of HAYNES 282 material was received from GE Energy Schenectady via GE
GRC Bangalore. The nominal composition of this alloy determined by X-ray fluorescence is chromium 0.20,
cobalt 0.11, titanium 0.021, aluminum 0.0156, silicon 0.0019, carbon 0.007, boron 0.0014, and nickel balance;
cast and forged coupons were sectioned to 110 mm length rods. Heat treatment comprised first solutionizing at
1,120°C for 1 h and air cooling to room temperature (RT), followed by aging treatment at 760°C for 8 h and
furnace cooling to RT.

METALLOGRAPHY AND CHARACTERIZATION


After heat treatment, samples were grinded to remove ∼2.5-mm-thick surface layer to eliminate any effect of
oxygen ingress. The metallographic preparation of all the samples was done by paper grinding and polishing,
followed by polishing with a diamond paste of different sizes and finally by colloidal silica. Initial microstructures
of the heat-treated bars were characterized by optical microscope. Leica DM 2700 M was employed to capture
optical micrographs. Grain size was calculated by linear intercept method using MIPAR version 1.1. Transmission
electron microscopy (TEM) samples were prepared using the usual mechanical thinning process, punching of
3 mm disc, followed by twin jet polishing. TEM investigation was carried out in JEOL 2100F operated at 200 kV.
Precipitates size distribution were determined from TEM images using Adobe Photoshop CS V3 with Fovea pro
plug-ins.47 For measurement, 4–5 images of various locations were used. Each image consisted of average 50–60
precipitates. After C-F tests fracture surfaces from the failed specimen were observed under FEG SEI FESEM
(MERLIN) to study the fracture morphology.

C-F TESTING DETAILS


Threaded cylindrical samples were fabricated from heat-treated coupons with 7-mm gage diameter and 14-mm gage
length as shown in figure 1. All C-F tests were carried out according to ASTM E2714-13(2020), Standard Test
Method for Creep-Fatigue Testing,48 at a strain rate 1×10−4 s−1 and at 760°C using an INSTRON 8862 servo-electric
machine equipped with a high-temperature furnace of maximum temperature 1,000°C. Under fully reversed ðR∈ =
−1Þ uniaxial strain controlled mode CF tests were carried out in air as atmospheric condition at 760°C. Frequency is
calculated for each test based on the dwell time and position. For this study, 3 waveforms (i.e., standard peak dwell
tests) were chosen with two different dwell positions (tensile and compression) as shown in figure 2. Temperature
during the testing was controlled by MTS model temperature controller. Two external K-type thermocouples were

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
164 MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282

FIG. 1
Schematic of the C-F test
specimen (all dimensions
are in mm).

FIG. 2
Schematic
representation of the
loading waveform for
tensile, compressive, and
both dwell tests.

wrapped with the specimen to be tested for continuous monitoring of the test temperature. The temperature along
the gage length was controlled with an error of less than +/−5°C. Contact type extensometer made up of ceramic
material having gage length of 12.5 mm and travel +2.5 mm (tensile)/−1.25 mm (compressive) was calibrated using
customary standard and was attached to the gage portion of the specimen for measuring strain. All specimens were
tested until failure. Failure criteria was set to be 20 % of the load drop. Fracture surfaces of the fatigued specimens
were preserved in vacuum desiccators for microscopic observation.

Results and Discussions


MICROSTRUCTURAL CHARACTERIZATION
Optical images of as-received microstructures are shown in figure 3A. Pronounced annealing twins are observed
as parallel lines. ASTM grain size number is found to be 4.8. Heat-treated microstructure before testing is

FIG. 3
Micrographs of as-
received specimen: (A)
optical micrograph
showing grains,
annealing twins, and
carbides; (B) TEM BF
image showing γ 0
precipitates.

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282 165

characterized using TEM. Figure 3B shows a TEM bright field (BF) image of γ 0 size distribution at 760°C aged for
8 h. A unimodal distribution of γ 0 precipitates with a mean size of 60 ± 7 nm is observed.
Material’s response under strain-controlled symmetric cyclic loading with dwell time can be characterized by
evolutions of peak stress, stress amplitude, stress relaxation, and inelastic strain amplitude during cycling. Some
derived parameters, like the degree of softening, normalized cycle ratio (NCR), and hysteresis energy can also be
analyzed to characterize the C-F response of the material. These analyses are presented in the following.
Subsequently, fracture mode as a function of dwell time and mode is discussed based on the fractographs.

PEAK STRESS
Evolution of peak stress with cycle number is presented in figure 4 (only the absolute values are shown). Whereas
figure 4A–C show the plots of 100-s dwell at tensile peak, compressive peak, and both peaks, respectively,
figure 4D–F show the plots of 1,000-s dwell at positions in the same order, respectively. Peak stress during cycling
has been found to be affected by two things: firstly, the generic phenomena of initial cyclic hardening followed by
slight softening (because of accumulated damage) of the material, and secondly, the effect of stress relaxation. The
effect of stress relaxation during dwell at different positions are much more pronounced in the case of 1,000-s
dwell (fig. 4D–F), whereas this effect is negligible in the case of 100-s dwell (fig. 4A–C), in which case initial cyclic
hardening followed by slight softening or attaining of steady state value is observed. The effect of stress relaxation
(observed for 1,000-s dwell) is explained in more detail in the subsequent text.
Tensile dwell for 1,000 s (fig. 4D) causes (1) the compressive peak stress to increase continuously through a
major part of the life before reaching stable values and (2) the tensile peak stress to continuously go down beyond
a few initial cycles. During dwell at the tensile peak strain, stress relaxes. This stress relaxation ensures the starting
stress point for the reverse loading to advance significantly toward the compressive side. This helps the com-
pressive stress to reach to a higher (negative) value under the full reversal of the strain (when the compressive
peak strain is reached). This increase in the compressive peak stress continues through the cycling. Hence, the
compressive peak stress continuously increases during cycling for the tensile dwell case. Also, as the reloading
starts from a higher (more negative) value of compressive stress at the compressive peak strain position, the
tensile stress attained at the peak tensile strain position (under the full loading half cycle) is lower as compared
to the previous cycle. This continues through the cycling and causes a continuous decrease in the peak tensile
stress.
Compressive dwell for 1,000 s (fig. 4E) causes (1) the compressive peak stress to continuously go down and
(2) the tensile peak stress to slightly increase until about 20 cycles, beyond which it decreases slowly during further
cycling. Compressive dwell causes stress relaxation of compressive peak stress at every cycle, which causes its
continuous decrease in subsequent cycles. Because of relaxation of this compressive stress at the compressive peak
strain, the reverse loading starts from a stress point that is more toward the tensile side. This helps the tensile
stress to reach to a higher value when full loading (strain) half cycle is imposed so that the tensile peak strain is
reached. Hence, like the case of tensile dwell, for the case of compressive dwell, tensile peak stress should increase.
However, this effect of increase in the opposite peak stress during dwell at a particular peak strain is found to be
much less when compressive dwell is applied as compared to tensile dwell. This suggests that the degree of soft-
ening introduced in the material because of compressive dwell is more than that which is due to tensile dwell.
During dwell at a peak strain, along with stress relaxation that is responsible for the opposite peak stress to
increase, time-dependent inelastic strain is also introduced in the material, which softens the material.
In the case of 1,000-s dwell at both the peak positions (fig. 4F), the opposite effects of stress relaxation during
dwell at each position cancel out to a certain degree. In the case of both 1,000-s dwell (fig. 4F), beyond the first 3
cycles, tensile peak stress decreases and the compressive peak stress increases slowly to reach respective steady
state values. From the previous discussion, we have seen that stress relaxation during dwell at a specific position
causes the corresponding peak stress to decrease and the opposite peak stress to increase during cycling. We have
also seen that the effect of increase in the opposite peak stress is much more by the tensile dwell. This explains the
previously observed phenomena for 1,000-s dwell at both the peak positions.

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
166 MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282

FIG. 4 Evolution of tensile and compressive peak stresses with cycles (A)–(C) for 100-s dwell (A) at the tensile peak,
(B) at the compressive peak, (C) at both the peaks, and (D)–(F) for 1,000-s dwell (D) at the tensile peak, (E) at the
compressive peak, and (F) at both the peaks.

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282 167

STRESS AMPLITUDE
The evolution of stress amplitude (half of the stress range [peak tensile stress – peak compressive stress]) with
elapsed cycles is depicted in figure 5. For 100-s dwell tests, irrespective of dwell position, initial cyclic hardening
followed by softening and saturation is observed. In the case of 1,000-s dwell, a continuous softening behavior is
observed. This transition from cyclic hardening response to continuous softening by increasing dwell time has
also been reported for some other superalloys like Haynes 230, Haynes 188, etc.11,25,26 The stress amplitude is
higher in 100-s dwell compared to 1,000-s dwell time. Both dwell shows much more vigorous softening phe-
nomenon. Among the single position dwell tests, the compressive dwell shows higher softening rate (higher slope
in the curve in fig. 5).
As discussed earlier, the effect of stress relaxation for 100-s dwell tests is much less. The 1,000-s dwell intro-
duces significant inelastic strain in the material, causing continuous softening. Softening rate (slope of the curves)
that is due to strain dwell decreases in the following order: both dwell, compressive dwell, and the tensile dwell.

FIG. 5
Evolution of stress
amplitude in different
loading waveforms.

FIG. 6
Degree of softening
based on inelastic strain
range for various
waveforms.

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
168 MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282

DEGREE OF SOFTENING
Degree of softening (S) has been quantified here based on inelastic strain range as follows:

Δε1in − Δεstin
S=j j (1)
Δεstin

where Δεstin is the inelastic strain range at the steady (or N f =2) cycle and Δε1in is the inelastic strain range of the
first (or initial) cycle. Inelastic strain is obtained by subtracting the elastic strain from the total strain.
Figure 6 compares softening phenomenon of various loading schemes. The degree of softening for 100-s
dwell tests is much less. The variation that is due to the dwell position is also minimal for the 100-s dwell tests,

FIG. 7
Life and degree of
softening for different
waveforms.

FIG. 8 Evolution of stress relaxation with cycle for various loading waveforms: (A) tensile stress state (tensile dwell and
tensile part of the both dwell) and (B) compressive stress state (compressive dwell and compressive part of the
both dwell).

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282 169

only showing the maximum softening in case of the both dwell test. For 1000-s dwell, however, the degree of
softening is much higher and is sensitive to the dwell position. Both dwell shows the maximum degree of soft-
ening. Compressive dwell shows higher degree of softening than the tensile dwell as discussed before. It is found
that the degree of softening relates well with C-F life for various waveforms as shown in figure 7. The higher the
degree of softening, the lower is the life. Softening is due to the inelastic strain generation at the time of stress
relaxation during strain dwell.25

STRESS RELAXATION
Stress relaxation is an important phenomenon to understand C-F behavior. The stress relaxation is mainly the
increased inelastic component of strain during dwell time in a strain cycling test. The stress relaxation is

FIG. 9
Variations in stress
relaxation per cycle
(SRcycle) and life for
various waveforms.

FIG. 10
Comparison of NCR
values for different
loading waveforms
tested at 760°C.

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
170 MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282

dependent on temperature, strain range, and dwell time.36 Figure 8 shows the evolution of stress relaxation with
cycle number for the tensile stress state (fig. 8A) and the compressive stress state (fig. 8B). Both dwell shows
higher stress relaxation irrespective of the stress state as compared to the single dwell cases. Expectedly, 1,000-s
dwell shows higher stress relaxation than 100-s dwell for all the dwell positions.

FIG. 11 Evolution of hysteresis loop of stress versus inelastic strain for (A) tensile 100 s, (B) compressive 100 s,
(C) both 100 s, (D) tensile 1,000 s, (E) compressive 1,000 s, and (F) both 1,000 s.

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282 171

When the stress relaxations of all the cycles are averaged over the corresponding number of cycles to failure
(life), we get an estimate of effective stress relaxation per cycle (SRcycle). Figure 9 shows the variations in SRcycle
and life for various waveforms. It is found that this stress relaxation per cycle also relates well with the C-F life.
The higher the stress relaxation per cycle, the lower is the life.

DWELL SENSITIVITY BEHAVIOR


The performance of a material under dwell fatigue tests can be diagnosed by a parameter called NCR. It is the ratio
of C-F life to that of fatigue life without dwell. It indicates sensitivity to time-dependent fatigue tests. If value of
NCR is less than 1, then the material is sensitive to C-F interaction. The higher the value of NCR, the less dam-
aging it is.8
LCF tests with no dwell were carried out on the same material earlier under the same set of conditions (fully
reversed strain controlled cycle with 0.4 % strain amplitude at 760°C, at a strain rate of 1 × 10−4 s−1), which we
reported in Barat et al.49 The values of LCF lives with and without dwell were used to calculate the NCR values,
which are shown in figure 10 for various waveforms. There is a trend with dwell position for both the dwell times
considered here. The order of detrimental effects of C-F interaction, moving from the least damaging to the most
damaging, is the tensile dwell, the compressive dwell, and the both dwell. Dwell position has significant effect on
the amount of time-dependent inelastic strain introduced in the material and the amount of stress relaxation per
cycle, both of which weaken the material.

EVOLUTION OF HYSTERESIS LOOP/ INELASTIC STRAIN WITH CYCLES


Cyclic hardening/softening can also be visualized through the evolution of the hysteresis loops of stress versus
inelastic strain. In C-F tests, inelastic strain consists of plastic as well as creep strain. Figure 11 shows the evolution
of hysteresis loops for different waveforms. Whereas figure 11A–C show loop evolution for 100-s dwell for the
dwell positions of tensile, compressive, and both respectively, figure 11D–F show the same for the 1,000-s dwell in
the same order for dwell positions. It is observed that with cycling, hysteresis loops shift in the direction of strain
hold, either in the case of tensile or compressive. Tensile dwell makes the loops shift toward the tensile strain side,
whereas the compressive dwell shifts the loops toward the compressive side. The 1,000-s dwell shows a significant
amount of inelastic strain generation with respect to the initial loop. Both peak dwell shows significantly higher

FIG. 12
Inelastic strain energy
density and life for
different waveforms.

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
172 MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282

inelastic strain range as compared to the single peak dwell from the initial stage of the tests (compare the 2nd
cycle). Evolution shows an increase in inelastic strain and a decrease in peak stress, signifying softening of the
material.
One can measure the inelastic strain energy density (energy per unit volume) for a specific cycle by meas-
uring the area within a hysteresis loop of stress versus inelastic strain.45 It is also known as hysteresis energy and is
denoted by W in . The higher value of W in signifies a larger amount of energy absorbed by the material as damage

FIG. 13 SEM fractographs after C-F tests: (A, B) tensile dwell – 100 s, (C, D) tensile dwell -1,000 s, (E, F) compressive dwell
- 100 s, and (G, H) both dwell -100 s.

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282 173

during a cycle. It is found that waveforms showing a higher value of W in (measured for a cycle at half-life) have
lower life (fig. 12). Both dwell shows more inelastic strain energy density compared to single dwell, as also was
observed by Singh and Sahu.42

DETERMINATION OF FRACTURE MORPHOLOGY


In cyclic loading, failure occurs in three stages, namely crack initiation, crack propagation, and final instant
fracture. Figure 13 shows fractographs of 100-s dwell at different dwell positions (fig. 13A and 13B for tensile
peak, fig. 13E and 13F for compressive peak, fig. 13G and 13H for both peaks) and 1,000-s dwell at tensile peak
position (fig. 13C and 13D). In the fractographs, crack initiation sites and features like tear ridges secondary cracks
are marked. Study of the fractographs reveals the following about the fracture mode as a function of dwell position
and dwell time. Single dwell for 100 s at either tensile or compressive peaks shows full transgranular fracture,
whereas both dwell for 100 s (fig. 13H) shows intergranular fracture along with transgranular fracture. Increasing
dwell time from 100 to 1,000 s (fig. 13D) converts the fracture mode from fully transgranular to a predominantly
intergranular mixed mode fracture. Increasing dwell time causing more propensity to fail through intergranular
cracking mode was reported by other researchers also.35,39,50 The cause of intergranular cracking can be attributed
to the combined effect of planar slip and oxidation during dwell at elevated temperature.50

Conclusions
Complex C-F behavior of a newly developed ultra-supercritical ST grade material HAYNES 282 is studied. The
major conclusions from this study are summarized as follows:

• The effect of stress relaxation during dwell at different positions is much more pronounced in the case of
1,000-s dwell, whereas this effect is negligible in the case of 100-s dwell, in which case the initial cyclic
hardening followed by slight softening or attaining of steady state is observed. The 1,000-s dwell introduces
significant inelastic strain in the material, causing continuous softening. Softening effect of the strain dwell
increases in the following order: tensile dwell, compressive dwell, and both dwell.
• Both dwell shows the maximum degree of softening. Compressive dwell shows a higher degree of softening
than the tensile dwell. The higher the degree of softening, the lower the life is.
• Stress relaxation per cycle also relates well with C-F life. The higher the stress relaxation per cycle, the lower
the life is.
• Both dwell is found to be the most damaging as determined by life and NCR, caused mainly by higher amounts
of inelastic strain introduced in the material and stress relaxation during strain dwell for this condition.
• Inelastic strain energy density is highest for both dwell conditions irrespective of dwell time.
• Fractographic evidence shows the occurrence of transgranular fracture mode for single dwell either at ten-
sile or compressive peak. However, both dwell for 100 s shows intergranular fracture along with trans-
granular fracture. Increasing dwell time from 100 to 1,000 s converts the fracture mode from fully
transgranular to a predominantly intergranular mixed mode fracture.

ACKNOWLEDGMENTS
The raw/processed data required to reproduce these findings can be found in the Experimental and Results and
Discussion sections of this article. Raw cyclic stress strain data may be available on request. The authors would like to
acknowledge the support of Council of Scientific and Industrial Research-National Metallurgical Laboratory
Jamshedpur for carrying out the C-F tests, General Electric Schenectady for providing the material, and
Central Research Facility – Indian Institute of Technology Kharagpur for the characterization facility.

References
1. M. Yonemura and M. Mitsuhara, “Damage Mechanism of Nickel-Based Creep-Resistant Alloys Strengthened by the
Laves Phase at the Grain Boundary,” Philosophical Magazine 98, no. 36 (2018): 3247–3266, https://doi.org/10.1080/
14786435.2018.1524161

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
174 MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282

2. R. Viswanathan, J. F. Henry, J. Tanzosh, G. Stanko, J. Shingledecker, B. Vitalis, and R. Purgert, “U.S. Program on Materials
Technology for Ultra-Supercritical Coal Power Plants,” Journal of Materials Engineering and Performance 14 (June 2005):
281–292, https://doi.org/10.1361/10599490524039
3. L. M. Pike, “Development of a Fabricable Gamma-Prime (γ 0 ) Strengthened Superalloy,” in Proceedings of the
International Symposium on Superalloys (Pittsburgh, PA: TMS, 2008), 191–200.
4. L. J. Chen, Z. G. Wang, G. Yao, and J. F. Tian, “An Assessment of Three Creep–Fatigue Life Prediction Methods for
Nickel-Based Superalloy GH4049,” Fatigue & Fracture of Engineering Materials & Structures 23, no. 6 (June 2000):
509–519, https://doi.org/10.1046/j.1460-2695.2000.00311.x
5. B. Fournier, M. Sauzay, C. Caës, M. Noblecourt, M. Mottot, A. Bougault, V. Rabeau, and A. Pineau, “Creep-Fatigue-
Oxidation Interactions in a 9Cr–1Mo Martensitic Steel. Part II: Effect of Compressive Holding Period on Fatigue
Lifetime,” International Journal of Fatigue 30, no. 4 (2008): 663–676, https://doi.org/10.1016/j.ijfatigue.2007.05.008
6. S. Holdsworth, “Creep-Fatigue Failure Diagnosis,” Materials 8, no. 11 (2015): 7757–7769, https://doi.org/10.3390/
ma8115418
7. E. G. Ellison, “A Review of the Interaction of Creep and Fatigue,” Journal of Mechanical Engineering Science 11, no. 3
(1969): 318–339, https://doi.org/10.1243/JMES_JOUR_1969_011_039_02
8. B. Ding, W. Ren, J. Peng, Y. Zhong, and J. Yu, “Influence of Dwell Time on the Creep–Fatigue Behavior of a Directionally
Solidified Ni-Based Superalloy DZ445 at 850°C,” Materials Science and Engineering: A 725 (May 2018): 319–328, https://
doi.org/10.1016/j.msea.2018.04.038
9. C. Dong, H. Yu, Z. Jiao, F. Kong, and Y. Chen, “Low Cycle Fatigue, Creep and Creep-Fatigue Interaction Behavior of a
TiAl Alloy at High Temperatures,” Scripta Materialia 144 (February 2018): 60–63, https://doi.org/10.1016/j.scriptamat.
2017.09.016
10. C. Stöcker, M. Zimmermann, H.-J. Christ, Z.-L. Zhan, C. Cornet, L. G. Zhao, M. C. Hardy, and J. Tong, “Micro-
structural Characterisation and Constitutive Behaviour of Alloy RR1000 under Fatigue and Creep–Fatigue Loading
Conditions,” Materials Science and Engineering: A 518, nos. 1–2 (August 2009): 27–34, https://doi.org/10.1016/j.
msea.2009.04.055
11. S. Y. Lee, Y. L. Lu, P. K. Liaw, L. J. Chen, S. A. Thompson, J. W. Blust, P. F. Browning, A. K. Bhattacharya, J. M.
Aurrecoechea, and D. L. Klarstrom, “Tensile-Hold Low-Cycle-Fatigue Properties of Solid-Solution-Strengthened
Superalloys at Elevated Temperatures,” Materials Science and Engineering: A 504, nos. 1–2 (March 2009): 64–72,
https://doi.org/10.1016/j.msea.2008.10.030
12. A. Pineau and S. D. Antolovich, “High Temperature Fatigue of Nickel-Base Superalloys – A Review with Special Emphasis
on Deformation Modes and Oxidation,” Engineering Failure Analysis 16, no. 8 (2009): 2668–2697, https://doi.org/10.
1016/j.engfailanal.2009.01.010
13. X. Chen, Z. Yang, M. A. Sokolov, D. L. Erdman III, K. Mo, and J. F. Stubbins, “Effect of Creep and Oxidation on Reduced
Fatigue Life of Ni-Based Alloy 617 at 850 °C,” Journal of Nuclear Materials 444, nos. 1–3 (2014): 393–403, https://doi.org/
10.1016/j.jnucmat.2013.09.030
14. R.-Z. Wang, B. Chen, X.-C. Zhang, S.-T. Tu, J. Wang, and C.-C. Zhang, “The Effects of Inhomogeneous Microstructure
and Loading Waveform on Creep-Fatigue Behaviour in a Forged and Precipitation Hardened Nickel-Based Superalloy,”
International Journal of Fatigue 97 (April 2017): 190–201, https://doi.org/10.1016/j.ijfatigue.2017.01.002
15. F. Tahir, S. Dahire, and Y. Liu, “Image-Based Creep-Fatigue Damage Mechanism Investigation of Alloy 617 at 950 °C,”
Materials Science and Engineering: A 679 (January 2017): 391–400, https://doi.org/10.1016/j.msea.2016.10.050
16. Z. Y. Yu, X. Z. Wang, Z. F. Yue, and X. M. Wang, “Investigation on Microstructure Evolution and Fracture Morphology of
Single Crystal Nickel-Base Superalloys under Creep-Fatigue Interaction Loading,” Materials Science and Engineering: A
697 (June 2017): 126–131, https://doi.org/10.1016/j.msea.2017.05.018
17. S. X. Li and D. J. Smith, “High Temperature Fatigue‐Creep Behaviour of Single Crystal SRR99 Nickel Base Superalloys:
Part II—Fatigue-Creep Life Behaviour,” Fatigue & Fracture of Engineering Materials & Structures 18, no. 5 (1995):
631–643, https://doi.org/10.1111/j.1460-2695.1995.tb01423.x
18. R. Brommesson, M. Ekh, and C. Persson, “Experimental Observations and Modelling of Cyclic and Relaxation Behaviour
of the Ni-Based Superalloy Haynes 282,” International Journal of Fatigue 87 (June 2016): 180–191, https://doi.org/10.
1016/j.ijfatigue.2016.01.027
19. R.-Z. Wang, S.-P. Zhu, J. Wang, X.-C. Zhang, S.-T. Tu, and C.-C. Zhang, “High Temperature Fatigue and Creep-Fatigue
Behaviors in a Ni-Based Superalloy: Damage Mechanisms and Life Assessment,” International Journal of Fatigue 118
(January 2019): 8–21, https://doi.org/10.1016/j.ijfatigue.2018.05.008
20. K. Barat, S. Sivaprasad, S. Kar, and S. Tarafder, “Creep Fatigue Interaction under Different Test Variables: Mechanics and
Mechanisms,” Journal of Testing and Evaluation 46, no. 6 (2018): 2521–2539, https://doi.org/10.1520/JTE20160399
21. R. Ahmed, P. R. Barrett, M. Menon, and T. Hassan, “Thermo-Mechanical Low-Cycle Fatigue-Creep of Haynes 230,”
International Journal of Solids and Structures 126–127 (November 2017): 90–104, https://doi.org/10.1016/j.ijsolstr.
2017.07.033
22. B. Ding, W. Ren, K. Deng, H. Li, and Y. Liang, “An Abnormal Increase of Fatigue Life with Dwell Time during Creep-
Fatigue Deformation for Directionally Solidified Ni-Based Superalloy,” High Temperature Materials and Processes 37,
no. 3 (1984): 277–284, https://doi.org/10.1515/htmp-2016-0079
23. T. C. Totemeier and H. Tian, “Creep-Fatigue–Environment Interactions in INCONEL 617,” Materials Science and
Engineering: A 468–470 (November 2007): 81–87, https://doi.org/10.1016/j.msea.2006.10.170

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282 175

24. Y. L. Lu, L. J. Chen, P. K. Liaw, G. Y. Wang, C. R. Brooks, S. A. Thompson, J. W. Blust, et al., “Effects of Temperature and
Hold Time on Creep-Fatigue Crack-Growth Behavior of HAYNES® 230® Alloy,” Materials Science and Engineering: A
429, nos. 1–2 (2006): 1–10, https://doi.org/10.1016/j.msea.2005.07.039
25. L. J. Chen, P. K. Liaw, Y. H. He, M. L. Benson, J. W. Blust, P. F. Browning, R. R. Seeley, and D. L. Klarstrom, “Tensile Hold
Low-Cycle Fatigue Behavior of Cobalt-Based HAYNES® 188 Superalloy,” Scripta Materialia 44, no. 6 (2001): 859–865,
https://doi.org/10.1016/S1359-6462(00)00702-8
26. Y. L. Lu, L. J. Chen, G. Y. Wang, M. L. Benson, P. K. Liaw, S. A. Thompson, J. W. Blust, et al., “Hold Time Effects on Low
Cycle Fatigue Behavior of HAYNES 230® Superalloy at High Temperatures,” Materials Science and Engineering: A 409,
nos. 1–2 (2005): 282–291, https://doi.org/10.1016/j.msea.2005.05.120
27. J. Zrník, J. Semeňák, V. Vrchovisnký, and P. Wangyao, “Influence of Hold Period on Creep–Fatigue Deformation
Behaviour of Nickel Base Superalloy,” Materials Science and Engineering: A 319–321 (December 2001): 637–642,
https://doi.org/10.1016/S0921-5093(01)01030-9
28. J. M. Silva, R. A. Cláudio, C. M. Branco, and J. M. Ferreira, “Creep-Fatigue Behavior of a New Generation Ni-Base
Superalloy for Aeroengine Usage,” Procedia Engineering 2, no. 1 (2010): 1865–1875, https://doi.org/10.1016/j.proeng.
2010.03.201
29. D. Hu and R. Wang, “Experimental Study on Creep–Fatigue Interaction Behavior of GH4133B Superalloy,” Materials
Science and Engineering: A 515, nos. 1–2 (2009): 183–189, https://doi.org/10.1016/j.msea.2009.02.049
30. T. Billot, P. Villechaise, M. Jouiad, and J. Mendez, “Creep–Fatigue Behavior at High Temperature of a UDIMET 720
Nickel-Base Superalloy,” International Journal of Fatigue 32, no. 5 (2010): 824–829, https://doi.org/10.1016/j.ijfatigue.
2009.07.003
31. K. Barat, S. Sivaprasad, S. K. Kar, and S. Tarafder, “Low-Cycle Fatigue of IN 718: Effect of Waveform,” Fatigue & Fracture
of Engineering Materials & Structures 42, no. 12 (2019): 2823–2843, https://doi.org/10.1111/ffe.13127
32. S. X. Li and D. J. Smith, “High Temperature Fatigue‐Creep Behaviour of Single Crystal SRR90 Nickel Base Superalloys:
Part 1—Cyclic Mechanical Response,” Fatigue & Fracture of Engineering Materials & Structures 18, no. 5 (1995):
617–629, https://doi.org/10.1111/j.1460-2695.1995.tb01422.x
33. F. Ellyin and Y. Asada, “Time-Dependent Fatigue Failure: The Creep-Fatigue Interaction,” International Journal of
Fatigue 13, no. 2 (1991): 157–164, https://doi.org/10.1016/0142-1123(91)90008-M
34. W. O. Ngala and H. J. Maier, “Creep-Fatigue Interaction of the ODS Superalloy PM 1000,” Materials Science and
Engineering: A 510–511 (2009): 429–433, https://doi.org/10.1016/j.msea.2008.06.056
35. D. Shi, J. Liu, X. Yang, H. Qi, and J. Wang, “Experimental Investigation on Low Cycle Fatigue and Creep–Fatigue
Interaction of DZ125 in Different Dwell Time at Elevated Temperatures,” Materials Science and Engineering: A 528,
no. 1 (2010): 233–238, https://doi.org/10.1016/j.msea.2010.08.089
36. D. Hu, Q. Ma, L. Shang, Y. Gao, and R. Wang, “Creep-Fatigue Behavior of Turbine Disc of Superalloy GH720Li at 650 °C
and Probabilistic Creep-Fatigue Modeling,” Materials Science and Engineering: A 670 (July 2016): 17–25, https://doi.org/
10.1016/j.msea.2016.05.117
37. R. K. Rai, J. K. Sahu, S. K. Das, N. Paulose, and C. Fernando, “Creep-Fatigue Deformation Micromechanisms of a
Directionally Solidified Nickel-Base Superalloy at 850°C,” Fatigue & Fracture of Engineering Materials & Structures
43, no. 1 (2019): 51–62, https://doi.org/10.1111/ffe.13028
38. R.-Z. Wang, X.-C. Zhang, J.-G. Gong, X.-M. Zhu, S.-T. Tu, and C.-C. Zhang, “Creep-Fatigue Life Prediction and
Interaction Diagram in Nickel-Based GH4169 Superalloy at 650 °C Based on Cycle-by-Cycle Concept,” International
Journal of Fatigue 97 (April 2017): 114–123, https://doi.org/10.1016/j.ijfatigue.2016.11.021
39. X. Chen, Z. Yang, M. A. Sokolov, D. L. Erdman III, K. Mo, and J. F. Stubbins, “Low Cycle Fatigue and Creep-Fatigue
Behavior of Ni-Based Alloy 230 at 850 °C,” Materials Science and Engineering: A 563 (February 2013): 152–162, https://
doi.org/10.1016/j.msea.2012.11.063
40. B. Fournier, M. Sauzay, C. Caës, M. Noblecourt, M. Mottot, L. Allais, I. Tournie, and A. Pineau, “Creep-Fatigue
Interaction in a 9 Pct Cr-1 Pct Mo Martensitic Steel: Part I. Mechanical Test Results,” Metallurgical and Materials
Transaction A 40, no. 2 (February 2009): 321–329, https://doi.org/10.1007/s11661-008-9686-z
41. M. Okazaki and Y. Yamazaki, “Creep-Fatigue Small Crack Propagation in a Single Crystal Ni-Base Superalloy, CMSX-2:
Microstructural Influences and Environmental Effects,” International Journal of Fatigue 21, Supplement 1 (1999):
S79–S86, https://doi.org/10.1016/S0142-1123(99)00058-4
42. R. K. Singh and J. K. Sahu, “Influence of Hold Type and Duration on Cyclic Deformation Behaviour of IN740 H,”
Materials Science and Technology 35, no. 4 (2019): 488–499, https://doi.org/10.1080/02670836.2019.1570662
43. R. T. Dewa, J. H. Park, S. J. Kim, and S. Y. Lee, “High-Temperature Creep-Fatigue Behavior of Alloy 617,” Metals 8, no. 2
(2018): 103, https://doi.org/10.3390/met8020103
44. G. R. Halford, M. H. Hirschberg, and S. S. Manson, “Creep-Fatigue Analysis by Strain-Range Partitioning” (paper pre-
sentation, National Pressure Vessel and Piping Conference, San Francisco, CA, May 1, 1971).
45. S. Mukherjee, K. Barat, S. Sivaprasad, S. Tarafder, and S. K. Kar, “Elevated Temperature Low Cycle Fatigue Behaviour of
Haynes 282 and Its Correlation with Microstructure – Effect of Ageing Conditions,” Materials Science and Engineering: A
762 (August 2019): 138073, https://doi.org/10.1016/j.msea.2019.138073
46. K. A. Rozman, J. J. Kruzic, and J. A. Hawk, “Fatigue Crack Growth Behavior of Nickel-Base Superalloy Haynes 282 at
550-750 °C,” Journal of Materials Engineering and Performance 24, no. 8 (August 2015): 2841–2846, https://doi.org/10.
1007/s11665-015-1588-9

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG
176 MUKHERJEE ET AL. ON CREEP-FATIGUE BEHAVIOR OF HAYNES 282

47. K. Barat, M. Ghosh, S. Sivaprasad, S. K. Kar, and S. Tarafder, “High-Temperature Low-Cycle Fatigue Behavior in
HAYNES 282: Influence of Initial Microstructure,” Metallurgical and Materials Transactions A: Physical Metallurgy
and Materials Science 49, no. 10 (July 2018): 5211–5226, https://doi.org/10.1007/s11661-018-4760-7
48. Standard Test Method for Creep-Fatigue Testing, ASTM E2714–13(2020) (West Conshohocken, PA: ASTM International,
approved May 1, 2020), https://doi.org/10.1520/E2714
49. K. Barat, M. Ghosh, S. Sivaprasad, S. K. Kar, and S. Tarafder, “High-Temperature Low-Cycle Fatigue Behavior in
HAYNES 282: Influence of Initial Microstructure,” Metallurgical and Materials Transactions A 49, no. 10 (2018):
5211–5226, https://doi.org/10.1007/s11661-018-4760-7
50. K. Prasad, R. Sarkar, P. Ghosal, and V. Kumar, “Simultaneous Creep–Fatigue Damage Accumulation of Forged Turbine
Disc of IN 718 Superalloy,” Materials Science and Engineering: A 572 (June 2013): 1–7, https://doi.org/10.1016/j.msea.
2013.02.003

&RS\ULJKWE\$670,QW
O DOOULJKWVUHVHUYHG 7KX1RY*07
Materials Performance and Characterization
'RZQORDGHGSULQWHGE\
1$/ 1$/ SXUVXDQWWR/LFHQVH$JUHHPHQW1RIXUWKHUUHSURGXFWLRQVDXWKRUL]HG

You might also like