You are on page 1of 18

International Journal of Fatigue 158 (2022) 106774

Contents lists available at ScienceDirect

International Journal of Fatigue


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

Multiaxial fatigue behaviour and damage mechanisms of P92 steel under


various strain amplitudes and strain ratios at high temperature
Ning Gao a, b, Wei Zhang a, b, *, Peng Yin a, b, Fei Liang a, b, Guodong Zhang c, Xianxi Xia c,
Yanfen Zhao c, Changyu Zhou a, b, *
a
School of Mechanical and Power Engineering, Nanjing Tech University, Nanjing 211816, China
b
Jiangsu Key Lab of Design and Manufacture of Extreme Pressure Equipment, Nanjing 211816, China
c
Suzhou Nuclear Power Research Institute, Suzhou 215004, China

A R T I C L E I N F O A B S T R A C T

Keywords: High temperature multiaxial fatigue tests at different strain amplitudes and strain ratios were performed at
Multiaxial fatigue 600 ◦ C. Dynamic strain ageing (DSA) analysis and microstructure characterization were used to analyze the
Strain amplitude cyclic deformation and damage mechanisms. Results reveal that the increases in strain amplitude and strain ratio
Strain ratio
accelerate the cyclic softening. Moreover, the DSA behaviour shows a distinct dependence on the strain
Cyclic deformation
amplitude, strain ratio and loading direction. Furthermore, the increase in strain amplitude promotes subgrains
Damage mechanism
growth and dislocation structure transformation. Particularly, the increase in strain ratio promotes the occur­
rence of planar slip of dislocation.

1. Introduction temperature. Nakamura et al. [10] studied the effects of phase angle on
multiaxial crack behaviour and fatigue life of Ti-6Al-4 V and found that
In the past decades, ultra-supercritical units have been widely the fatigue life of non-proportional specimen was about 1/10 of that of
developed and used in order to improve power generation efficiency, proportional specimen, and the number of cracks was about 10 times of
aiming to save resources and protect the environment [1-3]. The main that under proportional loading. Xu et al. [11] and Mei et al. [12]
steam temperature in the ultra-supercritical unit has reached or even considered that the loading path is the main factor affecting the multi­
exceeded 600 ◦ C, and the steam pressure has been higher than 31 MPa. axial fatigue performance. Among different paths, the fatigue perfor­
The critical operation conditions put forward stringent requirements for mance of the material under circular path loading is the worst. Shamsaei
the fatigue strength, creep strength, and corrosion resistance of gener­ et al. [13] investigated the effect of material structure and microstruc­
ator materials under high temperature and high pressure [4,5]. Owing to ture hardness on cyclic hardening of 1050 steel and 304 L stainless steel,
excellent properties at high temperature, P92 steel has been widely used and found that the increase of microstructure hardness leads to the
in manufacturing ultra-supercritical units [6]. During the service of decrease of cyclic hardening. In addition, the effect of loading sequence
ultra-supercritical units, the high-temperature components often suffer on the multiaxial fatigue behaviour of titanium was also studied [14].
from the interaction of complex high-temperature loading and different Colin et al.[15] studied the cyclic deformation of 304 L steel and 7075-
types of mechanical loading due to frequent start-ups and shutdowns. T6 aluminium under various loading amplitudes, and found that the
For complex structures, the stress state is not simple uniaxial stress, but a direction of overload affected the fatigue life. Ince et al.[16] studied the
multiaxial stress state is present [7-9]. Therefore, the multiaxial fatigue effects of plastic and elastic strains on multiaxial fatigue damage. They
behaviour of P92 steel at a high temperature is deserved more in­ believed that the axial plastic strain is conducive to opening the crack,
vestigations not only from the point of view of design but also from the and the torsional plastic strain is conducive to forming the core of the
damage evaluation. microcracks. On the other hand, the normal elastic strain helped to open
Recently, more and more researches have been focusing on the the crack and accelerated the crack growth. Ma et al. [17] carried out
multiaxial low cycle fatigue behaviours of materials. Generally, these multiaxial low cycle fatigue tests on TA2 under various multiaxial strain
studies were focused on the multiaxial fatigue behaviours at room ratios, and found that the increase of multiaxial strain ratio promoted

* Corresponding authors at: School of Mechanical and Power Engineering, Nanjing Tech University, Nanjing 211816, China.
E-mail addresses: zhang_wei@njtech.edu.cn (W. Zhang), changyu_zhou@163.com (C. Zhou).

https://doi.org/10.1016/j.ijfatigue.2022.106774
Received 30 November 2021; Received in revised form 18 January 2022; Accepted 27 January 2022
Available online 1 February 2022
0142-1123/© 2022 Elsevier Ltd. All rights reserved.
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 1. Microstructure of the as-received P92 steel: (a) OM graph, (b) TEM graph, and (c) schematic representation of the microstructure.

Fig. 2. Sizes of the specimen (mm).

Fig. 3. (a) Loading path and (b) loading waveform of the axial-torsion fatigue tests.

2
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

properties of duplex stainless steel (CF8M). Li et al. [26] performed an


experimental study on axial-torsion thermomechanical fatigue of nickel-
base alloys within the temperature range of 360 ◦ C to 650 ◦ C and found
that the fatigue life depended on the combined action of fatigue, creep,
and oxidation damage. To date, there are also some reports [7,27,28] on
the multiaxial properties of P92 steel. Nevertheless, due to the limitation
of test equipment, most of the studies can only apply axial loading to the
specimen. They cut a notch along the circumferential direction of the
cylindrical solid specimen, and obtained different multiaxial stress states
by changing the size of the notch. However, the notched specimen
cannot obtain the torsional stress state. Therefore, the effect of torsional
deformation on the material properties cannot be clarified through
notched specimens. To the best of the knowledge of the authors, re­
searches about the multiaxial fatigue behaviour of P92 steel under axial-
torsion loading at high temperature are still limited. Especially, re­
searches on the microscopic damage mechanism of P92 steel under
axial-torsion loading are rarely reported. Therefore, this work is devoted
to investigating the effects of both strain amplitude and strain ratio on
the cyclic deformation and damage mechanism of P92 steel at elevated
temperatures.
In the present work, multiaxial fatigue tests at different strain am­
plitudes (±0.2%, ±0.3%, ±0.4%, ±0.6%) and strain ratios (0, √3/10,
√3/5, 2√3/5, 3√3/5) are performed at an elevated temperature of
600 ◦ C. Afterwards, the cyclic stress–strain response and fatigue life of
the material under multiaxial fatigue loading are evaluated. Finally, the
fracture mechanism of the material is clarified by scanning electron
Fig. 4. Schematic diagram of the test device.
microscope (SEM), and the microstructures are observed by trans­
mission electron microscope (TEM) to clarify the damage mechanism of
the initial cyclic hardening in the axial direction. the material.
Based on the study of multiaxial fatigue at room temperature, a series
of studies on the multiaxial fatigue properties of materials at high 2. Material and methods
temperatures were carried out. Inoue et al. [18-21] systematically
studied the multiaxial fatigue behaviour of 2.25Cr-1Mo steel under The material studied in this work is P92 steel in the form of a pipe,
seven different loading conditions at 600 ◦ C. Shang et al. [22] conducted which has a diameter and wall thickness of 230 mm and 90 mm,
an experimental work on the multiaxial fatigue properties of nickel-base respectively. The chemical compositions are listed in wt.%: 0.012C;
alloy GH4169 at 650 ◦ C. It was shown that the strain loading path, 0.191Si; 0.418 Mn; 8.83 Cr; 0.462 Mo; 0.195 V; 0.0071 Ti; 0.0037B;
loading path sequence and loading parameters affected the cyclic 0.339 Ni; 0.052 Al; 0.0466 Cu; 0.0745 Nb; 0.048 N; 1.60 W; 0.0038P.
hardening or cyclic softening behaviours. Based on the critical plane Fig. 1(a) and (b) shows the microstructure of the material observed by
method, Shang et al. [23] defined a damage parameter to characterize optical microscope (OM) and TEM. The schematic representation of the
the multiaxial fatigue damage at high temperatures. Wang et al.[24] microstructure is presented in Fig. 1(c). As it can be seen that there are
further proposed a modified von Mises stress to calculate high- some points with high precipitate density, which are three-phase points
temperature multiaxial creep damage. Kwon et al. [25] focused on the formed due to the contact between the prior austenite grain boundaries
effect of material degradation on the axial-torsion multiaxial fatigue (PAGBs). Martensitic laths in different directions are distributed in the

Table 1
Detailed tests conditions for the axial-torsion fatigue tests of P92 steel.
Test No. Temperature Multiaxial strain ratio Axial strain amplitude Maximum torsional strain amplitude Axial strain rate Fatigue life
(◦ C) (%) (%) (s¡1) (N)

IF02600 600 0 ±0.2 / 1 × 10-3 6050


IF03600 600 0 ±0.3 / 1 × 10-3 894
IF04600 600 0 ±0.4 / 1 × 10-3 446
IF06600 600 0 ±0.6 / 1 × 10-3 107
TF02600√3/10 600 √3/10 ±0.2 ±0.035 1 × 10-3 4388
TF03600√3/10 600 √3/10 ±0.3 ±0.052 1 × 10-3 806
TF04600√3/10 600 √3/10 ±0.4 ±0.069 1 × 10-3 445
TF06600√3/10 600 √3/10 ±0.6 ±0.104 1 × 10-3 235
TF02600√3/5 600 √3/5 ±0.2 ±0.070 1 × 10-3 3225
TF03600√3/5 600 √3/5 ±0.3 ±0.104 1 × 10-3 629
TF04600√3/5 600 √3/5 ±0.4 ±0.138 1 × 10-3 440
TF06600√3/5 600 √3/5 ±0.6 ±0.208 1 × 10-3 173
TF026002√3/5 600 2√3/5 ±0.2 ±0.140 1 × 10-3 2535
TF036002√3/5 600 2√3/5 ±0.3 ±0.208 1 × 10-3 584
TF046002√3/5 600 2√3/5 ±0.4 ±0.276 1 × 10-3 372
TF066002√3/5 600 2√3/5 ±0.6 ±0.416 1 × 10-3 183
TF026003√3/5 600 3√3/5 ±0.2 ±0.210 1 × 10-3 1248
TF036003√3/5 600 3√3/5 ±0.3 ±0.312 1 × 10-3 545
TF046003√3/5 600 3√3/5 ±0.4 ±0.414 1 × 10-3 367
TF066003√3/5 600 3√3/5 ±0.6 ±0.624 1 × 10-3 183

3
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 5. Method of the preparation of the samples for SEM and TEM observations.

blocks of PAGBs, which is also a typical feature of P92 steel [29]. electrochemically polished with ethanol electrolyte containing 10%
Moreover, precipitates of M23C6 are distributed along the martensite perchloric acid to obtain the observable area. The microstructure
lath boundaries, and there are also some nano-type MX precipitates observation is conducted in an FEI Tecnai G2 F30 with a working voltage
distributed inside the lath [30], as shown in Fig. 1(b). In addition, some of 300 kV.
dislocations are observed in the martensite laths, where some disloca­
tions connect with each other. Therefore, some dislocation networks 3. Results
inside the laths are observed. The hierarchical microstructure contrib­
utes to the good properties of P92 steel under high temperatures. 3.1. Cyclic deformation response
To investigate the axial-torsion fatigue behaviour of P92 steel, a
hollow tubular specimen (Fig. 2) is designed according to the ASTM 3.1.1. Effect of strain amplitude
E2207 standard [31]. The gauge length, the outer diameter and the wall Fig. 6 shows the variation of peak tensile stress with respect to the
thickness of the specimen are 30 mm, 10 mm and 1 mm, respectively. number of fatigue cycles at different strain amplitudes. As it can be
Before testing, the outer and inner surfaces of all specimens are polished observed from the figure, the axial peak tensile stress of P92 steel shows
to avoid the influence of scratches on the initiation of fatigue cracks, a continuous softening behaviour under different loading conditions,
where the average surface roughness of the outer and inner sides are which is consistent with that of P92 steel under uniaxial loading [32,33].
limited within Ra 0.2 and Ra 0.4, respectively. Moreover, the increasing strain amplitude promotes the increase of peak
The axial-torsion fatigue test is performed in strain control mode. tensile stress. The torsional peak stress response is similar to the axial
During fatigue tests, the axial strain amplitude is set as ± 0.2%, ±0.3%, peak stress response, which also shows a continuous softening behaviour
±0.4% and ± 0.6%. A constant strain rate of 1 × 10-3 s− 1 is adopted. The in the whole fatigue process. Furthermore, the increasing strain ampli­
multiaxial strain ratio (λ) represents the ratio of the axial strain ampli­ tude induces the increase of torsional peak stress as well. Nevertheless,
tude to the torsional strain amplitude is set as 0, √3/10, √3/5, 2√3/5 the torsional peak stress is much lower than the axial peak stress. This
and 3√3/5. Fig. 3 shows the loading path and loading waveform of the may be because the plastic deformation, caused by the torsional
tests. The strain is measured using an axial-torsion extensometer pro­ component, is very small compared with that caused by the axial
duced by MTS (632.68F-08). Due to the limitation of the measurement component, as shown in Fig. 6(f). To quantitatively characterize the
range of the extensometer (the maximum torsion angle is ± 5◦ ), the cyclic softening behaviour, the curves of the axial and torsional stress
maximum strain ratio is set as 3√3/5. All the fatigue tests are carried response are fitted to determine the softening rate, as illustrated in
out at 600 ◦ C in an electro-hydraulic servo fatigue test system (MTS Fig. 7. It can be noted that the softening rate of the axial direction is
809). The temperature of the specimen is measured and controlled in much greater than that of the torsional direction, and the increase of
real-time by a K-type thermocouple, which is tied on the outer surface of strain amplitude accelerates the softening in both axial and torsional
the gauge section. When the temperature reaches 600 ◦ C, the specimen direction, which is detrimental to the fatigue life [34].
is held for 30 min to ensure that the temperature distributes uniformly Fig. 8 presents the stress–strain hysteresis loops along the axial and
between the outer and inner sides before fatigue tests, and the temper­ torsional directions of the half-life cycle under different strain ampli­
ature deviation is controlled within ± 2 ◦ C. Fig. 4 shows the detailed test tudes. It is noted that both the plastic strain energies (area of the hys­
device and Table 1 lists the detailed test conditions. teresis loop) in axial and torsional directions increase with increasing
After fatigue failure, the fracture characteristics of the failure spec­ strain amplitude. For axial stress, the peak stress with a strain amplitude
imens are observed by SEM and the microstructures of the failure of 0.2% is the smallest compared with other strain amplitudes. This is
specimens are observed by TEM, aiming to investigate the fracture and because when the strain amplitude is 0.2%, the plastic strain energy is
microstructural damage mechanisms. Fig. 5 presents the method of the very small and its softening rate is low. Thus the stress drop is slow, as
preparation of the samples for SEM and TEM observations. One part is shown in Fig. 6 and Fig. 8. Moreover, when comparing the cyclic
cut near the fracture location for SEM observation. Meanwhile, a thin stress–strain hysteresis loops in the axial and torsional direction, the
sheet, with a thickness of 0.3 mm along the inner wall of the specimen, is torsional stress is much smaller than the axial stress although the strain
cut for the TEM observation. These TEM samples are firstly mechanically is similar (Fig. 8(e1) and Fig. 8(e2)). Furthermore, stress serrations, a
polished to 80 μm with metallographic sandpaper, and then typical feature of dynamic strain ageing (DSA) [35], are also observed in

4
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 6. Variation of peak tensile stress of P92 steel under different strain amplitudes for various multiaxial strain ratios: (a) λ = 0, (b) λ=√3/10, (c) λ=√3/5, (d) λ =
2√3/5, (e) λ = 3√3/5 and (f) plastic strain at the half-life cycle.

the stress–strain hysteresis loops. DSA is mainly the result of the pinning and depinning of dislocations [41]. The stress drop is the difference
effect of solute atoms and moving dislocations [36,37]. In addition, between the highest stress and the lowest stress in the stress serration.
when the specimen is in a tensile state, the DSA phenomenon mainly The maximum value of the stress drop represents the locking strength of
occurs in the axial direction, but when the specimen is in a compressive the solute atomic air mass to the moving dislocation [42]. The value of
state, the DSA phenomenon mainly occurs in the torsional direction. stress drop and the number of stress serration can reflect the activity of
Previous studies [38-40] have shown that DSA affects fatigue properties DSA. When the value of stress drop is high and the number of stress
by reducing ductility and plasticity. To analyze the change of the DSA serrations is large, the DSA activity is strong. The number of stress ser­
effect, the number of stress serrations and maximum stress drop are rations and maximum stress drop under different strain amplitudes are
quantitatively analyzed. On the hysteresis loop of the half life cycle, the shown in Fig. 9(a) and (b), respectively. The results show that with
sum of the number of axial and torsional serrations is used as the number increasing strain amplitude, the number of stress serrations increases
of serrations. The serrations number has been proved to be related to the and shows an obvious accelerating trend, indicating the activity of DSA
microstructure evolution of materials, especially the periodic pinning increases. The maximum stress drop also has an obvious increasing trend

5
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 7. Variation of softening rate under different strain amplitudes: (a) axial softening rate, (b) torsional softening rate.

Fig. 8. Stress–strain hysteresis loops of the half-life cycle under different strain amplitudes of various strain ratios: (a) λ = 0, (b) λ=√3/10, (c) λ=√3/5, (d) λ =
2√3/5 and (e) λ = 3√3/5.

between the strain amplitude of 0.2% to 0.4%. However, the increasing life decreases with increasing axial strain amplitude. When the strain
trend of the maximum stress drop slows down or even decreases after amplitude is less than 0.4%, the fatigue life decreases rapidly with
strain amplitude exceeds 0.4%. This indicates that when the strain increasing strain amplitude, but when the strain amplitude is greater
amplitude exceeds 0.4%, the obstacle of dislocation movement is than 0.4%, the fatigue life decreases slowly. Researches on the uniaxial
weakened, which will be explained from the microscopic point of view fatigue behaviour of P92 steel has shown that as the applied strain
in the following sections. amplitude exceeds 0.4%, even if the applied strain amplitude increases,
Fig. 10 presents the evolution of fatigue life versus axial strain the change of fatigue life is not obvious [43]. From Fig. 10, we can also
amplitude. It can be observed that at different strain ratios, the fatigue obtain a similar conclusion as that under uniaxial fatigue loading,

6
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 8. (continued).

suggesting that the damage caused by the increase of strain amplitude sensitive to the change of the strain ratio. It is observed in Fig. 12(a) that
reaches saturation. In other words, the introduction of torsional stress/ when the strain ratio increases from 0 to √3/10, the axial softening rate
strain has little effect on the damage saturation behaviour at high strain increases rapidly. Because its loading state just changes from uniaxial to
amplitude. multiaxial loading, the introduction of the torsional stress is the major
reason for the change of the axial softening rate. When the multiaxial
3.1.2. Effect of strain ratio strain ratio is between √3/10 and √3/5, the elastic deformation ac­
Fig. 11 shows the evolution of peak tensile stress versus the number counts for a small part in the torsional component, and the plastic
of fatigue cycles at different strain ratios. Even if the axial strain deformation increases considerably, which promotes the transformation
amplitude remains unchanged, the axial peak tensile stress still alters of the microstructure. Therefore, the increasing trend of the axial soft­
due to the different strain ratios. It was found that the axial peak tensile ening rate slows down or even decreases. When the multiaxial strain is
stress decreases with increasing strain ratio, which is more obvious at √3/5 to 3√3/5, the plastic strain increases in the torsional direction,
the strain amplitude of 0.4% and 0.6%. However, torsional peak stress is and the axial softening rate increases slowly. Fig. 12(b) shows the
enhanced by increasing the strain ratio. The softening rate is also softening rate in the torsional direction. Different from the axial

7
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 9. (a) Number of stress serrations and (b) variation of maximum stress drop with strain amplitude.

significant degradation in fatigue life. The harmful effect of strain ratio


on the fatigue life is apparently degraded by increasing axial strain
amplitude. Particularly, the fatigue life keeps almost constant irre­
spective of the strain ratio at a strain amplitude of 0.6%. The results
indicate that the damage caused by torsional loading is relatively
smaller than that by axial loading.

3.2. Fracture features characterization

3.2.1. Fracture behaviour at different strain amplitudes


To identify the fracture mechanisms under various loading condi­
tions, four representative specimens with strain amplitudes of 0.2%,
0.3%, 0.4%, and 0.6% for a strain ratio equal to √3/5 were selected for
fracture morphology analysis, as shown in Fig. 16. It is observed that
there are multiple crack source regions on the inner and outer walls of
the specimen, where most of the main cracks initiate from the inner
surface and propagate to the outer surface, while a few cracks originate
Fig. 10. Relationship between fatigue life and axial strain amplitude. from the outer surface and extend to the inner wall. Finally, the two
cracks intersected near the outer wall, forming the fracture morphology
direction, the torsional softening rate shows an increasing trend, which characteristics of mainly cracking on the inner surface and supple­
is also consistent with the increase of plastic strain shown in Fig. 6(f). mented by cracking on the outer surface. With increasing strain ampli­
Fig. 13 compares the stress–strain hysteresis loops in axial and tude, the number of fatigue crack sources increases, which induces lower
torsional directions of the half-life cycle under different strain ratios. It is fatigue life. A certain number of secondary cracks can be observed under
noted that the increase of the multiaxial strain ratio has little impact on four different strain amplitudes. Previous studies [43,44] showed that
the axial plastic strain energy (area of the stress–strain hysteresis loop) the secondary crack of P92 steel starts from the crack tip and extends
at low strain amplitude (strain amplitudes of 0.2% and 0.3%). However, into the matrix along the transgranular direction. It can also see from the
at high strain amplitudes of 0.4% and 0.6%, the increase in the multi­ figure above that most of the secondary cracks are perpendicular to the
axial strain ratio reduces the axial plastic strain energy. For torsional fatigue crack growth direction. With the increase of strain amplitude,
plastic strain energy, the plastic strain energy increases significantly some secondary cracks tend to connect, making the secondary cracks
with increasing strain ratio. The sum of axial and torsional plastic strain more obvious. The secondary crack generally extends to the tearing
energy is taken as the total plastic strain energy. Fig. 14 plots the evo­ ridge. Tearing ridge is the main characteristic sign of mode I fracture
lution of the total plastic strain energy at different strain amplitudes and under low cycle fatigue loading. In the case of mode I fracture, fatigue
strain ratios. It is observed that the increases of the strain ratio and strain stripes perpendicular to the crack propagation direction are generated,
amplitude enhance the total plastic strain energy, unavoidably inducing as shown in Fig. 16. Moreover, it is observed that when the strain
large fatigue damage on the material. amplitude gradually increases from 0.2% to 0.4%, at the same magni­
Fig. 15 shows the evolution of fatigue life versus strain ratio. It is fication, the fatigue stripes gradually become sparse, indicating that the
noted that the increase of strain ratio results in the reduction of fatigue spacing of fatigue stripes is gradually increasing. According to the
life. With increasing strain ratio, torsional plasticity contributes greatly research of Wareing and Vaughan [45], the longer the crack, the greater
to the damage of materials, resulting in the obvious reduction in fatigue the fringe spacing, and the fringe spacing represents the macro growth
life. However, what stands out in the figure is that the effect of strain rate of the crack. In other words, the increase of strain amplitude pro­
ratio on the fatigue life is also affected by the strain amplitude. At a low motes the crack growth rate, which is characterized by the increase of
strain amplitude of 0.2%, the increase in strain ratio results in a fatigue fringe spacing. In the fracture morphology, some peculiar fatigue
stripes of tire traces can also be observed, which produce fatigue stripes

8
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 11. Variation of peak tensile stress of P92 steel under different strain ratios for various strain amplitudes: (a) Δε/2 = 0.2%, (b) Δε/2 = 0.3%, (c) Δε/2 = 0.4%,
(d) Δε/2 = 0.6%.

Fig. 12. Variation of softening rate under different multiaxial strain ratios: (a) axial softening rate, (b) torsional softening rate.

in two directions respectively under the action of torsional shear stress as shown in Fig. 17. It can be noted from the low magnification images
and axial tensile and compressive stress. The fatigue stripes in two di­ that the number of crack sources decreases with increasing multiaxial
rections cross each other to form the fatigue stripes of wheel tire traces, strain ratio, which shows that the increase of shear strain weakens the
which is also the mixed type fatigue fracture characteristics of type I and crack sources. The enlarged view of the crack source area also validates
type II. it. With the increase of shear strain, the crack source becomes flatter,
and the height drop of morphology on both sides decreases obviously.
3.2.2. Fracture behaviour at different strain ratio Moreover, it can be seen that there are many small cleavage planes with
To further study the influence of strain ratio on the fatigue crack different sizes, shapes, and heights in the crack propagation region. In
initiation and propagation behaviour, the fracture morphologies under addition, there are parallel and continuous fatigue stripes on each small
different strain ratios for a strain amplitude equal to 0.3% are observed, cleavage plane, but the fatigue stripes on the cleavage planes of several

9
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 13. Hysteresis loops of P92 steel under different multiaxial strain ratios: (a) Δε/2 = 0.2%, (b) Δε/2 = 0.3%, (c) Δε/2 = 0.4%, (d) Δε/2 = 0.6%.

10
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

also a certain number of carbides in the grain, which are arranged lin­
early, showing the position of the initial lath boundary. However, the
coarsening of carbides after multiaxial fatigue is not obvious, because
the coarsening of precipitates is mainly related to time [50]. In the
present work, the time of the tests is short (the maximum test time is
about 10 h). Hence the coarsening of precipitates hardly occurs. Addi­
tionally, most of the dislocations in the martensitic laths of P92 steel in
the as-received state are high-density dislocations (Fig. 1(b)). This high-
density dislocation is mainly formed by the interaction between
austenite grain boundary and martensite lath during martensitic trans­
formation. After multiaxial fatigue, most dislocations in the micro­
structure show the form of a dislocation network (Fig. 19), which is a
sign of the decrease of dislocation density [51]. It is noted that with the
increase in strain amplitude, the free dislocation region between laths
appears more frequently and the dislocation density decreases gradu­
ally. The dislocation lines in the subgrain appear to pass through the
Fig. 14. Relationship between strain energy density and strain ratio. subgrain boundaries. At low strain amplitude, their arrangement is
mainly cellular, and at high strain amplitude, the dislocation shows a
parallel band. At the subgrain boundary, dislocations show a large
number of disordered entanglements to form dislocation walls, and the
dislocation wall becomes more prominent at high strain amplitude.

3.3.2. Microstructure evolution under different strain ratios


Fig. 20 presents the microstructures of P92 steel under different
strain ratios for an axial strain amplitude of 0.3%. It is observed that the
width of martensitic lath increases and dislocation density inside the
laths also decreases with the increase of the multiaxial strain ratio. The
annihilation of free dislocations and dislocations with opposite symbols
in the lath is an important reason for the reduction of dislocations [52].
Under a low strain ratio, some knitted dislocations can be observed,
which are the characteristics of dissolution of subgrain boundary [53].
However, knitting dislocations cannot be observed at a high strain ratio,
which can be ascribed to the fact that a high strain ratio accelerates the
movement of free dislocations and dissolves the boundary of the sub­
grain. Moreover, with the increase of the strain ratio, some ladder and
banded dislocations appear in the microstructure, which is a planar
Fig. 15. Relationship between fatigue life and multiaxial strain ratio.
dislocation [54]. It shows that the increase of the strain ratio promotes
the cross slip in axial and torsional directions, resulting in this planar slip
adjacent faults are not parallel and continuous [46]. When the fatigue of dislocation.
fringes in the crack growth zone show good regularity, continuity, and
consistency, it shows that it is a kind of ductile fatigue fringes. When 4. Discussions
there is a large amount of plastic deformation in fatigue, the ductile
fatigue stripe also produces deformation and is no longer continuous 4.1. Effect of strain amplitude on the multiaxial fatigue damage behaviour
[47]. This is the reason why the faults between small cleavage planes
show fracture characteristics, and this phenomenon will become more The results section has shown that during the multiaxial fatigue
and more obvious with the increase in strain ratio.
process, both the axial and torsional plastic deformations increase with
increasing strain amplitude. Meanwhile, fatigue life is significantly
3.3. Microstructure evolution reduced. Notably, as the strain amplitude exceeds 0.4%, even if the
strain amplitude increases, damage caused by the increase of strain
3.3.1. Microstructure evolution under different strain amplitudes amplitude reaches saturation. Generally, fatigue damage is mainly
To understand the microstructural damage mechanisms under caused by the accumulation of plastic deformation [55]. The reduction
different loadings, the microstructures of P92 steel after fatigue failure of dislocation and the growth of subgrain are the main microstructural
under different strain amplitudes for a strain ratio equal to √3/5 are reasons for the fatigue damage.
observed, as shown in Fig. 18. An overview of these figures indicates At different strain amplitudes, P92 steel shows three continuous
that compared with the as-received material (Fig. 1(b)), the martensite softening stages during the fatigue process, as shown in Fig. 6. The dy­
laths tend to widen. Compared with the microstructure of strain am­ namic recovery caused by the rearrangement of dislocations in the laths
plitudes of 0.2% and 0.3%, there are some obvious subgrain structures contributes to the phenomenon. However, the softening rate of P92 steel
with increasing axial strain amplitudes to 0.4% and 0.6%. Benjamin shows obvious differences under different strain amplitudes. Therefore,
et al. [48] revealed that the growth of subgrain is the result of the the evolution trend of microstructure under different strain amplitude is
dissolution of the low angle boundaries containing subgrain and lath varied. It is observed from the microstructures in Fig. 19 that at low
boundaries. During the fatigue process of P92 steel, the subgrains in strain amplitude, the dislocation presents a cellular shape, which is a
martensite lath are growing continuously, and their number and size are low-energy structure. The formation of this cellular dislocation weakens
increasing [49]. High strain amplitude accelerates this process due to the cyclic softening effect of P92 steel [44], so it shows a lower softening
high plastic strain. Therefore, it is easier to observe subgrain under high rate at low strain amplitude. At high strain amplitude, the growth of
strain amplitude. Meanwhile, carbides, which are mainly distributed in subgrains after lath boundary dissolution can be observed in Fig. 18, and
the interior of grains and on the lath boundaries, are observed. There are the microstructure recovery is relatively uniform. This is because the

11
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 16. Fracture morphology of P92 steel under different axial strain amplitudes for a strain ratio equal to √3/5: (a) 0.2%, (b) 0.3%, (c) 0.4%, (d) 0.6%.

recovery of subgrain is related to the local plastic deformation at each contains chromium and other elements [57]. The interaction between
cycle [56]. At high strain amplitude, the local plastic deformation is also atoms in these elements and dislocations can effectively lock disloca­
large, which promotes the recovery of microstructure. The microstruc­ tions and produce a pinning effect. Studies on the high-temperature
ture formed by this recovery mechanism increases the average free path fatigue of P92 steel have shown that the precipitates in the lath inevi­
of dislocations [44], and dislocations pass through the subgrain tably aggregate and coarsen during the fatigue process [58,59].
boundary and show strip characteristics, promoting the increase of Although the test period in this study is short, no obvious precipitate
softening rate at high strain amplitude. aggregation and coarsening are observed. However, the internal solute
Considering the fact that the DSA is the result of the gravitational atoms also tend to aggregate. Because large plastic deformation accel­
interaction between moving dislocations and solute atoms [42], it is not erates the recovery of microstructure, as shown in Fig. 19(a) and (b), the
difficult to deduce that the DSA phenomenon is also closely related to aggregation of solute atoms is promoted. The frequency of pinning and
the evolution of microstructure. It is well known that the precipitate depinning between dislocation and solute atoms is thus enhanced.

12
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 17. Fracture morphology of P92 steel under different multiaxial strain ratios for an axial strain amplitude equal to 0.3%: (a) λ = 0, (b) λ=√3/10, (c) λ=√3/5,
(d) λ = 2√3/5 and (e) λ = 3√3/5.

Therefore, the number of stress serrations increase with increasing strain deformation [61]. This has an inhibitory effect on the DSA effect.
amplitude, as shown in Fig. 9(a). When the dislocation and solute atoms Therefore, when the strain amplitude is 0.6%, the maximum stress drop
are pinned, there is threshold stress that makes the dislocation free from tends to slow down, as shown in Fig. 9(b). The evolution of fatigue life in
the pinning effect [60]. When the threshold stress is overcome and the Fig. 10 also proves this phenomenon. Because the dislocation wall
dislocation is separated from the pinning, the dislocation continues to formed by dislocation accommodates plastic deformation, the fatigue
move and bind with the next solute atom and then remove the pinning. damage of plastic deformation on the microstructure is reduced. At the
Eventually, these dislocations annihilate with the dislocations of oppo­ same time, the decrease of DSA activity further reduces the damage. This
site symbols or form cellular structures with lower energy. However, explains the phenomenon that the effect of strain amplitude on the fa­
when the strain amplitude is large enough, dislocation forms a dislo­ tigue life is reduced when the strain amplitude exceeds 0.4%. The
cation wall, as shown in Fig. 19(c) and (d), which accommodates plastic detailed schematic damage mechanisms is shown in Fig. 21.

13
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 18. Microstructure of P92 steel under different axial strain amplitudes for a strain ratio equal to √3/5: (a) 0.2%, (b) 0.3%, (c) 0.4%, (d) 0.6%.

4.2. Effect of strain ratio on the multiaxial fatigue damage behaviour plastic strain, as shown in Fig. 6(f). Also due to this reason, the DSA
mainly appears in the axial direction, as shown in Fig. 13. Moreover,
Regarding the effect of strain ratio, results show that the total plastic since axial plastic strain plays a dominant role, the fracture morphology
strain energy is accumulating with increasing strain ratio, resulting in under a low strain ratio is closer to the uniaxial fatigue.
large fatigue damage, as shown in Fig. 14. However, it is found from At a relatively high strain ratio, the subgrain boundaries have dis­
Fig. 13 that with increasing strain ratio, the change of axial plastic strain solved, the carbide on the boundary distributes isolated in the grains, as
energy is relatively small, whereas the increasing trend of torsional shown in Fig. 20(d) and (e). These carbides are easy to attract disloca­
plastic strain energy is obvious, suggesting that the change in total tions for pinning, which promotes the DSA effect. Hence in Fig. 13, the
plastic strain energy is mainly brought about by the torsional component DSA can be obviously observed in the torsional direction at a high strain
with increasing strain ratio. Results also show that at high strain ratios, ratio. Because the DSA restricts the free movement of internal disloca­
the damage caused by the torsional part reaches saturation, which is tions, it weakens the softening of P92 steel. Thus, at a high strain ratio,
similar to the axial behaviour. With increasing strain ratio, torsional the axial softening rate decreases firstly and then increases again, as
plastic deformation plays a great role in the cyclic deformation and shown in Fig. 12(a). Moreover, some laddered dislocations are also
microstructure evolution, but its influence is still limited. observed in Fig. 20(d) and (e), and slip bands in the linear direction are
At a relatively low strain ratio, some braided dislocations near the observed in Fig. 20(d). These are typical characteristics of planar slip.
boundary are observed in Fig. 20(a), (b), and (c). Braided dislocations Because the pinning effect of dislocations in DSA limits the movement of
react with subgrain boundaries, resulting in the dissolution of subgrains. dislocations on the slip plane, making dislocations move in the form of
The increase of the strain ratio promotes this process. Obviously, the planar slip [62]. The main slip system is fixed and cannot rotate, which
dissolution of subgrains promotes the cyclic softening of materials, is more likely to occur during fatigue load, while the secondary slip
which is well reflected in Fig. 12. In addition, at a low strain ratio, the system is not easy to start [63]. The planar slip phenomenon is easier to
torsional plastic strain can be ignored as it is much smaller than the axial be observed at a high strain ratio (Fig. 20), indicating that the increase of

14
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 19. Dislocation morphology of P92 steel under different axial strain amplitudes for a strain ratio equal to √3/5: (a) 0.2%, (b) 0.3%, (c) 0.4%, (d) 0.6%.

the strain ratio strengthens the secondary slip under multiaxial fatigue. mechanisms of P92 steel at a high temperature have been evaluated. The
In addition, due to the accumulation of plastic deformation, a large influences of strain amplitude and strain ratio on the softening behav­
number of carbides gather near the boundaries of laths or grains, which iour, cyclic response, and fatigue life have been studied. In addition, the
is a favourable position for microcrack nucleation. Therefore, under the correlation between the mechanical properties and microstructure
combined effect of carbide and planar slip, the fracture morphology of evolution has been comprehensively discussed. The following conclu­
the specimen shows the characteristics of a broken small cleavage sur­ sions can be drawn:
face, as shown in Fig. 17. Secondary cracks are easy to occur between the
small cleavage planes of fatigue stripes in different directions, which (1) P92 steel shows three-stages softening characteristics in both
reduces the fatigue strength of P92 steel. The detailed damage mecha­ axial and torsional directions. The increases in strain amplitude
nism process is also shown in Fig. 21. and strain ratio accelerate the softening behaviour. However, the
In regard to the fatigue life (Fig. 15), at a low axial strain amplitude, effect of the axial strain amplitude reaches saturated after 0.4%
the fatigue life decreases rapidly with increasing strain ratio, while at a amplitude. Moreover, the effect of strain ratio is weakened with
high axial strain amplitude, the fatigue life is insensitive to the variation increasing axial strain amplitude. Axial plasticity plays a leading
of strain ratio. Because the axial plastic strain is very small at low axial role in the multiaxial fatigue damage of P92 steel.
strain amplitude (Fig. 6(f)), therefore the change in torsional plastic (2) The DSA behaviour shows a significant dependence on the
strain has a great impact on fatigue life. However, at a high axial strain loading conditions. At a low strain ratio, the DSA occurs in the
amplitude, the axial plastic deformation increases greatly (Fig. 6(f)), and axial direction when the specimen is stretched. Whereas at a high
the impact of the change in torsional plastic strain on fatigue life is strain ratio, the DSA appears in the torsional direction when the
reduced, demonstrating that the fatigue damage is dominated by axial specimen is compressed.
plastic deformation. In other words, the effect of torsional plastic (3) The fracture morphology of P92 steel under high-temperature
deformation plays a second role next to the axial plastic deformation. multiaxial fatigue is a mixed fracture mode of mode I and mode
II. When the torsional plasticity is less, the fracture morphology is
5. Conclusion similar to that of uniaxial fracture, mainly mode I fracture. When
the torsional plastic strain increases, the fatigue features show
In this work, the axial-torsion fatigue behaviour and damage tire marks, the cleavage surface is more broken.

15
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 20. Microstructure of P92 steel under different strain ratios for an axial strain amplitude equal to 0.3%: (a) λ = 0, (b) λ=√3/10, (c) λ=√3/5, (d) λ = 2√3/5 and
(e) λ = 3√3/5.

16
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

Fig. 21. Schematic microstructural damage mechanisms under different strain amplitudes and strain ratios.

(4) The dislocation structure and martensite structure are signifi­ [6] Viswanathan R, Bakker W. Materials for ultrasupercritical coal power
plants—Boiler materials: Part 1. J Mater Eng Perform 2001;10:96–101. https://doi.
cantly dependent on the multiaxial fatigue loading. The increase
org/10.1361/105994901770345394.
of strain amplitude accelerates the growth of subgrains in the [7] Chang Y, Xu H, Ni Y, Lan X, Li H. The effect of multiaxial stress state on creep
laths, promotes the dislocations transform from cellular to behavior and fracture mechanism of P92 steel. Mater Sci Eng, A 2015;636:70–6.
banded shape and reduces free dislocations. The increase of strain https://doi.org/10.1016/j.msea.2015.03.056.
[8] Razak NA. Experimental Study of Multiaxial Stress State on Creep Rupture Life and
ratio enhances the dislocation slip, promotes the movement of Ductility of P91 Steel. J Fail Anal and Preven 2021;21(6):1991–9. https://doi.org/
free dislocations, resulting in the reduction of dislocation density 10.1007/s11668-021-01270-z.
and dissolution of the subgrain boundaries. [9] Ren F, Tang X, Li B. Microstructure evolution analysis on creep behavior of Grade
91 steel under multiaxial state of stress. IOP Conf Ser: Earth Environ Sci 2019;233:
022013. https://doi.org/10.1088/1755-1315/233/2/022013.
[10] Nakamura H, Takanashi M, Itoh T, Wu M, Shimizu Y. Fatigue crack initiation and
Declaration of Competing Interest growth behavior of Ti–6Al–4V under non-proportional multiaxial loading. Int J
Fatigue 2011;33(7):842–8. https://doi.org/10.1016/j.ijfatigue.2010.12.013.
The authors declare that they have no known competing financial [11] Xu C, Tao T, Ke A. NON-PROPORTIONAL CYCLIC HARDENING BEHMIORS OF
1Cr18Ni9Ti STAINLESS STEEL. Acta Mech Sin 2001.
interests or personal relationships that could have appeared to influence [12] Mei J, Dong P, Kalnaus S, Jiang Y, Wei Z. A path-dependent fatigue crack
the work reported in this paper. propagation model under non-proportional modes I and III loading conditions. Eng
Fract Mech 2017;182:202–14. https://doi.org/10.1016/j.
engfracmech.2017.07.026.
Acknowledgements [13] Shamsaei N, Fatemi A. Effect of microstructure and hardness on non-proportional
cyclic hardening coefficient and predictions. Mater Sci Eng, A 2010;527(12):
The authors gratefully acknowledge the financial supports of the 3015–24. https://doi.org/10.1016/j.msea.2010.01.056.
[14] Shamsaei N, Gladskyi M, Panasovskyi K, Shukaev S, Fatemi A. Multiaxial fatigue of
National Natural Science Foundation of China (No. 52005250), China titanium including step loading and load path alteration and sequence effects. Int J
Postdoctoral Science Foundation (No. 2021 M691559) and Jiangsu Fatigue 2010;32(11):1862–74. https://doi.org/10.1016/j.ijfatigue.2010.05.006.
Planned Projects for Postdoctoral Research Funds (No. 2020Z320). [15] Colin J, Fatemi A. Variable amplitude cyclic deformation and fatigue behaviour of
stainless steel 304L including step, periodic, and random loadings. Fatigue Fract
Eng Mater Struct 2010;33:205–20. https://doi.org/10.1111/j.1460-
References 2695.2009.01429.x.
[16] Ince A, Glinka G. A generalized fatigue damage parameter for multiaxial fatigue
[1] Yoshizawa M, Igarashi M. Long-term creep deformation characteristics of advanced life prediction under proportional and non-proportional loadings. Int J Fatigue
ferritic steels for USC power plants. Int J Press Vessels Pip 2007;84(1-2):37–43. 2014;62:34–41. https://doi.org/10.1016/j.ijfatigue.2013.10.007.
https://doi.org/10.1016/j.ijpvp.2006.09.005. [17] Ma T-H, Chang Le, Guo S, Kong L-R, He X-H, Zhou C-Y. Comparison of multiaxial
[2] Pandey C, Saini N, Mahapatra MM, Kumar P. Study of the fracture surface low cycle fatigue behavior of CP-Ti under strain-controlled mode at different
morphology of impact and tensile tested cast and forged (C&F) Grade 91 steel at multiaxial strain ratios. Int J Fatigue 2020;140:105818. https://doi.org/10.1016/j.
room temperature for different heat treatment regimes. Eng Fail Anal 2017;71: ijfatigue.2020.105818.
131–47. https://doi.org/10.1016/j.engfailanal.2016.06.012. [18] Inoue T, Yoshida F, Niitsu Y, Ohno N, Uno T, Suzuki A. Inelastic stress-strain
[3] Maddi L, Shivhare R, Kumar V, Goel M, Ramesh M, Ballal A. Effect of tempering response of 214Cr-1Mo steel under combined tension-torsion at 600◦ C. Nucl Eng
time on the microstructure and stress rupture properties of P92 steel. Mater Today: Des 1994;150:107–18. https://doi.org/10.1016/0029-5493(94)90055-8.
Proc 2021;44:34–8. https://doi.org/10.1016/j.matpr.2020.06.125. [19] Inoue T, Kishi S, Koto H, Takahashi Y. Fatigue-creep life prediction of 214Cr-1Mo
[4] Hyde CJ, Sun W, Hyde TH, Saad AA. Thermo-mechanical fatigue testing and steel under combined tension-torsion at 600◦ C. Nucl Eng Des 1994;150:119–27.
simulation using a viscoplasticity model for a P91 steel. Comput Mater Sci 2012;56: https://doi.org/10.1016/0029-5493(94)90056-6.
29–33. https://doi.org/10.1016/j.commatsci.2012.01.006. [20] Inoue T, Sakane M, Fukuda Y, Igari T, Miyahara M, Okazaki M. Fatigue-creep life
[5] Mao J, Li X, Wang D, Zhong F, Luo L, Bao S, et al. Experimental study on creep- prediction for a notched specimen of 214Cr•1Mo steel at 600◦ C. Nucl Eng Des
fatigue behaviors of chinese P92 steel with consideration of several important 1994;150:141–9. https://doi.org/10.1016/0029-5493(94)90058-2.
factors. Int J Fatigue 2021;142:105900. https://doi.org/10.1016/j. [21] Inoue T, Okazaki M, Igari T, Sakane M, Kishi S. Evaluation of fatigue-creep life
ijfatigue.2020.105900. prediction methods in multiaxial stress state: The second report of the benchmark

17
N. Gao et al. International Journal of Fatigue 158 (2022) 106774

project (B) by the Subcommittee on Inelastic Analysis and Life Prediction of High [43] Zhang W, Wang X, Li X, Gong J, Wahab MA. Influence of prior low cycle fatigue on
Temperature Materials. JSMS. Nuclear Eng Des 1991;126(1):13–21. https://doi. microstructure evolution and subsequent creep behavior. Int J Fatigue 2018;109:
org/10.1016/0029-5493(91)90201-R. 114–25. https://doi.org/10.1016/j.ijfatigue.2018.01.001.
[22] Shang D-G, Sun G-Q, Chen J-H, Cai N, Yan C-L. Multiaxial fatigue behavior of Ni- [44] Zhang Z, Hu Z-F, Fan L-K, Wang B. Low Cycle Fatigue Behavior and Cyclic
based superalloy GH4169 at 650◦ C. Mater Sci Eng, A 2006;432(1-2):231–8. Softening of P92 Ferritic-martensitic Steel. J Iron Steel Res Int 2015;22(6):534–42.
https://doi.org/10.1016/j.msea.2006.06.014. https://doi.org/10.1016/S1006-706X(15)30037-6.
[23] Shang D, Sun G, Yan C, Chen J, Cai N. Creep-fatigue life prediction under fully- [45] Wareing J, Vaughan HG. The relationship between striation spacing, macroscopic
reversed multiaxial loading at high temperatures. Int J Fatigue 2007;29(4):705–12. crack growth rate, and the low-cycle fatigue life of a Type 316 stainless steel at
https://doi.org/10.1016/j.ijfatigue.2006.06.010. 625◦ C. Metal Science 1977;11(10):439–46. https://doi.org/10.1179/
[24] Wang X-W, Shang D-G, Guo Z-K. Multiaxial creep–fatigue Life Prediction Under msc.1977.11.10.439.
Variable Amplitude Loading at High Temperature. J of Materi Eng and Perform [46] Yin SM, Yang F, Yang XM, Wu SD, Li SX, Li GY. The role of twinning–detwinning
2019;28(3):1601–11. https://doi.org/10.1007/s11665-019-03919-1. on fatigue fracture morphology of Mg–3%Al–1%Zn alloy. Mater Sci Eng, A 2008;
[25] Kwon JD, Park JC. Multiaxial Fatigue Life Prediction of Duplex Stainless Steels 494(1-2):397–400. https://doi.org/10.1016/j.msea.2008.04.056.
with Thermal Aging at 430◦ C under Axial-Torsional Load. KEM 2004;270–273: [47] Forsyth PJE. Fatigue damage and crack growth in aluminium alloys. Acta Metall
1183–8. https://doi.org/10.4028/www.scientific.net/KEM.270-273.1183. 1963;11(7):703–15. https://doi.org/10.1016/0001-6160(63)90008-7.
[26] Li D-H, Shang D-G, Zhang C-C, Liu X-D, Li F-D, Li Z-G. Thermo-mechanical fatigue [48] Benjamin F, Maxime S, Alexandra R, Françoise B, André P. Microstructural
damage behavior for Ni-based superalloy under axial-torsional loading. Mater Sci evolutions and cyclic softening of 9%Cr martensitic steels. J Nucl Mater 2009;
Eng, A 2018;719:61–71. https://doi.org/10.1016/j.msea.2018.02.029. 386–388:71–4. https://doi.org/10.1016/j.jnucmat.2008.12.061.
[27] Ni YZ, Lan X, Xu H, Mao XP. Finite element analysis and experimental research on [49] Wang X, Zhang W, Ni J, Zhang T, Gong J, Wahab MA. Quantitative description
notched strengthening effect of P92 steel. Mater High Temp 2014;31(2):185–90. between pre-fatigue damage and residual tensile properties of P92 steel. Mater Sci
https://doi.org/10.1179/1878641314Y.0000000010. Eng, A 2019;744:415–25. https://doi.org/10.1016/j.msea.2018.12.029.
[28] Mao XP, Yu Y, Li C, Huang SD, Xu H, Ni YZ. Study on Creep Behaviors of T92 Steel [50] Hättestrand M, Andrén H-O. Evaluation of particle size distributions of precipitates
under Multiaxial Stress State. Adv Mater Res 2014;860–863:774–9. https://doi. in a 9% chromium steel using energy filtered transmission electron microscopy.
org/10.4028/www.scientific.net/AMR.860-863.774. Micron 2001;32(8):789–97. https://doi.org/10.1016/S0968-4328(00)00086-X.
[29] Zhang Z, Hu Z, Schmauder S, Mlikota M, Fan K. Low-Cycle Fatigue Properties of [51] Moorthy V, Choudhary BK, Vaidyanathan S, Jayakumar T, Rao KBS, Raj B. An
P92 Ferritic-Martensitic Steel at Elevated Temperature. J Mater Eng Perform 2016; assessment of low cycle fatigue damage using magnetic Barkhausen emission in
25(4):1650–62. https://doi.org/10.1007/s11665-016-1977-8. 9Cr–1Mo ferritic steel. Int J Fatigue 1999;21:263–9. https://doi.org/10.1016/
[30] Kimura M, Yamaguchi K, Hayakawa M, Kobayashi K, Kanazawa K. Microstructures S0142-1123(98)00079-6.
of creep-fatigued 9–12% Cr ferritic heat-resisting steels. Int J Fatigue 2006;28(3): [52] Sauzay M, Brillet H, Monnet I, Mottot M, Barcelo F, Fournier B, et al. Cyclically
300–8. https://doi.org/10.1016/j.ijfatigue.2005.04.013. induced softening due to low-angle boundary annihilation in a martensitic steel.
[31] ASTM E 2207 – 02: Standard practice for strain-controlled axial-torsional fatigue Mater Sci Eng, A 2005;400-401:241–4. https://doi.org/10.1016/j.
testing with thin-walled tubular specimens, Annual Book of ASTM Standards, West msea.2005.02.092.
Conshohocken, PA, 2008. [53] Eggeler G. The effect of long-term creep on particle coarsening in tempered
[32] Saad AA, Sun W, Hyde TH, Tanner DWJ. Cyclic softening behaviour of a P91 steel martensite ferritic steels. Acta Metall 1989;37(12):3225–34. https://doi.org/
under low cycle fatigue at high temperature. Procedia Eng 2011;10:1103–8. 10.1016/0001-6160(89)90194-6.
https://doi.org/10.1016/j.proeng.2011.04.182. [54] Nishino S, Hamada N, Sakane M, Ohnami M, Matsumura N, Tokizane M.
[33] Wang X, Jiang Y, Gong J, Zhao Y, Huang X. Characterization of Low Cycle Fatigue MICROSTRUCTURAL STUDY OF CYCLIC STRAIN HARDENING BEHAVIOUR IN
of Ferritic–Martensitic P92 Steel: Effect of Temperature. Steel Res Int 2016;87(6): BIAXIAL STRESS STATES AT ELEVATED TEMPERATURE. Fatigue Fract Eng Mater
761–71. https://doi.org/10.1002/srin.201500218. Struct 1986;9(1):65–77. https://doi.org/10.1111/j.1460-2695.1986.tb01212.x.
[34] Fournier B, Dalle F, Sauzay M, Longour J, Salvi M, Caës C, et al. comparison of [55] Xu J, Huo M, Xia R. Effect of cyclic plastic strain and flow stress on low cycle
various 9–12%Cr steels under fatigue and creep-fatigue loadings at high fatigue life of 316L(N) stainless steel. Mech Mater 2017;114:134–41. https://doi.
temperature. Mater Sci Eng, A 2011;528(22-23):6934–45. https://doi.org/ org/10.1016/j.mechmat.2017.07.014.
10.1016/j.msea.2011.05.046. [56] Fournier B, Sauzay M, Caes C, Noblecourt M, Mottot M, Bougault A, et al.
[35] Nagesha A, Kannan R, Sastry GVS, Sandhya R, Singh V, Bhanu Sankara Rao K, et al. Creep–fatigue–oxidation interactions in a 9Cr–1Mo martensitic steel. Part I: Effect
Isothermal and thermomechanical fatigue studies on a modified 9Cr–1Mo ferritic of tensile holding period on fatigue lifetime. Int J Fatigue 2008;30(4):649–62.
martensitic steel. Mater Sci Eng, A 2012;554:95–104. https://doi.org/10.1016/j. https://doi.org/10.1016/j.ijfatigue.2007.05.007.
msea.2012.06.021. [57] Shankar V, Valsan M, Rao KBS, Kannan R, Mannan SL, Pathak SD. Low cycle
[36] Yin P, Zhang W, Guo S, Wen J, Zhang G, Xue F, et al. Thermomechanical fatigue fatigue behavior and microstructural evolution of modified 9Cr-1Mo ferritic steel.
behaviour and damage mechanisms in a 9% Cr steel: Effect of strain rate. Mater Sci Materials Science and Engineering A, Structural Materials: Properties,
Eng, A 2021;815:141308. https://doi.org/10.1016/j.msea.2021.141308. Microstructure and Processing 2006;437(2):413–22.
[37] Chang Le, Li X, Wen J-B, Zhou B-B, He X-H, Zhang G-D, et al. Thermal-mechanical [58] Korcakova L, Hald J, Somers MAJ. Quantification of Laves phase particle size in
fatigue behaviour and life prediction of P92 steel, including average temperature 9CrW steel. Mater Charact 2001;47(2):111–7. https://doi.org/10.1016/S1044-
and dwell effects. J Mater Res Technol 2020;9(1):819–37. https://doi.org/ 5803(01)00159-0.
10.1016/j.jmrt.2019.11.022. [59] Guo X, Gong J, Jiang Y, Rong D. The influence of long-term aging on
[38] Shankar V, Kumar A, Mariappan K, Sandhya R, Laha K, Bhaduri AK, et al. microstructures and static mechanical properties of P92 steel at room temperature.
Occurrence of dynamic strain aging in Alloy 617M under low cycle fatigue loading. Mater Sci Eng, A 2013;564:199–205. https://doi.org/10.1016/j.
Int J Fatigue 2017;100:12–20. https://doi.org/10.1016/j.ijfatigue.2017.03.001. msea.2012.10.024.
[39] Rao CV, Srinivas NCS, Sastry GVS, Singh V. Dynamic strain aging, deformation and [60] Reppich B. On the attractive particle–dislocation interaction in dispersion-
fracture behaviour of the nickel base superalloy Inconel 617. Mater Sci Eng, A strengthened material. Acta Mater 1998;46(1):61–7. https://doi.org/10.1016/
2019;742:44–60. https://doi.org/10.1016/j.msea.2018.10.123. S1359-6454(97)00234-6.
[40] Rahman MS, Priyadarshan G, Raja KS, Nesbitt C, Misra M. Characterization of high [61] Gerland M, Violan P. Secondary cyclic hardening and dislocation structures in type
temperature deformation behavior of INCONEL 617. Mech Mater 2009;41(3): 316 stainless steel at 600◦ C. Mater Sci Eng 1986;84:23–33. https://doi.org/
261–70. https://doi.org/10.1016/j.mechmat.2008.10.003. 10.1016/0025-5416(86)90219-3.
[41] Nagesha A, Kannan R, Srinivasan VS, Sandhya R, Choudhary BK, Laha K. Dynamic [62] Li B, Zheng Y, Liu C, Li Q, Zhang Z, Chen Xu. Torsional thermomechanical fatigue
Strain Aging and Oxidation Effects on the Thermomechanical Fatigue Deformation behavior of 316LN stainless steel. Mater Sci Eng, A 2020;789:139676. https://doi.
of Reduced Activation Ferritic-Martensitic Steel. Metall Mater Trans A 2016;47(3): org/10.1016/j.msea.2020.139676.
1110–27. https://doi.org/10.1007/s11661-015-3293-6. [63] Franciosi P. The concepts of latent hardening and strain hardening in metallic
[42] Pham MS, Holdsworth SR. Dynamic strain ageing of AISI 316L during cyclic single crystals. Acta Metall 1985;33(9):1601–12. https://doi.org/10.1016/0001-
loading at 300◦ C: Mechanism, evolution, and its effects. Mater Sci Eng, A 2012; 6160(85)90154-3.
556:122–33. https://doi.org/10.1016/j.msea.2012.06.067.

18

You might also like