You are on page 1of 8

International Journal of Fatigue 87 (2016) 351358

Contents lists available at ScienceDirect

International Journal of Fatigue


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

Damage propagation mechanism in low-cycle creep fatigue


of CuCrZr alloy
Masaya Deguchi a,b,, Hirobumi Tobe a, Eiichi Sato a
a
Institute of Space and Astronautical Science, Japan Aerospace Exploration Agency, 3-1-1 Yoshinodai, Chuo, Sagamihara 252-5210, Japan
b
Department of Materials Engineering, Graduate School of Engineering, The University of Tokyo, 7-3-1 Hongo, Bunkyo, Tokyo 113-8656, Japan

a r t i c l e i n f o a b s t r a c t

Article history: This paper describes a characteristic damage propagation mechanism in low-cycle creepfatigue of
Received 22 December 2015 Cu0.7Cr0.09Zr (mass%), as investigated by creepfatigue tests including strain controlled fatigue and
Received in revised form 12 February 2016 stress-holding type creep, and following microstructural observations by scanning electron microscopy
Accepted 16 February 2016
(SEM). The total stress-holding time until rupture in the creepfatigue test was shorter than one-tenth
Available online 24 February 2016
of the rupture life in the simple creep test, and the rupture life of the specimen in the creepfatigue test
was shorter than half of that in the simple fatigue test. The SEM images suggest that the connection
Keywords:
between fatigue crack propagating along grain boundaries and intergranular creep voids rapidly
Copper alloys
Creepfatigue
accelerates crack propagation.
Crack paths Crown Copyright 2016 Published by Elsevier Ltd. All rights reserved.
Strain controlled fatigue

1. Introduction steady-state combustion. Engines in conventional launching rock-


ets are used after being subjected to several combustion tests. Very
The copper-based dilute CuCrZr alloy is one of the well- low cycle life of this thermal fatigue has long been considered one
known precipitation-strengthened alloys [1]. This alloy exhibits of the critical points for such engines [5]. Actually, after the quali-
good fatigue resistance because its zirconium additive results in fication test (QT) of another model of the engines, cracks were gen-
lower stacking fault energy; it also exhibits a high thermal conduc- erated after a few dozen combustion cycle [6].
tivity [2]. To make use of these characteristics, researchers have To evaluate the thermal fatigue of CuCrZr alloy during the
investigated CuCrZr alloys as candidate heat-transfer materials aforementioned QT combustion cycle of the engine, numerical
for the next-generation fusion reactor known as the International analyses were conducted by JAXAs Engineering Digital Innovation
Thermonuclear Experimental Reactor (ITER) [3]. (JEDI) center [6]: they simulated the interaction between thermo-
The Japan Aerospace Exploration Agency (JAXA) has started fluid behavior and structural responses through developing a new
developing a new liquid rocket engine, LE-9, with a combustion analysis procedure, that is high-fidelity multi-physics coupled sim-
chamber containing an inner cylinder made of CuCrZr alloy ulation to which three-dimensional finite element method (3D-
because of its intermediate strength, good fatigue resistance, and FEM) was applied. Fig. 2(a) shows a calculated temperaturetime
excellent thermal conductivity [4]. To prevent the combustion profile at the region subjected to the most severe condition: the
chamber from melting during operation, the inner cylinder has inner surface at the center of the groove wall at the nozzle throat
grooves with thin walls for liquid hydrogen as coolant. During of the engine. At this position, the temperature rapidly increases
engine combustion, this thin wall is exposed on one side to liquid immediately after ignition (from A to B), followed by a slight over-
hydrogen and on the other side to combustion gases (Fig. 1). Thus, shoot (from B to C). Afterwards, the temperature remains at
in one engine operation cycle, the CuCrZr alloy is subjected to approximately 753 K for 265 s during steady-state combustion
thermal fatigue consisting of both one cyclic deformation corre- (from C to D), and finally rapidly decreases to 73 K as soon as com-
sponding to start up and shut down and creep deformation during bustion ends (from D to E). The strain and stress histories are
shown in Fig. 2(b). The strain reaches its minimum value ( 2.2%)
at the highest temperature (B); with the subsequent temperature
Corresponding author at: Institute of Space and Astronautical Science, Japan overshoot, the wall is elongated above the tensile yield (C). After-
Aerospace Exploration Agency, 3-1-1 Yoshinodai, Chuo, Sagamihara 252-5210, wards, the strain gradually increases under an almost constant
Japan. Tel.: +81 50 3362 7539; fax: +81 42 759 8461. stress of 150 MPa during steady-state combustion (from C to D)
E-mail addresses: deguchi.masaya@ac.jaxa.jp (M. Deguchi), tobe@isas.jaxa.jp and finally reaches its maximum value (+1.0%) at approximately
(H. Tobe), sato@isas.jaxa.jp (E. Sato).

http://dx.doi.org/10.1016/j.ijfatigue.2016.02.025
0142-1123/Crown Copyright 2016 Published by Elsevier Ltd. All rights reserved.
352 M. Deguchi et al. / International Journal of Fatigue 87 (2016) 351358

CuCrZr alloy [8]. Strain holding, however, produces stress relax-


ation and does not correctly simulate the combustion cycle shown
in Fig. 2. In order to exactly replicate the thermal fatigue of CuCr
Zr alloy in one combustion cycle, our research group conducted
creepfatigue tests in which a stress-holding process was intro-
duced [9]. The results indicated that the introduction of the
stress-holding process resulted in a large decrease in lifetime to
less than 20 cycles; by contrast, the NIMS database [8] reports a
lifetime of larger than 60 cycles under similar conditions. This dra-
matic decrease in lifetime suggests that characteristic damage
propagation occurs via the stress-holding process during low-
cycle creepfatigue.
Our group conducted microstructural observations of the frac-
tured specimens used in the creepfatigue tests and compared
their microstructures to those of samples used in simple creep
tests and simple fatigue tests [10]. The results imply that creep
voids nucleated at grain boundaries and fatigue cracks from the
surface of the specimen connect with each other, resulting in final
fracture. However, the damage propagation process should be
traced following the progress in the creepfatigue cycles.
Fig. 1. Image of a liquid rocket engine and schematic of its combustion chamber
wall. In this study, after reviewing the results of our own mechanical
creepfatigue, simple creep, and simple fatigue tests [10], we used
a scanning electron microscope to observe the microstructures of
CuCrZr alloy specimens subjected to interrupted creepfatigue
tests. These observations enabled us to elucidate the process and
mechanism of damage propagation in the alloy.

2. Experimental

2.1. Material

The alloy used in this study had a nominal chemical composi-


tion of Cu0.7Cr0.09Zr (mass%); it is the identical material as that
used in Ref. [10], and is the nominally same but a different-rod
material than that used in Ref. [9]. The alloy was received as a
bar that was fabricated through vacuum induction melting, hot
forging, solution treatment, and aging under appropriate condi-
tions, which are summarized in the NIMS data sheet [8].

2.2. Mechanical tests

All of the mechanical tests were conducted using a hydraulic-


servo fatigue testing machine (EFH-10-5-40, Saginomiya, Tokyo,
Japan). The specimen had a gauge length of 13 mm, a diameter of
6 mm, and a sufficiently large shoulder radius of 20 mm for high-
temperature low-cycle fatigue tests according to the applicable
Fig. 2. Results of 3D-FEM simulation for one engine-combustion cycle: (a)
JIS standard [11]. The specimen was heated at a high-frequency
temperaturetime profile and (b) strain/stresstime profiles. induction heating facility (GVL-153BC, Fuji Electronic Industrial,
Saitama, Japan). An R-type thermocouple welded onto the shoulder
of the specimen was used for controlling and measuring the spec-
the lowest temperature (E). The shut-down behavior (after D) is imen temperature. The strain in the gauge length was measured
not well-estimated in this calculation. These analyses clearly reveal using a high-temperature extensometer (No. 2632-055, Instron,
that the CuCrZr alloy is subjected not only to severe fatigue Norwood, United States). The aforementioned apparatus were con-
deformation (from A to B and from D to E) but also to tensile defor- tained in a vacuum chamber.
mation with a very low strain rate (from C to D) in one engine com- Before conducting the tests, the chamber was evacuated to a
bustion cycle. The former strain amplitude is 3%, which is pressure below 10 Pa; the specimen was then heated to 753 K at
substantially larger than the strain range observed in typical a rate of 100 K min 1 and maintained at 753 K for 30 min to reach
creepfatigue phenomena in power plants, etc. [3]. The latter is temperature equilibrium.
metallurgically equivalent to stress-holding creep deformation. According to the results of the 3D-FEM analyses shown in Fig. 2
In general, such thermal fatigue has been examined by low- (b), the conditions of the mechanical tests were determined as
cycle fatigue experiments under constant temperature, into which summarized in Table 1. The engine cycle condition was replaced
a strain-holding process is introduced [7]. Creepfatigue life on the with creepfatigue consisting of strain-controlled fatigue (com-
basis of stress-holding-type creepfatigue data had been assessed pressed and tensile strains of 2.2% and 1.0%, respectively, with
[5]. Later, the National Institute for Materials Science (NIMS) con- a strain rate of 1.0  10 3 s 1) and a stress-holding creep in ten-
structed a material database of the creepfatigue properties of the sion (stress of 150 MPa for holding time of 270 s) under a constant
M. Deguchi et al. / International Journal of Fatigue 87 (2016) 351358 353

Table 1
Conditions used in the three types of executed mechanical tests.

Test mode Temperature (K) Creep Fatigue


1
Stress hold period Compression (%) Tension (%) Strain rate (s )
Creep 753 150 MPa
3
Fatigue 753 2.2 +1.0 1.0  10
3
Creepfatigue 753 150 MPa for 270 s 2.2 +1.0 1.0  10

Fig. 3. Results of the creepfatigue, simple creep, and simple fatigue tests: (a) stress history and (b) strain history of the creepfatigue test, (c) strain history of the simple
creep test, and (d) stress history of the simple fatigue test.

temperature equal to that during the steady-state combustion where (a) and (b) show the stress and strain histories, respectively,
(753 K). To examine the effect of superposition of creep and fati- of the creepfatigue test, (c) shows the strain history of the simple
gue, simple tensile creep and simple fatigue tests were conducted creep test, and (d) shows the stress history of the simple fatigue
under the same conditions of the stress-holding process and the test. The total holding time until rupture in the creepfatigue test
fatigue process of the creepfatigue test, respectively. was 4900 s (270 s  18 cycles + 40 s), whereas the rupture life in
Several creepfatigue tests were interrupted at the end of the the simple creep test was approximately 58,000 s. The rupture life
cycles of N/Nf = 0.4, 0.6, and 0.8, where N is the cycle number in the creepfatigue mode was shorter than one-tenth of that in
and Nf is that at fracture. the simple creep mode. The rupture life in the creepfatigue test
(19 cycles) was shorter than one-half of that in the simple fatigue
2.3. Microstructural observation test (42 cycles).1
These rupture lives indicate that the creepfatigue life is much
The fracture surfaces of the specimens after the mechanical shorter than that predicted by the linear cumulative damage rule:
tests were observed by a field-emission scanning electron micro- material fractures when the summation of the creep damage (Uc)
scopy (FE-SEM) (JSM-7100F, JEOL, Tokyo, Japan). The cross- and fatigue damage (Uf) reaches unity (Uc + Uf = 1) [12]. Here, Uc
sections along the loading axis of the central part of the interrupted and Uf are defined as Uc = tcf/tsc and Uf = Ncf/Nsf, respectively,
and fractured specimens were then observed also by an FE-SEM. where tcf and tsc are rupture life (time) in the creepfatigue and
The cross-sections were mechanically polished and buff-finished simple creep tests, respectively, and Ncf and Nsf are rupture life
using a mixture of 28 mass% ammonia water and colloidal silica (cycle) in the creepfatigue and simple fatigue, respectively. The
with a particle size of 0.04 lm. Finally, these specimens were present results indicate that fracture in the creepfatigue test
etched for 1015 s in an ethanol solution of 1 mass% FeCl3. occurred with Uc = 0.08 and Uf = 0.45, i.e., Uc + Uf = 0.53. The life
was only approximately one-half of the value expected one on
3. Results and discussion the basis of the rule.

3.1. Creepfatigue test 1


Another creep-fatigue test under the same condition showed the rupture life of
21 cycles and total holding time until rupture of 5400 s (270 s  20 cycles + 0 s). In
High temperature tensile tests revealed that the alloy has 0.2% our groups earlier work [9], where the nominally same but a different-rod material
than that used in this study was used, the rupture life in creepfatigue was 20 cycles
proof stress of 180 MPa and tensile strength of 201 MPa at 753 K and total holding time until rupture of 5175 s (270 s  19 cycles + 45 s) and the
under vacuum with strain rate of 1.0  10 4 s 1. Fig. 3 shows the rupture life in simple creep was 58,000 s, respectively, both of which are similar to
results of the creepfatigue, simple creep, and simple fatigue tests, the results in this study.
354 M. Deguchi et al. / International Journal of Fatigue 87 (2016) 351358

Fig. 3(a) and (b) shows that the specimen fractured during the
stress-holding period (creep deformation) in the 19th cycle. The
magnitude of the maximum and minimum stresses in a cycle,
which is necessary for fatigue deformation gradually decreased.
The strain induced during creep deformation gradually increased
with increasing cycle number until the 14th cycle and then rapidly
increased after the 15th cycle until fracture. To show these phe-
nomena more visibly, Fig. 4 shows the stress/straintime profiles
and the stressstrain hysteresis loops of the 2nd and 17th cycles,
which correspond to N/Nf of 0.1 and 0.9, respectively. The maxi-
mum stresses necessary for fatigue deformation in both compres-
sive and tensile deformation at the 17th cycle ( 187 MPa and
168 MPa) were lower than those at the 2nd cycle ( 203 MPa and
192 MPa). The creep strain in the 17th cycle (e = 1.2%) was much
larger than that in the 2nd cycle (e = 0.33%). These results indicate
the damage accumulated from the 2nd until the 17th cycle.
Fig. 5(a) shows the strain histories of the stress-holding region
extracted from the strain histories shown in Fig. 3(b) for the 1st,
7th, 11th, 15th, 17th, and 18th cycles. Transient creep was evident
in every cycle and yielded more than half of the total creep strain
except the last two cycles. Creep strain increased with increasing
cycle number in both the transient and steady-state regions. Fig. 5. Results of the creepfatigue test: (a) creep strain histories in the 1st, 7th,
Fig. 5(b) shows the change in the total creep strain in a cycle. These 11th, 15th, 17th, and 18th cycles and (b) change in the total creep strain in a cycle.
results indicate that the creep strain gradually increased until the
15th cycle and then drastically increased immediately after the
15th cycle, corresponding to an N/Nf of 0.8. Therefore, we focused
our subsequent microstructural observations of the damage
around the 15th cycle.

3.2. Comparison with simple creep test and simple fatigue test

Fig. 6(a) shows a comparison between the creep curve of the


simple creep test and the accumulative creep curve of the creep
fatigue test; the latter represents stacks of the extracted creep
curve from every cycle. Instead of a large difference in rupture life,
the rupture strains were in a similar order, i.e., 13% and 24% in the
creepfatigue and simple creep tests, respectively. This rapid
accumulation of creep strain in the creepfatigue test is mainly
attributed to transient creep, and is considered as the origin of
small rupture life in the creepfatigue test.

Fig. 6. (a) Comparison between the accumulative creep curve of the creepfatigue
test and the creep curve of the simple creep test; (b) comparison between the
maximum and minimum stresses in each cycle in the creepfatigue test and the
simple fatigue test.

Fig. 6(b) shows the maximum and minimum stresses in each


cycle in the creepfatigue and simple fatigue tests. The initial
behavior until the 15th cycle is the same as each other in the
two tests, where the maximum and minimum stresses gradually
decrease with increasing cycle number. In the creepfatigue test,
however, a substantial decrease in the maximum stress is observed
immediately after the 15th cycle, whereas no such decrease is
observed in the simple fatigue test. Immediately after this decrease
in the maximum stress, the specimen in the creepfatigue test
fractured. Notably, not only this precipitous decrease in the maxi-
mum tensile stresses but also a substantial increase in creep strain
(Fig. 5(b)) occurred after the 15th cycle in the creepfatigue test.
Fig. 4. Comparison between the 2nd and 17th cycles of the creepfatigue test: (a) We therefore consider that substantial damage has been accumu-
stress/straintime profiles and (b) stressstrain hysteresis loop. lated until the 15th cycle.
M. Deguchi et al. / International Journal of Fatigue 87 (2016) 351358 355

3.3. Microstructural observations of fracture surfaces Fig. 8(a) and (b) shows the fracture surfaces of the specimens
subjected to the simple creep and simple fatigue tests, respectively.
Fig. 7(a) shows the fracture surface of the specimen subjected to Fig. 8(c) and (e) is magnified images corresponding to the rectan-
the creepfatigue test, and Fig. 7(b)(e) is magnified images corre- gles denoted by C and E in Fig. 8(a), respectively, and Fig. 8
sponding to the rectangles denoted by B, C, D, and E in Fig. 7(a), (d) and (f) is those denoted by D and F in Fig. 8(b), respectively.
respectively. Fig. 7(b)(d) clearly shows intergranular cracks being Fig. 8(c) and (d) shows that the cracks in both the simple creep
predominant without striations from crack initiation at the surface and simple fatigue tests initiated from the specimen surface as
until the cracks reach the specimen center. Thus, under this condi- pure intergranular cracks without striations. Fig. 8(e) shows that,
tion, the cracks not only initiated but also propagated in the inter- in the vicinity of the specimen center in the simple creep test, only
granular mode. Fig. 7(e) shows the final ductile-fracture surface in a dimple pattern is identified; no intergranular cracks are
the vicinity of the specimen center; this surface was formed by observed. In contrast, Fig. 8(f) shows that there are only intergran-
tension immediately before fracture. This observation is quite sim- ular cracks but no dimple pattern in the vicinity of the specimen
ilar to those reported by our group in the previous works [9,10]. center in the simple fatigue test.

Fig. 7. SEM micrographs of the specimen fractured in the creepfatigue test: (a) the entire fracture surface; (be) high-magnification micrographs corresponding to the
rectangles denoted by BE in (a), respectively.

Fig. 8. SEM micrographs of the fractured specimens (a) in the simple creep test and (b) in the simple fatigue test; (c, e) high-magnification micrographs corresponding to the
rectangles denoted by C and E in (a), respectively; (d, f) high-magnification micrographs corresponding to the rectangles denoted by D and F in (a), respectively.
356 M. Deguchi et al. / International Journal of Fatigue 87 (2016) 351358

Fatigue crack initiation in wavy slip materials and planar slip observation because significant changes in creep strain and maxi-
materials differs; in the former case, transgranular cracking due to mum stress were conspicuous in Figs. 5 and 6(b), respectively.
persistent slip bands is known to occur [13,14]. Although CuCr Fig. 9(a)(d) shows the history of creep-void growth in
Zr alloy is classified as a planar-slip material [2], the occurrence of sequence. Numerous-voids much smaller than 1 lm were
transgranular or intergranular cracking depends on the loading fre- observed on grain boundaries at the 7th cycle (a); these voids grew
quency, atmosphere, and temperature [1517]. For instance, at to 35 lm in length at the 11th cycle (b) and to 1015 lm at the
lower frequencies and higher temperatures, intergranular cracking 15th cycle (c). Finally, some elongated voids of 30 lm were
tends to occur. It is clear that CuCrZr alloy exhibited intergranular observed in the fractured specimen (d). Dose and Cadek also
cracking in the present condition. Although studies on the transition reported the formation of numerous grain-boundary voids during
from transgranular to intergranular cracking in CuCrZr alloy have creep deformation [18]. Changes in the length of the creep void
not been reported, the cracking mode might change under different av and accumulative creep strain eCREEP are plotted in Fig. 9(e),
conditions corresponding to a different engine design. which indicates that creep-void growth strongly influenced the
increase in accumulative creep strain; both gradually increased
until fracture. These results indicate that a substantial change in
3.4. Microstructural observations of cross-sections creep-void size was not observed even at N/Nf  0.8.
Fig. 10(a)(c) shows the history of fatigue-crack propagation in
Notably, both the grain size and the amount of twins were con- sequence, where the cracks selected for representation here were
firmed to be approximately similar between the as-received and the longest ones in each specimen. Fig. 10(d) shows the vertical
tested specimens through optical micrography. cross-section of the fractured specimen. The crack depths in
The cross-sections of the fractured and interrupted specimens specimens with N/Nf of 0.4, 0.6, and 0.8 were 40 lm, 75 lm,
in the creepfatigue tests were observed by an FE-SEM (Figs. 9 and 280 lm, respectively; however, the crack depth in the case of
and 10), where the tests were interrupted at the end of the 7th, N/Nf = 1.0 reached as large as a few millimeters because the
11th, and 15th cycles, corresponding to N/Nf of 0.4, 0.6, and 0.8, specimen diameter was 6 mm. That is, the crack depth gradually
respectively. The specimen subjected to 15 cycles was chosen for increased in earlier cycles and then drastically increased

Fig. 9. (ad) SEM micrographs of the cross-sections of the interrupted specimens in the creepfatigue tests: (a) the 7th cycle, (b) 11th cycle, (c) 15th cycle, (d) fractured
specimen. (e) Plot of creep strain and creep-void size changes with increasing cycle number.
M. Deguchi et al. / International Journal of Fatigue 87 (2016) 351358 357

Fig. 10. (ad) SEM micrographs of the cross-sections of the interrupted specimens in the creepfatigue tests: (a) the 7th cycle, (b) 11th cycle, (c) 15th cycle, (d) fractured
specimen. (e) Change in crack depth and the maximum stress normalized by that in the 1st cycle.

immediately before fracture. The changes in crack depth c and the


maximum stress normalized by that in the 1st cycle RrMAX are plot-
ted in Fig. 10(e). This plot indicates that the propagating crack depth
attributed to the maximum stress changes until the 15th cycle, and
then, crack depth increases drastically after the 15th cycle. On the
other hand, Fig. 10(d) shows numerous elongated creep voids near
the fracture surface similarly to aforementioned Fig. 9(d).
Fig. 11 presents a magnified image of the crack tip in Fig. 10(c);
this high-magnification SEM image shows the fatigue crack tip of
the creepfatigue-tested specimen interrupted at the 15th cycle.
As evident in Fig. 10(a)(b), the crack tips of the specimens sub-
jected to 7 or 11 cycles are narrow and sharp, whereas that of the
specimen subjected to 15 cycles exhibits a shape with an expanded
width and an elliptic tip. The crack tip is considered to connect with Fig. 11. High-magnification SEM micrograph of the fatigue crack tip in the creep
an elongated creep void at approximately the 15th cycle because fatigue-tested specimens interrupted at the 15th cycle. Note that the shape of the
both the size and morphology of this crack tip are similar to those crack tips at the 15th cycle is different from that at previous cycles; Fig. 10(a and b)
of the creep void shown in Fig. 9(c). The repetition of this connec- shows sharp crack tips.

tion between a fatigue crack and elongated creep voids, both of


which are formed at grain boundaries, begins to occur at approxi- 3.5. Damage propagation mechanism
mately the 15th cycle and drastically accelerates the failure of the
specimen in the creepfatigue test compared to the failures of the Nucleation and growth of creep voids on grain boundaries
simple-creep and the simple-fatigue-test specimens. occurred alone in the simple creep test; by contrast, only propaga-
358 M. Deguchi et al. / International Journal of Fatigue 87 (2016) 351358

Fig. 12. Schematic of the proposed damage propagation process under the creepfatigue conditions: (a) earlier term (N/Nf = 0.4), (b) intermediate term (N/Nf = 0.6), and (c)
later term (N/Nf = 0.8).

tion of fatigue cracks along grain boundaries occurred in the simple connect with grown creep voids in the later term of the cycles.
fatigue test. However, both nucleation and growth of creep voids These results suggest that the repetition of this connection
and propagation of fatigue cracks occurred during the creepfatigue between fatigue cracks and creep voids accelerates cracks to prop-
tests. This result is consistent with previously published results for agate rapidly inside the specimen, which were attributed to the
stainless steel and nickel-based alloys [12,15,19], except that the short lifetime in creepfatigue of the CuCrZr alloy.
fatigue cracks observed in the present study were classified as pure
intergranular cracks. Acknowledgement
Fig. 12 shows schematics of the proposed damage propagation
process in the creepfatigue tests of CuCrZr alloy. As shown in This research was supported by Grants-in-Aids for Scientific
Fig. 12(a), numerous minute voids nucleated on grain boundaries Research from the Japan Society for the Promotion of Science.
and fatigue cracks began to propagate from the surface of the spec-
imen intergranularly during the early term of the test (N/Nf = 0.4, References
the 7th cycle). In Fig. 12(b), the small voids grew larger and fatigue
cracks simultaneously continued to propagate along grain bound- [1] Butterworth GJ, Forty CBA. A survey of the properties of copper alloys for use as
fusion reactor materials. J Nucl Mater 1992;189:23776.
aries during the intermediate term (N/Nf = 0.6, the 11th cycle). [2] Batra IS, Dey GK, Kulkarni UD, Banerjee S. Microstructure and properties of a
Fig. 12(c) shows a tip of the fatigue crack propagating inside the CuCrZr alloy. J Nucl Mater 2001;299:91100.
specimen connected with grown voids during the latter term (N/ [3] Li M, Werner E, You JH. Low cycle fatigue behavior of ITER-like diverter target
under DEMO-relevant operation conditions. Fusion Eng Des 2015;90:8896.
Nf = 0.8, the 15th cycle). The crack depth reached several millimeters [4] Atsumi M, Yoshikawa K, Ogawara A, Onga T. Development of LE-X engine. MHI
in a few cycles after the 15th cycle because the fatigue crack tip Tech Rev 2011;48(4):408.
began to connect with grown creep voids. These connections [5] Kitade S, Igari T, Fukahori O, Ohigashi H. Creepfatigue life assessment for
combustion-chamber of the first stage engine of H-II rocket. MHI Tech Rev
between creep voids and fatigue crack tips arose extensively inside 1990;27(6):51520 [in Japanese].
the specimen subjected to more than 15 cycles; these connections [6] Nishimoto M, Yamanishi N, Yoshimura S, Kasahara N, Akiba H. Failure
accelerated damage propagation during creepfatigue of the Cu mechanism of combustion chamber of liquid rocket engine under severe
multi-physics conditions (residual deformation of chamber throat). Trans
CrZr alloy.
ASME 2012;78A:4052 [in Japanese].
Hales [19] proposed four failure mechanisms depending on the [7] Chen X, Yang Z, Sokolov MA, Erdman III DL, Mo K, Stubbins JF. Effect of creep
total strain amplitude and the number of cycles. In the case of Cu and oxidation on reduced fatigue life of Ni-based alloy 617 at 850 C. J Nucl
Mater 2014;444:393403.
CrZr alloy at the temperature and stress amplitude conditions
[8] NIMS Space Use Materials Strength Data Sheet No. 13. Data sheet in low-cycle
classified as Case 2, fatigue crack damage initiates and then creep fatigue properties of CuCrZr alloy; 2009. <http://smds.nims.go.jp/MSDS/pdf/
cavity damage initiates after the critical amount of accumulative sheet/SP13J.pdf>.
creep strain is achieved. [9] Yanagi S, Imai S, Kawai N, Sato E. Thermal fatigue characteristics of copper
alloy for rocket engine combustion chamber. J Jpn Inst Copper 2012;51:6670
[in Japanese].
4. Conclusions [10] Deguchi M, Hori Y, Tobe H, Sato E. Investigation in damage process during
creep fatigue of copper based alloy for rocket engine combustion chamber. J
Jpn Inst Copper 2015;54:6772 [in Japanese].
In this study, low-cycle creepfatigue tests with a stress- [11] Method of high temperature low cycle fatigue testing for metallic materials, JIS
holding process were conducted at 753 K for Cu0.7Cr0.09Zr Z2279-1992.
[12] Skelton RP, Gandy D. Creepfatigue damage accumulation and interaction
(mass%) alloy to simulate a combustion cycle of a liquid rocket
diagram based on metallographic interpretation of mechanisms. Mater High
engine; simple creep and simple fatigue tests were also conducted Temp 2008;25(1):2754.
under the corresponding conditions. The total stress-holding time [13] Liang F-L, Laird C. Control of intergranular fatigue cracking by slip homogeneity
until rupture in the creepfatigue test was 4900 s, which was in copper I: effect of grain size. Mater Sci Eng A 1989;117:95102.
[14] Liang F-L, Laird C. Control of intergranular fatigue cracking by slip homogeneity
shorter than one-tenth the rupture life in the simple creep test of in copper I: effect of loading mode. Mater Sci Eng A 1989;117:10313.
58,000 s. The rupture life of the specimen in the creepfatigue test [15] Gell M, Leverant GR. Mechanisms of high-temperature fatigue. In: Carden AE,
(19 cycles) is shorter than one-half of that in the simple fatigue test McEvily AJ, Wells CH, editors. Fatigue at elevated
temperatures. Philadelphia: American Society for Testing and Materials; 1973.
(42 cycles). These results indicate that the creepfatigue life cannot p. 3767.
be predicted by the conventional linear cumulative-damage rule. [16] Dahal J, Maciejewski K, Ghonem H. Loading frequency and microstructure
The surfaces of the fractured specimens in the creepfatigue, interactions in intergranular fatigue crack growth in a disk Ni-based
superalloy. Int J Fatigue 2013;57:93102.
simple creep, and simple fatigue tests were observed by SEM. [17] Carroll LJ, Cabet C, Carroll MC, Wright RN. The development of microstructural
The cross-sections of the fractured and interrupted specimens of damage during high temperature creepfatigue of a nickel alloy. Int J Fatigue
creepfatigue tests were also observed by SEM. The observations 2013;47:11525.
[18] Dose F, Cadek J. An assessment of time to creep fracture due to diffusional
showed that creep voids were generated on grain boundaries and growth of grain boundary voids. Mater Sci Eng 1972;9:3559.
then grew, and that fatigue crack simultaneously propagated from [19] Hales R. A quantitative metallographic assessment of structural degradation of
the specimen surfaces along grain boundaries. In addition, magni- type 316 stainless steel during creepfatigue. Fatigue Eng Mater 1980;3
(4):33956.
fied SEM image clearly showed that the fatigue crack tip began to

You might also like