You are on page 1of 8

International Journal of Fatigue 84 (2016) 916

Contents lists available at ScienceDirect

International Journal of Fatigue


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

Effect of grain size on fatigue-crack growth in 2524 aluminium alloy


Deyan Yin a,b, Huiqun Liu a,b,, Yuqiang Chen c,, Danqing Yi a,b, Bo Wang a,b, Bin Wang a,b, Fanghua Shen a,b,
Shang Fu a,b, Cong Tang a,b, Suping Pan a,b
a
School of Materials Science and Engineering, Central South University, Changsha, Hunan 410083, PR China
b
Key Laboratory of Nonferrous Metal Materials Sciences and Engineering, Ministry of Education, Central South University, Changsha, Hunan 410083, PR China
c
Hunan University of Science and Technology, Xiangtan, Hunan 411201, PR China

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

Article history: In this study, industrial 2524 aluminium alloy plates with various grain sizes (0.8298 lm) were pre-
Received 25 August 2015 pared by cold rolling and heat treatment. The fatigue-crack-growth rate was studied as a function of grain
Received in revised form 6 November 2015 size through fatigue tests and microstructural observations. The results showed that grain refinement led
Accepted 13 November 2015
to a decrease in the resistance against fatigue-crack growth. Besides, the levels of crack closure in coarse-
grained samples were higher than those in fine-grained ones at low values of the range of the stress
intensity factor K, DK. This phenomenon was predicted and explained well by the crack-deflection model.
Keywords:
2015 Elsevier Ltd. All rights reserved.
Grain-size tailoring
Fatigue-crack growth
2524 aluminium alloy

1. Introduction fatigue-growth resistance with larger grain size is attributed to


the decrease in driving force of changes in the crack path, which
AlCuMg (2000 series) aluminium alloys are widely used in is induced by the microstructure and possible contact between
aircrafts because of their outstanding damage-tolerance and facets of rough cracks [16]. It is well known that this beneficial
fatigue-resistance properties. An excellent example is the 2524 effect is typically more pronounced at a low range of the stress
alloy, which was developed by Alcoa in the 1990s and has been intensity factor K, DK, where both the size of cyclic plastic zones
used as wing and fuselage skin on Boeing 777 and Airbus A380 and the opening displacement of cyclic crack tips are smaller than
[14]. The safety of an aircraft is closely related to the fatigue prop- the size of grains [17].
erty of the alloy used. Earlier studies revealed that experimental In order to study the effects of grain size on fatigue behaviour,
conditions such as applied loading and environmental factors play researchers have obtained alloys with different grain sizes using
significant roles on an alloys fatigue properties [5,1,6,7]. However, different methods such as electrodeposition, equal-channel angu-
recent studies indicated that intrinsic factors, including physical lar pressing, and cryomilling [1822]. These methods have inevita-
metallurgical properties, of the alloy have stronger effects on its bly introduced different characteristics, such as dislocations and
fatigue properties [812]. defects, which increase the complexity when comparing the effects
Among the physical metallurgical properties, grain size is one of of grain size on the fatigue behaviour of these alloys. Besides, as it
the most important factors. Grain size significantly influences both is difficult to obtain sufficient number of large samples by these
fatigue-crack initiation and propagation behaviour [13,14]. On the methods, some works have been conducted on non-standard sam-
basis of previous work on microcrystalline metals (the grain sizes ples, which might have affected the validity of the experiments.
were typically above 1 lm) and ultra-fine-grained crystalline met- Therefore, it is important to obtain materials with significantly dif-
als (the grain sizes were typically in the range of 100 nm to 1 lm), ferent grain sizes under the same processing conditions to examine
it is generally recognized that a structure with relatively larger their effects on the fatigue behaviour of the 2524 alloy.
grain size tends to have a higher fatigue-crack-growth threshold In our present work, we successfully prepared 2524 aluminium
(DKth) and lower crack-growth rate [14,15]. The increase in alloy plates with widely varying grain sizes through conventional
cold rolling and heat treatment. The fatigue-crack-growth beha-
viour of these plates was investigated in detail with a fatigue test-
Corresponding authors at: School of Materials Science and Engineering, Central
ing machine, an optical microscope, a transmission electron
South University, Changsha, Hunan 410083, PR China (H. Liu). Tel./fax: +86 731
88836320. microscope (TEM), a scanning electron microscope (SEM), and a
E-mail addresses: liuhuiqun@csu.edu.cn (H. Liu), yqchen1984@163.com three-dimensional microscope. The aim of this work is to reveal
(Y. Chen).

http://dx.doi.org/10.1016/j.ijfatigue.2015.11.011
0142-1123/ 2015 Elsevier Ltd. All rights reserved.
10 D. Yin et al. / International Journal of Fatigue 84 (2016) 916

the influence of grain size on the fatigue-crack-growth behaviour TEM and OM, as shown by the results in Figs. 2 and 3, respectively.
of 2524 aluminium alloy. The average measured grain sizes of the four samples are shown in
Table 2. The mean grain size of sample 1 was about 0.8 lm, which
2. Materials and experimental methods was associated with the low-density dislocation and without pre-
cipitates at the interior of the grains. The grain structures of sam-
A commercial hot-rolled 2524 alloy plate with thickness of ples 2 and 3 were relatively equiaxed; the measured grain size of
6 mm was provided by Southwest Aluminium Co. Ltd., China. Its sample 2 was 25, 16, and 29 lm in the longitudinal (L), short trans-
nominal composition was 4.2% Cu, 1.41% Mg, 0.56% Mn, 0.08% Fe, verse (S), and T directions, respectively; the grain size of sample 3
0.06% Si (in wt.%), and the balance was Al. The processing and was 59, 46, and 66 lm in the L, S, and T directions, respectively.
heat-treatment parameters of plates with four different grain sizes The grain structure of sample 4 was the most coarse, with grain
(sample 1, sample 2, sample 3, and sample 4) are listed in Table 1. sizes of 324, 298, and 345 lm in the L, S, and T directions, respec-
Tensile tests were carried out on a CSS-4400 testing machine tively. As shown by their tensile properties, listed in Table 3, the
along the transverse (T) direction. Fatigue-crack propagation and four samples obeyed the HallPetch relationship and clearly
crack-closure tests were carried out on an 250-kN fatigue machine revealed the effect of strengthening by grain refinement. For exam-
(8803, Instron, USA) using the middle-tension sample M(T) (see ple, sample 1, which had ultra-fine grains, had a higher ultimate
Fig. 1). Specimens were taken from alloy plates along the T direc- strength of 500 MPa, while the strength decreased for samples
tion and pre-cracked under mode-I. A sinusoidal loading was with larger grain sizes.
applied at a frequency of 10 Hz and load ratio of 0.1. The length
of the fatigue crack was measured by an optical microscope 3.2. Fatigue-crack-growth rate and fatigue-crack closure
(1 lm) attached on the fatigue machine. The following equation
was used to determine the stress intensity factor: The curves of the fatigue-crack-growth rate versus DK (the
r applied stress intensity factor range) for the four samples are plot-
DP pa pa
DK sec ; 1 ted in Fig. 4. Sample 1 exhibited higher fatigue-crack-growth rates
B 2w 2
with ultrafine grain size (0.8 lm). With increases in grain size, DKth
where P is the load (in MPa); B and w are the thickness (in cm) and increased and the fatigue-crack-growth rate decreased signifi-
width of the sample, respectively; a is the crack length (in cm); and cantly, especially in the region near DKth. When DK > 27 MPa m1/2,
a = 2a/w. the fatigue-crack-growth rates of samples 2, 3, and 4 were almost
The morphology of the fracture surface on each sample after the same. Since the grain size effect on fatigue crack propagation is
cyclical deformation was observed by a three-dimensional micro- more obvious under low DK, thus the following analyses focus on
scope (Kh7700, Hirox, Japan). Metallographic specimens were pre- the regime DK < 27 MPa m1/2.
pared in a standard procedure and etched in a solution consisting The crack-closure results of the four samples are shown in Fig. 5
of HF (2 mL), HCl (3 mL), HNO3 (5 mL), and H2O (200 mL). The in terms of DKeff/DK and Fig. 6 in terms of Kop/Kmax, where DKeff =
microstructure was observed by multifunction optical microscopy Kmax  Kop is the effective stress intensity factor, Kmax is the maxi-
and using an SEM (JSM-5600LV, JEOL, Japan). The microstructure of mum stress intensity level, and Kop is the stress intensity level
some samples was characterized by a TEM (TecnaiG2 20, FEI, USA) measured at the onset of closure. It can be seen that DKeff of sample
operating at 200 kV. Disks with diameters of 3 mm were cut from 1 had a higher value than those of the other three samples. The
fatigue-tested samples for TEM observation. These disks were first value of DKeff decreased with increases in grain size, which implies
ground to a thickness of 0.1 mm and then double-jet polished elec- that sample 1 with finer grains had higher fatigue-crack-growth
trolytically in a solution of 33% HNO3 and 67% methanol at 30 C. rate than those with coarser grains. This is attributed to the higher
closure level for larger grains at low DK levels (i.e. increasing DKeff/
DK and decreasing Kop/Kmax).
3. Results

3.1. Grain structure and tensile properties 3.3. Fatigue-crack fracture

The microstructures of the 2524 aluminium alloy samples with Three-dimensional reconstructions of the fracture-surface mor-
different grain sizes (samples 1, 2, 3 and 4) were characterized by phology of four samples at DK = 12 MPa m1/2 are shown in Fig. 7.
Line profile of these fracture surfaces were drawn according to
the method in ref. [23].
Table 1
The standard deviation of asperity heights (r0) is often used to
The processing and heat treatment routes of 2524 alloy plate samples.
quantify the surface roughness, which is given as followings [24]:
Samples The processing and heat treatment routes
 Pn 
Sample 1 6 mm hot rolled plate ? solid solution (500 C/1 h) ? cold i zi
rolling to 1.2 mm ? annealing (380 C/1 min) ? 12% pre-
r0 2
n1
deformation before natural aging for 96 h
Sample 2 6 mm hot rolled plate ? cold rolling to 3.6 mm ? solid
solution (500 C/30 min) ? cold rolling to 1.2 mm ? solid
wherein zi is the height of asperity and n is the total number of
solution (500 C/30 min) ? 12% pre-deformation before asperities.
natural aging for 96 h A statistical work on the asperity height was carried out and the
Sample 3 6 mm hot rolled plate ? cold rolling to 1.2 mm ? annealing standard deviations of asperity heights (r0) for four samples were
(330 C/30 min + 380 C/30 min + 420 C/30 min) ? solid
calculated to be 3.5 lm, 10.3 lm, 27.3 lm and 68.6 lm, respec-
solution (500 C/1 h) ? 12% pre-deformation before natural
aging for 96 h tively. It reveals that the fracture-surface roughness increases sig-
Sample 4 6 mm hot rolled plate ? eight passes cold rolling with 10% nificantly with the increase of grain size. r0 value of sample 4 is
reduction for each pass, and intermediate annealing (330 C/ almost 20 times greater than that of sample 1. But the fracture-
30 min + 380 C/30 min + 420 C/30 min) after each rolling surface roughness increases with the grain size is apparently
pass ? solid solution(500 C/1 h) ? 12% pre-deformation
before natural aging for 96 h
non-linear, since the grain size of sample 4 is about 300 times than
that of sample 1.
D. Yin et al. / International Journal of Fatigue 84 (2016) 916 11

Fig. 1. The fatigue crack propagation test sample (All dimensions in mm).

The scanning electron and optical micrographs of crack-


propagation profiles of four samples, observed along the S direc-
tion, are shown in Fig. 8. The crack-propagation path in sample 1
was smooth because of its ultra-fine grains (Fig. 8a). For sample
2, 3 and 4, although some intergranular growth was involved,
transgranular growth of the crack was predominant, as shown in
400nm Fig. 8(b)(d).
Since no precipitates were observed in sample 1, at the same
time, sample 2, 3 and 4 were T3 tempered, GP zone was the main
strengthening phase. The GP zone is believed to have less resis-
tance to dislocation slipping, and thus can hardly shorten slip
length. Therefore, as the crack propagates through a grain, it is
quite straight and scarcely deflected (Fig. 8(b)(d)). However, the
crack always deflects when it crosses grain boundaries. This is
probably due to the reason that the crack deviates to the most
favourable crystallographic orientation [25] (e.g., {1 1 1}Al) during
its propagation.
Moreover, closure of the fatigue crack occurred in sample 4, and
the crack path was more tortuous than that in the fine-grained
sample, as shown in Fig. 8(d). The crack closure in sample 4
showed a short periodic characteristic, especially at the position
Fig. 2. TEM image of sample1 showing fine grain size. of the bending crack.

(a) (b)

300m (c)

Fig. 3. Optical microstructure of (a) sample 2, (b) sample 3 and (c) sample 4.
12 D. Yin et al. / International Journal of Fatigue 84 (2016) 916

Table 2
Grain size (lm) of 2524 alloy plate samples.

Direction Sample 1 Sample 2 Sample 3 Sample 4


L 25 59 324
S 0.8 16 46 298
T 29 66 345

Table 3
Mechanical properties of 2524 alloy plate samples (T direction).

Samples Yield strength, r0.2 Tensile strength, rb Elongation, d


(MPa) (MPa) (%)
Sample 1 442 500 19
Sample 2 325 443 23
Sample 3 305 405 23
Sample 4 286 383 24

Fig. 6. Fatigue crack closure level (DK eff =DK) vs. applied DK of four samples at room
temperature, load ratio R = 0.1, frequency 10 Hz.

fatigue-crack growth and better damage tolerance. Thus, we can


conclude that grain refinement in an alloy serves to reduce the tor-
tuosity of the crack path, especially in the regime near DKth.
In general, the effective driving force required to propagate a
deflected crack is typically smaller than that for propagating a
straight crack at the same DK [16,26]. On the other hand, compared
with the straight crack, the deflected crack has a longer path when
propagating across the same distance, that is, the propagation rate
of the deflected crack is lower than that of the straight crack.
Additionally, the crack-closure effect further improves the appar-
ent driving force for the propagation of a deflected crack [8,16].
The collective effect of these factors was the reason the coarse-
grained sample in our study had greater resistance to fatigue frac-
ture than the fine-grained one.
Considering the effects of deflection and closure of periodically
Fig. 4. Fatigue crack growth rates vs. applied DK of four alloys at room temperature, deflected cracks, a crack-deflection model [16] was introduced to
load ratio R = 0.1, frequency 10 Hz. analyse the experimental results and predict the fatigue-crack-
growth rates. Schematic diagrams showing crack propagation
paths are shown in Fig. 9, where h is the kink angle, D is the
extended distance of tilted crack along the kink, and E is the
extended distance of the crack on the propagated plane and normal
to the tensile direction.
In Fig. 9, the local tensile opening (Mode I) and sliding (Mode II)
stress intensity factors for crack growth along the kink, kl and k2,
respectively, can be expressed as followings:
 
h
k1  cos3 KI 3
2

   
h h
k2  sin cos2 KI 4
2 2

wherein DKI is the nominal far-field stress intensity factor range.


Considering the deflections in crack path, the effective stress
intensity range (DKdef), in each segment (of span D + E) of this
non-planar crack, can be given by
Fig. 5. Fatigue crack closure level (Kop/Kmax) vs. applied DK of four samples at room 

temperature, load ratio R = 0.1, frequency 10 Hz. D cos2 2h E ~ DK I


DK def DK I D 5
DE
4. Discussion As well as known, for Mode I fatigue cracks are loaded in purely
uniaxial far-field cyclic tension, substantial Mode II displacements
In this study, a sample with coarser grain size showed crack develop around the crack tip if the crack faces are not flat [27,28]
deflections and had a higher level of crack closure. This implies that (Mode II displacements were measured to exceed Mode I displace-
the coarse-grained sample possessed higher resistance against ments near the crack tip in some investigations [29]). It could lead
D. Yin et al. / International Journal of Fatigue 84 (2016) 916 13

(a) Crack propagation direction (b)

Crack propagation direction


(c) (d)

Crack propagation direction


(e) (f)

Crack propagation direction


(g) (h)

Fig. 7. Three-dimensional metallographic photographs and surface line profiles of fatigue fracture surface morphology of four samples at DK 12 MPa  m1=2 (a) three-
dimensional metallographic photograph of sample 1, (b) line profile of sample 1, (c) three-dimensional metallographic photograph of sample 2, (d) line profile of sample 2, (e)
three-dimensional metallographic photograph of sample 3, (f) line profile of sample 3, (g) three-dimensional metallographic photograph of sample 4 and (h) line profile of
sample 4.
14 D. Yin et al. / International Journal of Fatigue 84 (2016) 916

(a)

(b)

(c)

(d)

200m

(e) (f) (g)

Fig. 8. SEM and metallographic photographs of fatigue crack profile for (a) sample 1, (b) sample 2, (c) sample 3, (d) sample 4, (e) crack closure in sample 4, (f) crack closure in
sample 4 and (g) crack closure in sample 4.

D uI
uII

(a) (b)
E

Fig. 9. Schematic diagrams of crack deflection (a) in fully opened condition (b) in premature closure condition led by mismatch between fracture surfaces.

the crack undergo a premature closure at positive (tensile) loads Combining Eqs. (5)(8), the effects of crack deflection, crack closure,
due to kinematically irreversible mismatch between two fracture and the ratio of ranges of global and local effective stress intensity
surfaces [30], as illustrated in Fig. 9(b). factor can be expressed as follows [8]:
uI and uII are defined as normal (Mode I) displacement and in-
plane (Mode II) displacement, respectively. The mismatch degree  !  1 !
DK eff D
cos2 2h 1 k tan h 2
(k) when the fracture surface begins to contact after unloading E
1 9
can be expressed as [16] DK D
E
1 1 k tan h

uII A statistical work on the topographical parameters (D/E and h)


k 6
uI of sample 2, 3 and 4 was performed based on the observation in
Fig. 7. The results are shown in Table 4. As can be seen, the value
In terms of stress intensity factor, the magnitude of the closure
of D/E increases obviously with the increase of grain size. This is
effect can be expressed as [16]
 1
k tan h 2
DK cl DK I 7 Table 4
1 k tan h Topographical parameters associated with alloy fracture surfaces.

Since DKdef could be further deduced by fracture surface mismatch, Samples Samples 2 Samples 3 Samples 4
the effective stress intensity (DKeff) becomes D/E 3.3 5.7 7.8
h 48.7 47.5 47.9
~ DK I  DK cl
DK eff D 8 k 0.048 0.068 0.481
D. Yin et al. / International Journal of Fatigue 84 (2016) 916 15

consistent with previous work [31] that larger grain size makes
more tortuous fatigue crack path. However, it can also be noted
that the grain size has little or no effect on the value of h. Since fati-
gue cracks are generally prone to propagate along slip bands [12],
it is reasonable to expect that the h value is more sensitive to grain
orientation rather than grain size.
If k is assumed to remain constant during the crack propagation,
its value can be calculated by the following equation:
8"  #2 9
1 < D
cos2 h
1 =
tan h 
E

2
  1 10
k : D cos2 h 1  eff D 1
DK ;
E 2 DK E

The calculated value of k for sample 2, 3 and 4 are shown in


Table 4. It shows that the k value increases apparently as the grain
size increases, which suggest that the crack-closure caused by frac-
ture surface mismatch is more evident for the alloy with coarser
grains.
Fig. 11. Fatigue crack growth rates predicted by the crack deflection model and the
Considering the extra length of the crack during deflection, the experimental results of sample 3. The crack of growth rate of sample 1 is assumed
straight-crack-growth rate (da/dn)L can be modified as [8] to be equal to (da/dn)L in Eq. (11).

  
da D cos h E da
11
dn DE dn L

Assuming that the crack-growth rate of the straight crack is


known and the value of k remains the same during the whole crack
propagation, we can then use Eqs. (9) and (11) to predict the crack-
growth rate as a function of DK for the deflection crack.
As mentioned above, the fatigue cracks of sample 1 were
approximately straight (in Fig. 8), and the crack-growth rate of
sample 1 can be taken as the growth rate of the straight crack,
(da/dn)L. Therefore, the crack-growth rates of sample 2, sample 3,
and sample 4 can be estimated.
Figs. 1012 show the crack-growth rates predicted by the crack-
deflection model. As can be seen in Figs. 10 and 11, when the stress
factors are at low DK levels, the calculated crack growth rates are
near between sample 2 and 3. But at high DK levels
(DK > 27 MPa m1/2), the predicted fatigue-crack-growth rates devi-
ate from the experimental data. This coincides with the results in
Fig. 4 which shows that the crack-growth rates are insensitive to Fig. 12. Fatigue crack growth rates predicted by the crack deflection model and the
the grain size when DK > 27 MPa m1/2. experimental results of sample 4. The crack of growth rate of sample 1 is assumed
However, as shown in Fig. 12, only when DK < 22 MPa m1/2, the to be equal to (da/dn)L in Eq. (11).
calculated fatigue-crack-growth rates are close to that of sample 4.
When DK < 22 MPa m1/2, the predicted fatigue-crack-growth rates
It is generally known that the growth behaviour of a crack is
deviate from the experimental data obviously.
closely related to the size of the plastic zone near crack tip
[32,33]. If the plastic zone size is large enough, the crack growth
behaviour will be not related to microstructure but to mechanical
properties of alloy [34]. Under uniaxial cyclic loading condition, the
plastic zone size (rp) can be expressed as [8]
 2
K
rp / 12
r0:2
As can be seen, rp increases with the increase of K, but decreases
with the increase of r0.2.
For a given r0.2, the fatigue-crack-growth rate is less sensitive to
the crack-closure at a high K level due to larger rp. This resulted the
gradual deviation of the predicted fatigue-crack-growth rates of
sample 2 and 3 from the experimental data when
DK > 27 MPa m1/2. Compared to sample 2 and 3, this deviation
appeared in sample 4 at an even lower DK level (i.e., 22 MPa m1/2).
The reason is that the smaller r0.2 value of sample 4 produced a lar-
ger rp and thus weakened the crack-closure effect.
Fig. 10. Fatigue crack growth rates predicted by the crack deflection model and the
Besides, it can be noticed that the measured fatigue-crack-
experimental results of sample 2. The crack growth rate of sample 1 is assumed to growth rates are obviously larger than the predicted ones when
be equal to (da/dn)L in Eq. (11). DK > 30 MPa m1/2. This could also be attributed to that the low
16 D. Yin et al. / International Journal of Fatigue 84 (2016) 916

strength of sample 4 accelerated the crack growth rate at the later [9] Chen YQ, Yi DQ, Jiang Y, Wang B, Liu HQ. Concurrent formation of two different
type precipitation-free zones during the initial stage of homogenization. Phil
stage of crack propagation, so, at such high DK levels the crack
Mag 2013;93:226978.
growth behaviour mainly depended on the mechanical properties [10] Agnew SR, Vinogradov AY, Hashimoto S, Weertman JR. Overview of fatigue
of alloy. performance of Cu processed by severe plastic deformation. J Electron Mater
1999;28:103844.
[11] Zhou MZ, Yi DQ, Liu HQ, Liu WJ, Zheng F. Enhanced fatigue crack propagation
5. Conclusions resistance of an AlCuMg alloy by artificial aging under influence of electric
field. Mater Sci Eng, A 2010;527:40705.
[12] Chen YQ, Pan SP, Zhou MZ, Yi DQ, Xu DZ, Xu YF. Effects of inclusions, grain
Plates of the 2524 aluminium alloy with significantly different boundaries and grain orientations on the fatigue crack initiation and
grain sizes were obtained by conventional processing methods, propagation behaviour of 2524T3 Al alloy. Mater Sci Eng, A 2013;580:1508.
and their fatigue crack growth rates as functions of grain size were [13] Mughrabi H, Hoppel HW. Cyclic deformation and fatigue properties of
ultrafine grain size materials: current status and some criteria for
investigated. The following conclusions were drawn.
improvement of the fatigue resistance. Mater Res Soc Symp Proc 2001;634.
Sample 1 with ultrafine grains (0.8 lm) had higher fatigue- B2.1.1B2.1.12.
crack-growth rates than other alloy samples with coarser grains. [14] Kim Y-W, Bidwell LR. Effects of microstructure and aging treatment on the
fatigue crack growth behaviour of high strength P/M aluminum alloy X7091.
As the grain size increased among the samples, DKth increased
In: High-strength powder metallurgy aluminum alloys: proceedings of a
and the fatigue-crack-growth rate decreased significantly, espe- symposium sponsored by the Powder Metallurgy Committee of the
cially in the region near DKth. The levels of crack closure in a Metallurgical Society of AIME, held at the 111th AIME Annual Meeting,
coarser-grained alloy were higher than that in a fine-grained alloy, Dallas, Texas, February 1718, 1982/edited by Michael J. Koczak, Gregory J.
Hildeman; p. 10724.
especially at low DK levels. The crack path in the coarse-grained [15] Chowdhury PB, Sehitoglu H, Rateick RG, Maier HJ. Modeling fatigue crack
alloy was more tortuous than that in the fine-grained alloy. The growth resistance of nanocrystalline alloys. Acta Mater 2013;61:253147.
crack deflection model was used to predict the crack-growth rate [16] Suresh S. Fatigue crack deflection and fracture surface contact:
micromechanical models. Metall Trans A 1985;16:24960.
as a function of DK for the deflection crack, and it provided a good [17] Hanlon T, Kwon Y-N, Suresh S. Grain size effects on the fatigue response of
explanation for the decrease in fatigue-crack-growth rate in the nanocrystalline metals. Scripta Mater 2003;49:67580.
coarse-grained sample. [18] Valiev RZ, Islamgaliev RK, Alexandrov IV. Bulk nanostructured materials from
severe plastic deformation. Prog Mater Sci 2000;45:10389.
[19] Plekhov O, Paggi M, Naimark O, Carpinteri A. A dimensional analysis
Acknowledgments interpretation to grain size and loading frequency dependencies of the Paris
and Wohler curves. Int J Fatigue 2011;33:47783.
[20] Zhou F, Rodriguez R, Lavernia EJ. Thermally stable nanocrystalline AlMg alloy
This research was supported by National Natural Science
powders produced by cryomilling. Mater Sci Forum 2003;386388:40914.
Foundation of China (Grant No. 51071177, and Grant No. [21] Han BQ, Lee Z, Nutt SR, Lavernia EJ, Mohamed FA. Mechanical properties of and
51405153), the Major State Basic Research Project of China (Grant ultrafine-grained Al7.5%Mg alloy. Metall Mater Trans A 2003;34:60313.
[22] Wang Y, Jiang S, Wang M, Wang S, Xiao TD, Strutt PR. Abrasive wear
No. 2012CB619506) and the Key Projects in the National Science
characteristics of plasma sprayed nanostructured alumina/titania coatings.
and Technology (No. 2014BAC03B05). The hot-rolled aluminium Wear 2000;237:17685.
alloy plate used in this experimental study was made available [23] Kamp N, Gao N, Starink MJ, Parry MR, Sinclair I. Analytical modelling of the
by Southwest Aluminium Co. Ltd. influence of local mixed mode displacements on roughness induced crack
closure. Int J Fatigue 2007;29:897908.
[24] Sehitoglu H, Garca AM. Contact of crack surfaces during fatigue: part 2.
References Simulations. Metall Mater Trans A 1997;28:227789.
[25] Sangida MD, Maierb HJ, Sehitoglu H. The role of grain boundaries on fatigue
[1] Srivatsan TS, Kolar D, Magnusen P. Influence of temperature on cyclic stress crack initiation an energy approach. Int J Plast 2011;27:80121.
response, strain resistance, and fracture behaviour of aluminum alloy 2524. [26] Cavaliere P. Fatigue properties and crack behaviour of ultra-fine and
Mater Sci Eng, A 2001;314:11830. nanocrystalline pure metals. Int J Fatigue 2009;31:147689.
[2] Chen YQ, Yi DQ, Jiang Y, Wang B, Xu DZ, Li SC. Twinning and orientation [27] Pokluda J. Dislocation-based model of plasticity and roughness-induced crack
relationships of T-phase precipitates in an Al matrix. J Mater Sci closure. Int J Fatigue 2013;46:3540.
2013;48:322531. [28] Pokluda J, Pippan R. Analysis of roughness-induced crack closure based on
[3] Maduro LP, Baptista CARP, Torres MAS, et al. Modeling the growth of LT and TL- asymmetric crack-wake plasticity and size ratio effect. Mater Sci Eng, A
oriented fatigue cracks in longitudinally and transversely pre-strained Al 2007;462:3558.
2524T3 alloy. Procedia Eng 2011;10:12149. [29] Davidson DL, Lankford J. Mixed-mode crack opening in fatigue. Mater Sci Eng
[4] Golden PJ, Grandt Jr AF, Bray GH. A comparison of fatigue crack formation at 1983;60:2259.
holes in 2024T3 and 2524T3 aluminum alloy specimens. Int J Fatigue [30] Ishihara S, Sugai Y, Mcevily AJ. On the distinction between plasticity- and
1999;21:S2119. roughness-induced fatigue crack closure. Metall Mater Trans A
[5] Baptista CARP, Adib AML, Torres MAS, Pastoukho VA. Describing fatigue crack 2012;43:308696.
growth and load ratio effects in Al 2524 T3 alloy with an enhanced exponential [31] Turnbull A, Rios ERDL. The effect of grain size on fatigue crack growth in an
model. Mech Mater 2012;51:6673. aluminum magnesium alloy. Fatigue Fract Eng Mater Struct 1995;18:135566.
[6] Srivatsana TS, Kolara D, Magnusen P. The cyclic fatigue and final fracture [32] Garca AM, Sehitoglu H. Contact of crack surfaces during fatigue: Part 1.
behaviour of aluminum alloy 2524. Mater Des 2002;23:12939. Formulation of the model. Metall Mater Trans A 1997;28:226375.
[7] Fonte MA, Stanzl-Tschegg SE, Holper B, Tschegg EK, Vasudevan AK. The [33] Kamp N, Parry MR, Singh KD, Sinclair I. Analytical and finite element modelling
microstructure and environment influence on fatigue crack growth in 7049 of roughness induced crack closure. Acta Mater 2004;52:34353.
aluminum alloy at different load ratios. Int J Fatigue 2001;23:S3117. [34] Gall K, Sehitoglu H, Kadioglu Y. FEM study of fatigue crack closure under
[8] Suresh S. Fatigue of materials. 2nd ed. Cambridge: Cambridge University Press; double slip. Acta Mater 1996;44:395565.
1998.

You might also like