Вы находитесь на странице: 1из 10

Available online at www.sciencedirect.

com

ScienceDirect
Acta Materialia 69 (2014) 236–245
www.elsevier.com/locate/actamat

In situ electron microscopy investigation of void healing


in an Al–Mg–Er alloy at a low temperature
M. Song a, K. Du a,⇑, S.P. Wen b, Z.R. Nie b, H.Q. Ye a
a
Shenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese Academy of Sciences, Shenyang 110016, People’s Republic
of China
b
School of Materials Science and Engineering, Beijing University of Technology, Beijing 100022, People’s Republic of China

Received 25 September 2013; received in revised form 10 January 2014; accepted 2 February 2014
Available online 28 February 2014

Abstract

Using in situ transmission electron microscopy and electron tomography, we have studied the healing process of submicron-scale
voids embedded in a cold-rolled Al–Mg–Er alloy. The results show that voids are healed successfully within 50 min at a relatively
low temperature of 453 K. Quantitative analysis of the in situ micrographs reveals that the void-healing process involves several stages:
an initial fast-healing stage, then a slow-healing stage, and finally a rapid healing near the end. The different healing rates are likely
related to varying surface curvatures due to the evolution of void geometry during the healing process. Because the voids are embedded
inside Al alloy grains, lattice diffusion is considered to dominate the entire healing process. Mg enrichment was observed at the healed
voids immediately after the healing. This indicates that the faster diffusion of Mg atoms in the Al matrix enhances void healing in the
Al–Mg–Er alloy. Post-mortem experiments verify that voids in bulk cold-rolled Al–Mg–Er alloy can also be healed by 1 h of annealing at
473 K, consistent with the in situ experiments. The fatigue crack growth resistance and plasticity of the cold-rolled Al alloy are improved
significantly after annealing at 473 K.
Ó 2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: In situ; Al–Mg–Er alloy; Void healing; Transmission electron microscopy (TEM); Electron tomography

1. Introduction materials properties and help develop materials processing


techniques.
Extensive studies [1–3] have revealed that void defects In recent years, studies on void healing have reported
usually cause stress concentration and crack nucleation significant progress, particularly in the areas of theoretical
around them under external loadings, which would greatly analysis and simulation. Key issues with void healing
shorten the incubation period of crack initiation. Voids involve the provision of material and energy. The main
also have adverse effects on fatigue crack propagation sources of energy include heating [9], electropulsing [10]
and crack closure, whereas most mechanical fractures in and external loading [11]. The sources of material primarily
engineering are reportedly associated with fatigue crack involve surface [12], lattice [13] and interface diffusion [14].
initiation and growth in metals [4]. Therefore, investiga- Simulations of pure metals have demonstrated that temper-
tions on the evolution of voids in materials have attracted ature has a significant effect on the process of void healing.
significant interest [5–8], and these investigations will facil- When the temperature is reduced, the void size reduction
itate an understanding of the recovery mechanism of rate decreases dramatically. Only for pure metals without
pre-existing dislocations can voids be healed at relatively
⇑ Corresponding author. Tel.: +86 24 8397 0725; fax: +86 24 2389 1320. low temperatures, such as 400 K for Cu nanoplates [15]
E-mail address: kuidu@imr.ac.cn (K. Du). and 673 K for body-centered cubic Fe crystals [16]. Based

http://dx.doi.org/10.1016/j.actamat.2014.02.004
1359-6454/Ó 2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
M. Song et al. / Acta Materialia 69 (2014) 236–245 237

on the difference in chemical potentials between a spherical properties were determined with a loading rate of
void surface and the grain boundary interface, Sun [13] has 2 mm min1. Fatigue crack growth rates were measured
simulated the shrinkage of a spherical void at the center of by using standard center crack tensile (CCT) samples.
a quasi-spherical grain dominated by lattice self-diffusion The test dimensions and notch lengths of the CCT samples
in pure Fe at high temperatures ranging from 973 to are 225 mm  60 mm  6 mm and 1.5 mm, respectively. A
1173 K for a-Fe and from 1200 to 1400 K for c-Fe and pre- sinusoidal load–time function was adopted, with the
dicted that only small, micron-sized spherical voids could stress ratio R set to 0.1 and the stress in the interval of
be wholly healed within hours at high temperature. A large 1.6  104–2.0  104 N. The oscillation frequency was
number of experiments on inorganic non-metallic materials 15 Hz.
have shown that void healing can occur under annealing at
high temperatures, ranging from 1073 to 1723 K [17–20]. In 2.3. Microstructure characterization
contrast, a relatively small number of experiments have
been performed on void healing in metals or alloys, and In situ heating experiments were conducted on a
these mainly involve post-mortem observations at high JEM2100 transmission electron microscope with a Gatan
temperatures [21]. Takahashi et al. [22] have analyzed the model 628 single-tilt heating holder at an accelerating volt-
shrinkage process of artificially introduced microscale age of 200 kV. A Tecnai G2 F30 transmission electron
cylindrical holes in Cu at high temperatures by post-mor- microscope, equipped with a high-angle annular-dark-field
tem scanning electron microscopy (SEM) observations. (HAADF) detector and an X-ray energy-dispersive spec-
They investigated void junctions by SEM and concluded trometry (EDS) system, was used at 300 kV for bright-field
that voids with different morphologies could be predicted imaging, HAADF scanning transmission electron micros-
by suitable models. Nevertheless, because of the small size copy (STEM) imaging, and composition analysis. The elec-
(<10 lm) of voids and the surface imaging nature of SEM tron tomography was performed on the Tecnai G2 F30
techniques, it is difficult to observe the void-shrinkage pro- equipped with a Fishione 2040 tomography sample holder,
cess directly by SEM. Therefore, few experimental in situ and the data was acquired using Inspect3D software for
studies on void healing have been reported. In this study, STEM tomography. 3-D reconstruction and volume analy-
in situ heating transmission electron microscopy (TEM) sis were performed with Amira 5.0.1 software (Visage Imag-
has been employed to study the void-healing process in a ing, Berlin). During the experiment, the sample was tilted
cold-rolled Al–Mg–Er alloy [23] at low temperatures. Via from –60° to +60° in steps of 2°. TEM thin-foil specimens
in situ heating at 453 K we have observed the healing pro- parallel to the rolling plane were prepared by cutting, grind-
cess of voids embedded in an Al–Mg–Er alloy in real time ing, polishing and ion-milling. A SUPRA35 field emission
and analyzed the healing kinetics accordingly. scanning electron microscope, operated at 20 kV, was used
to investigate the fracture surface of fatigue cracks.
2. Experimental
3. Results
2.1. Sample preparation
3.1. Microstructure of the cold-rolled Al alloy
The experimental Al–Mg–Er alloys were produced by
the semicontinuous chill-casting method. The as-cast alloys Bright-field TEM images show the distribution of gray
were homogenized at 743 K for 24 h to eliminate interden- lath-like precipitates and “white features” between them
dritic segregation, followed by hot-rolling at 743 K, anneal- in the present cold-rolled Al–Mg–Er alloy (Fig. 1a). Based
ing at 453 K for 2 h and cold-rolling into sheets of 6 mm on a large number of EDS analyses (more than 20), the
thickness. The thickness reductions brought about by the chemical composition of the lath-like precipitates was
hot-rolling and cold-rolling processes were 92% and 67%, revealed to be Al and Mn with an atomic percentage ratio
respectively. The chemical composition of the alloy was of Al:Mn of approximately 6:1. Hence, the lath-like precip-
analyzed by inductively coupled plasma spectroscopy as itates are presumably Al6Mn and will be described as such
5.80% Mg, 0.67% Mn, 0.31% Er, 0.10% Zr (wt.%), and bal- hereinafter. Between the Al6Mn precipitates, the “white
ance was Al. To compare with in situ heating TEM sam- features” have a projected dimension of about
ples, annealed (CA) Al alloy samples were prepared from 400 nm  50 nm. A large number of TEM images revealed
the cold-rolled (CR) alloy by annealing at 473 K for differ- that the “white features” were located only between Al6Mn
ent durations (10, 20, 30, 40, 50 and 60 min, respectively). precipitates and had projected dimensions of the order of
tens of nanometers to submicrometers. HAADF STEM
2.2. Performance tests imaging shows that the “white features” in bright-field
TEM images turn into areas of dark contrast in HAADF
All mechanical tests of the CR and the CA alloys were STEM images (Fig. 1b). Fig. 1c shows the intensity profile
carried out under the same experimental conditions in an between the two arrows in the HAADF STEM image
ambient environment. The tensile and fatigue specimens (Fig. 1b). The intensity profile shows a “deep well” charac-
were machined parallel to the rolling direction. Tensile teristic, as distinguished from a “shallow dish” feature. In
238 M. Song et al. / Acta Materialia 69 (2014) 236–245

Fig. 1. Microstructure of the CR alloy. (a) A bright-field TEM image of the CR alloy showing the gray lath-like precipitates and “white features” between
them. (b) An HAADF STEM image of the CR alloy showing bright-contrast Al6Mn precipitates and voids between them. (c) Image intensity profile
between the two arrows in (b).

addition, the EDS counts generated from the “white fea- surface renders of the void viewed from different directions
tures” are far below those of the surrounding Al matrix. by rotating the sample (see also Supplementary Movie S1).
Based on the above observations, the “white features” When the tilting angle is 0°, the projected void shows a
should correspond to voids embedded in the CR alloy. dog-bone shape with a length of 180 nm and a projected
Because Al6Mn precipitates are nondeformable particles width varying from 30 nm at the center to 60 nm at the
[24], which are much harder than the surrounding Al ends (Fig. 2d). With an increase in the tilting angle, the pro-
matrix [25,26], it is easy for Al6Mn precipitates to be bro- jected width increases up to 110–180 nm (Fig. 2f). It is clear
ken during the cold-rolling deformation. When the broken from the 3-D surface renders that there are a considerable
fragments move along with the Al matrix, voids would be number of surface sites with a small curvature radius,
formed between them [27]. which is related to void healing, as discussed below.
The 3-D geometry of voids was determined by electron
tomography [28] (Fig. 2). Fig. 2c shows 3-D reconstructed 3.2. In situ heating experiment
surface renders of the void (gold) and the Al6Mn precipi-
tates (purple) on two sides of the void; everything is embed- To investigate the evolution of voids embedded in the
ded in an electron-transparent Al matrix. Fig. 2d–f shows CR alloy during an annealing treatment, in situ heating

Fig. 2. HAADF STEM images and surface renders of a void between Al6Mn precipitates embedded in the Al matrix. (a) HAADF STEM image of the
void with 0° tilting. (b) HAADF STEM image of the void with 50° tilting by rotating clockwise around the arrow indicated in (a). (c) Electron
tomography image of the void (gold) and the Al6Mn precipitates (purple) at two sides of the void; everything is embedded in an electron-transparent Al
matrix. (d–f) Surface renders of the void, obtained by electron tomography, shown by rotating counterclockwise around the x-axis indicated in (c). (For
interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)
M. Song et al. / Acta Materialia 69 (2014) 236–245 239

TEM has been employed to study the void-healing process. needed to complete the healing (the mean healing time is
Voids embedded inside Al grains that were 2–3 lm away 2205 s with a standard deviation of 750 s) varied consider-
from the sample perforation were chosen for the in situ ably for these voids. The scatterings of the initial void size
heating experiment at 453 K so that this experiment could and the healing time are possibly due to the varied size and
reflect the void evolution in the bulk material. Fig. 3 shows orientation of Al6Mn precipitates in Al alloy and the differ-
snapshots of a void during the in situ heating. During the ences in fracture surfaces of the Al6Mn precipitates after
initial stage of this experiment, it is obvious that voids heal cold rolling, as these determine the geometry and size of
quickly at sites with a smaller surface curvature radius the voids. In order to clearly illustrate the healing kinetics,
(Fig. 3a and b). During this stage, small isolated voids the void projected area and the healing duration have been
are formed at the necking region of the original voids normalized based on the initial size (denoted by the initial
and they disappear after a short time (Fig. 3a–c), behavior projected area here) of the voids and the heating duration
that is similar to the characteristics described by Gupta for needed for completing the healing, respectively, and then
the observation of cylindrical voids in polycrystalline averaged and plotted as an AN–tN curve in Fig. 4b. Here,
alumina [29]. Then, the healing rate slows down. After error bars indicate the standard deviation of normalized
the voids reach a certain size, the speed of healing projected areas at different heating times. The time deriva-
increases, and the voids are healed instantaneously tive of the normalized projected area (inset of Fig. 4b)
(Fig. 3e–h, see also Supplementary Movie S2). shows that the AN–tN curve can be divided into three dis-
tinct stages: an initial fast-healing stage (from s0 to s1), then
3.3. Variation of the projected area and volume of voids with a slow-healing stage (from s1 to s2), and finally a rapid-
the annealing time healing stage (from s2 to s3). Table 1 summarizes the pro-
jected area and heating duration of the voids at the begin-
Due to necking, the void in Fig. 3a was divided into two ning and the end of the void healing, as well as two nodes
parts, A and B (see also Fig. 4a). For each part, Fig. 4a for the three distinct healing stages, respectively.
shows how the projected areas A of voids A and B vary In order to reveal the change in the volume properties of
with the heating time. The curves reveal that the void-heal- voids as heating progresses, the in situ heating experiment
ing rate varies significantly in different periods. At the has also been combined with electron tomography charac-
beginning of the healing process, the healing rate is high; terization. When a void in a TEM specimen reached a cer-
then it slows down; but, towards the end of healing, the tain healing stage during the in situ heating experiment, the
healing rate increases significantly. The evolution of void specimen was transferred to the tomography specimen
projected area with heating duration was also measured holder and characterized there by electron tomography.
for 20 voids by in situ heating experiments. Table 1 shows In this way, the projected area and volume of the void
that the initial size (the mean projected area is 2716 nm2 were measured from HAADF STEM and reconstructed
with a standard deviation of 894 nm2) as well as the time 3-D images, respectively. After the electron tomography

Fig. 3. TEM images showing the healing evolution of voids embedded in the CR alloy during the in situ heating experiment.
240 M. Song et al. / Acta Materialia 69 (2014) 236–245

Fig. 4. (a) The void projected area as a function of time during the in situ heating experiment. (b) The normalized projected area (AN) as a function of the
normalized heating duration (tN) measured from 20 voids by in situ heating experiments. (Inset) the derivative of the normalized projected area with
respect to the normalized heating duration. s1 and s2 are the ending times of the first and the second stages, respectively.

Table 1 in healing rate between the second and third stages is less
Mean values and standard deviations of void projected areas and heating significant in the volume–heating duration curves com-
durations measured from 20 voids by in situ heating experiments. The
pared to the projected area. It should be noted that,
values correspond to nodes between three distinct healing stages (s1, s2), as
well as the beginning (s0) and the end (s3) of void healing process. t, because heating and cooling procedures were needed when
heating duration; rt, standard deviation of the heating duration; A, void transferring the specimen between the in situ heating pro-
projected area; rA, standard deviation of the projected area. cess and electron tomography characterization, the actual
Normalized heating duration Time (s) Projected area (nm2) heating duration the TEM specimen experienced deviated
tN t ± rt A ± rA slightly from the nominal heating duration, especially in
s0 = 0 0 2716 ± 894 the later healing stages.
s1 = 0.26 565 ± 192 1914 ± 605
s2 = 0.92 2036 ± 693 664 ± 207 3.4. Composition analysis
s3 = 1 2205 ± 750 0

After 40 min of heating at 453 K, bright-field TEM and


HAADF STEM imaging show that voids are healed suc-
observation, the TEM specimen was transferred back to cessfully, but EDS elemental maps indicate an inhomoge-
the heating holder for further in situ testing. Hence, the neous distribution of Al and Mg (Fig. 7a and d).
variations of void projected area and volume were determined Compared with the Al alloy matrix, there is enrichment
for different heating durations. Fig. 5 shows that the void in Mg and deficiency in Al in the regions where the voids
was divided into two parts, C and D, after 2 min of used to be. With prolonged annealing, the segregation
in situ heating at 453 K. Electron tomography images decreases (Fig. 7b and e). After 2 h of annealing at
(Fig. 5b–d, f–h, j–l, n–p, r–t and v–x) present the evolution 453 K, both Al and Mg are distributed homogeneously
of the 3-D geometries of voids C and D with heating time. (Fig. 7c and f).
According to the variation of the projected area with heat-
ing time (Fig. 6), Fig. 5a would be roughly associated with 3.5. Post-mortem experiment
the first healing stage, and Fig. 5e, i and m with the second
stage for void C. Similarly, for void D, Fig. 5a and b would The annealed bulk Al–Mg–Er alloys, which were
be associated with the first healing stage, and Fig. 5i, m and obtained by annealing the CR alloy at 473 K for different
q with the second stage. The change in void thickness durations, have also been investigated. HAADF STEM
(Fig. 5h, l and p) along the STEM viewing direction is images of the CR alloy (e.g. Fig. 8a) show that dog-
insignificant over 40 min of heating. Fig. 5q–u corresponds bone-shaped voids were located between Al6Mn precipi-
to the third stage, where void D changed from ellipsoidal to tates. Annealing for 20 min (Fig. 8b and c) caused the void
spherical in shape with a rapid decrease in void thickness. regions attached to Al6Mn precipitates and with small radii
Fig. 6 shows the variation of void volume with the heat- of curvature to shrink, and also healed the narrow necking
ing time. In agreement with the projected area, the void regions of voids, which also had small radii of surface cur-
volume clearly shows an initial fast-healing stage and then vature. The number of voids as well as their size (denoted
a slow-healing stage. During the final rapid-healing stage, by the projected area) decrease as the heating duration
the volume change accelerated, particularly when (e.g. Fig. 8d–f) is extended from 30 to 50 min. After
approaching the end of void healing, although the change 60 min of annealing, the HAADF STEM images (e.g.
M. Song et al. / Acta Materialia 69 (2014) 236–245 241

Fig. 5. HAADF STEM images and surface renders of voids obtained at different in situ heating durations at 453 K. (a, e, i, m, q, u) HAADF STEM
images of the voids obtained at 2, 5, 13, 40, 67 and 75 min of heating, respectively. (b, f, j, n, r, v), (c, g, k, o, s w) and (d, h, l, p, t, x) Surface renders of
voids C and D (indicated in (a)), at different heating durations, shown by rotating counterclockwise around the x-axis indicated in (b) for 0°, 45° and 90°,
respectively.

Fig. 8g) show no voids between Al6Mn precipitates. The


post-mortem experiment results indicate that voids in the
bulk CR Al–Mg–Er alloy can also be healed after 1 h of
low-temperature annealing. The healing process is consis-
tent with the in situ heating of thin-foil samples (e.g.
Fig. 3a–h), while the void geometries during the annealing
process are qualitatively consistent with those of the in situ
heating experiments.

3.6. Mechanical properties of the CR and CA alloys

The mechanical properties and fatigue crack growth rate


were characterized for the CR and CA alloys; all tests were
performed under the same experimental conditions. The
results are summarized in Table 2. The tensile strength of
the CR alloy is higher than that of the CA alloy, whereas,
Fig. 6. The projected area and volume of voids C and D indicated in as expected, the fracture elongation of the CR alloy is
Fig. 5a determined from HAADF STEM and reconstructed 3-D images, much less than that of the CA alloy. Nevertheless, the fati-
respectively, as functions of time during the in situ heating experiment. gue crack growth rates da/dN of the CA alloy are much
242 M. Song et al. / Acta Materialia 69 (2014) 236–245

Fig. 7. The enrichment of Mg at the healed voids. (a–c) EDS maps of Al (right) and Mg (middle), respectively, obtained from the respective areas of the
HAADF images (left). (d–f) Al and Mg EDS line-scan profiles scanned along the red dashed line of the HAADF images in (a–c), respectively.

Fig. 8. HAADF STEM images of the annealed bulk Al–Mg–Er alloys after different annealing times.

Table 2
Mechanical and fatigue properties of the CR and the CA alloys obtained at room temperature. DK, stress intensity factor range.
Specimen Tensile strength (Mpa) Fracture elongation (%) Fatigue crack growth rate da/dN (mm/cycle)
DK = 6.0 Mpa m1/2 DK = 10.0 Mpa m1/2 DK = 20.0 Mpa m1/2
5 4
CR 515 9.0 4.5  10 2.6  10 1.8  103
CA 445 16.5 2.4  105 1.8  104 0.8  103

lower under the same stress intensity factor range DK. The CR alloy and the CA alloy taken from corresponding loca-
results indicate that the plasticity of the CR alloy has been tions. The results suggest that the fracture surface of the
improved significantly after annealing for 1 h at 473 K. CR alloy is rougher due to the large local stress concentra-
Fig. 9a and b shows the fatigue fracture surfaces of the tion in the CR alloy.
M. Song et al. / Acta Materialia 69 (2014) 236–245 243

Fig. 9. SEM micrographs of fatigue fracture surfaces of the CR and the CA alloys taken from the corresponding locations.

4. Discussion

As described above, the void-healing process can be


divided into several apparent stages that are closely related
to the variation of the surface morphology of the voids
during the annealing. The results of the in situ experiment
indicate that voids tend to heal first at the corners with a
smaller surface curvature radius in the initial fast-healing
stage. This can be explained, approximately, by the con-
verted GibbsThompson equation [30]: Fig. 10. Schematic model illustrating the evolution of void healing during
lr ¼ lb þ 2rV m =rRT : ð1Þ the in situ experiment.

Here, lr and lb, represent the chemical potentials of voids d correspond to the initial fast-healing stage and the slow-
with a radius r and of infinite size, respectively; r is the spe- healing stage, respectively. Fig. 10d, e and e, f show the
cific surface energy; Vm is the molar volume of the material; final rapid-healing stages of voids A and B.
r is the radius of a void; R is the gas constant; and T is the Electron tomography combined with in situ heating
absolute temperature. Therefore, the voided regions with a confirms that the variation in void volume during the heal-
smaller curvature radius have a larger driving force for dif- ing process can be divided into different stages, most signif-
fusion and, consequently, a greater healing rate. The nar- icantly the initial fast-healing stage and the slow-healing
row neck of the void, which also has a small surface stage. This is consistent with the relationship revealed by
curvature radius, shrinks and is divided into several small the variation in projected area with heating time under con-
voids, and these small voids soon disappear. As a result, tinuous in situ heating observations. Although the distinc-
the void evolves into two ellipsoid-like parts, where each tion between the second and third stages is not so clear
part is attached to an Al6Mn precipitate. Therefore, the from the void volume, a tendency to accelerated healing
remaining curvature radii r are the larger ones for the void is obvious near the end of void healing (e.g. Fig. 6). This
surface, and the curvature radii of the ellipsoid-like voids is likely due to the small surface curvature radii of the voids
are more uniform than the initially dog-bone-shaped void. in this phase, which increase the driving force for diffusion.
Consequently, the healing process enters the slow-healing The difficulty of identifying the third healing stage is also
stage. As the ellipsoid-like voids continually shrink during attributed to the experimental setup. As previously dis-
the second stage, the average curvature radius r decreases cussed, cooling and heating processes occur when the spec-
with time t. The reason for the significant increase in the imen is transferred from the in situ heating experiment to
healing rate is associated with the small size of the voids the electron tomography observation. This would influence
entering the rapid-healing phase. Because of the small ra- the actual heating experienced by the voids being observed,
dius of curvature in this phase, the transport of materials thus creating error in the estimated heating durations for
is significantly increased per unit area at the same time. void volume–heating time curves. This uncertainty over
The thickness along the viewing direction decreases the time-scale assessment is particularly critical for the
quickly, while the ellipsoidal voids become spherical in determination of the third healing stage, due to its short
shape. Meanwhile, because the average curvature radius r duration (3 min, as shown in Table 1). In this regard, fur-
decreases with time t in this phase, the driving force for dif- ther research with in situ 3-D observation would be helpful
fusion increases. Fig. 10 shows a schematic of the evolution when electron tomography processing has been developed
of voids during the in situ experiment. Fig. 10a–c and c and to be fast enough to capture void evolution in real time.
244 M. Song et al. / Acta Materialia 69 (2014) 236–245

significant extent to the large number of voids in the CR


alloy. Owing to the local stress concentration around voids
in the CR alloy, cracks are likely to nucleate more easily
under external loadings. Meanwhile, voids have adverse
effects on the fatigue crack propagation path or crack clo-
sure. Therefore, successful healing of voids after annealing
at 473 K for 1 h results in a significant improvement in the
performance of the CA alloy.

5. Conclusions

The alloy studied here has a large number of voids


between the Al6Mn precipitates; these voids are embedded
in the matrix of the CR alloy, with the projected dimen-
sions from tens of nanometers to the submicrometer scale.
Fig. 11. The diffusivity of diffusants in the Al matrix from data given in The voids can be healed successfully after 50 min of in situ
Refs. [33–36]. heating at 453 K. The progress of void healing can be cat-
egorized into several apparent stages: an initial fast-healing
Despite the fact that there are significant differences in stage, a slow-healing stage and a rapid-healing near the
the healing rates of the different stages, the overall evolu- end. The void-healing process is completed by a combina-
tion of void healing involves lattice, interface and surface tion of lattice, interface and surface diffusions, and is con-
diffusion but presumably is controlled by the lattice diffu- trolled by the lattice diffusion.
sion. During the process of void healing, grain boundaries The filled void areas were deficient in Al and enriched in
may serve as an inexhaustible source of atoms and a sink Mg compared with the surrounding matrix, but after 2 h of
for vacancies decomposed by voids, with the driving force in situ heating at 453 K, the Al and Mg were distributed
provided by the difference between the chemical potentials homogeneously in the matrix. The enrichment of Mg is
of the grain boundary interface and the internal surface of related to the higher diffusion rate of Mg in the Al matrix,
voids [13]. The transportation of materials is controlled by which enhances the void healing in the Al alloy and is par-
lattice diffusion. The role of internal surface diffusion is to ticularly important for void healing at low temperatures.
keep the spherical void shape unchanged during the healing Post-mortem experimental results indicate that voids in
process [31]. Based on the above analysis, the voids embed- the CR alloy can be healed by an annealing treatment at
ded inside grains can only be eliminated by lattice diffusion 473 K for 1 h, consistent with the in situ experiments.
because of the transportation of atoms and vacancies. Fur- Voids have adverse effects on the fatigue performance of
thermore, lattice diffusion is much slower than interface the CR alloy, whereas the plasticity of the CR alloy can
diffusion [32]. Hence, the overall evolution of voids in the be improved significantly after annealing at 473 K for 1 h.
present research is controlled by lattice diffusion.
After 40 min of in situ heating at 453 K, the elemental Acknowledgements
maps indicate that the filled void areas are deficient in Al
and enriched with Mg compared with the surrounding The authors acknowledge financial support from the
matrix; This inhomogeneous distribution can be eliminated Special Funds for the Major State Basic Research Projects
by increasing the in situ heating time to 2 h. Fig. 11 shows of China (Grant No. 2012CB619503) and the Natural Sci-
the diffusion rate of diffusants in the Al matrix. Previous ences Foundation of China (Grant Nos. 51171188,
experimental data [33–36] indicate that there is a linear 51390473, 51221264, and 11332010). The authors are also
relationship between ln D and 1/T for Mg in an Al matrix grateful to J.K. Qiu, H. Xu, J. Wang, J.P. Cui, Y.T. Zhou,
between the temperatures of 573 and 928 K. When this lin- J.F. Liu, N. Lu and Y.J. Xu for stimulating discussions,
ear relationship is extrapolated to 453 K, the diffusion rate and to X.L. Liu for providing the alloy samples.
of Mg in an Al matrix is revealed to be much higher
(6 times) than the self-diffusion of Al atoms. Therefore, Appendix A. Supplementary material
the enriching of Mg in the healed voids is associated with
the higher diffusion rate of Mg compared with Al in the Supplementary data associated with this article can be
Al matrix. This indicates that the faster diffusion of Mg found, in the online version, at http://dx.doi.org/10.1016/
atoms in the Al matrix enhances the void healing in the j.actamat.2014.02.004.
Al alloy, which is particularly important at low
temperatures. References
It is noteworthy that compared with the CA alloy, both
the higher fatigue crack growth rates and greater roughness [1] Lados DA, Apelian D. Mater Sci Eng A 2004;385:187.
of the fatigue fracture surface are likely related to a [2] Li P, Maijer DM, Lindley TC, Lee PD. Mater Sci Eng A 2007;460:20.
M. Song et al. / Acta Materialia 69 (2014) 236–245 245

[3] Caton MJ, Jones JW, Boileau JM, Allison JE. Metall Mater Trans A [20] Liu SP, Ando K, Kim BS, Takahashi K. Int Commun Heat Mass
1999;30:3055. 2009;36:563.
[4] Suresh S. Fatigue of materials. 2nd ed. Cambridge: Cambridge [21] Brett J, Seigle L. Acta Metall 1963;11:467.
University Press; 1998. [22] Takahashi Y, Ueno F, Nishiguchi K. Acta Metall 1988;36:3007.
[5] Petrov AI, Razuvaeva MV. Phys Solid State 2007;45:907. [23] Meng G, Li BL, Li HM, Huang H, Nie ZR. Mater Sci Eng A
[6] Salac D, Lu W. Int J Solids Struct 2008;45:3793. 2009;516:131.
[7] Nosonovsky M, Amano R, Luccia JM, Rohatgib K. Phys Chem [24] Bae DH, Ghosh AK. Acta Mater 2002;50:511.
Chem Phys 2009;11:9530. [25] Cao B, Joshib SP, Ramesh KT. Scr Mater 2009;60:619.
[8] Gittins A. Acta Metall 1968;16:517. [26] Dai LH, Ling Z, Bai YL. Compos Sci Technol 2001;61:1057.
[9] Takahashi Y, Inoue K, Nishiguchi K. Acta Metall Mater [27] Azuma M. Structural control of void formation in dual phase steels,
1993;41:3077. PhD thesis, Technical University of Denmark; 2013.
[10] Qin RS, Su SX. J Mater Res 2002;17:2048. [28] Midgley PA, Ward EPW, Hungrı́a AB, Thomas JM. Chem Soc Rev
[11] Beere WB, Greenwood GB. Met Sci 1971;5:107. 2007;36:1477.
[12] Makin MJ. J Nucl Mater 1978;71:300. [29] Gupta TK. J Am Ceram Soc 1978;61:191.
[13] Sun J. J Mater Eng Perform 2002;11:322. [30] Peng ZA, Peng XG. J Am Chem Soc 2002;124:3343.
[14] Ma Q. Acta Metall 1998;46:1669. [31] Zhang HL, Sun J. Mater Sci Eng A 2004;382:171.
[15] Wang MF, Du GJ, Xia DY. Key Eng Mater 2013;531:454. [32] Yu HH, Suo Z. J Mech Phys Solids 1999;47:1131.
[16] Wei DB, Han JT, Tieu K, Jiang ZY. Scr Mater 2004;51:583. [33] Volin TE, Balluffi RW. Phys Status Solidi 1968;25:163.
[17] Choi SR, Tikare V. Mater Sci Eng A 1993;171:77. [34] Dais S, Messer R, Seeger A. Mater Sci Forum 1987;15:419.
[18] Ando K, Houjyou K, Chu MC, Takeshita S, Takahashi K, Sakamoto [35] Fujita T, Horita Z. Philos Mag A 2002;82:2249.
S, et al. J Eur Ceram Soc 2002;22:1339. [36] Rothman SJ, Peterson NL, Nowicki LJ, Robinson LC. Phys Status
[19] Nam KW, Kim JS. Mater Sci Eng A 2010;527:3236. Solidi B 1974;63:29.

Вам также может понравиться