You are on page 1of 12

Available online at www.sciencedirect.

com

Acta Materialia 56 (2008) 230–241


www.elsevier.com/locate/actamat

Microstructural evolution and nanostructure formation


in copper during dynamic plastic deformation
at cryogenic temperatures
Y.S. Li, N.R. Tao *, K. Lu *

Shenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese Academy of Sciences, Shenyang 110016, China

Received 1 May 2007; received in revised form 10 September 2007; accepted 12 September 2007
Available online 5 November 2007

Abstract

The microstructural evolution and formation mechanism of nanostructures in bulk pure Cu samples induced by dynamic plastic
deformation (DPD) at high strain rates and cryogenic temperatures were investigated using transmission electron microscopic charac-
terization. Three different mechanisms were identified for the plastic deformation and microstructural refinement, including dislocation
manipulation and rearrangement, deformation twinning forming nanoscale twin/matrix (T/M) lamellae in bundles, and shear banding in
the T/M lamellae. An increasing tendency of deformation twinning and shear banding was observed at higher strains. For strain e = 2.1,
a mixed nanostructure is formed in the DPD Cu bulk sample with nanoscale T/M lamellae making up about 33% of the volume and
nano-sized grains making up about 67%. The nanograins can be classified into three types in terms of their origin: (i) nanograins derived
from fragmentation of nanoscale T/M lamellae with an average transverse size of about 47 nm; (ii) nanograins in shear bands with an
average transverse size of about 75 nm; and (iii) nanograins derived from dislocation cells with an average transverse size of about
121 nm. The high density of deformation twins induced by high strain rates and cryogenic temperatures in DPD, distinct from that
in conventional severe plastic deformation, plays a crucial role in formation of the nano-sized grains.
 2007 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Nanostructured materials; Copper; Dynamic plastic deformation; Deformation twinning; Shear banding

1. Introduction are other processes known to produce true nanostructured


pure metals based on plastic straining, such as ball milling
Owing to the growing interest in nanostructured materi- (BM) [6], friction deformation under large sliding loads [7],
als in recent years, numerous synthesis techniques have surface mechanical attrition treatment (SMAT) [8] and
been proposed [1]. Severe plastic deformation (SPD) pro- cryogenic rolling [9]. However, these processes are either
cesses (e.g. equal channel angular pressing (ECAP) and inadequate for producing bulk samples and/or time-
high-pressure torsion (HPT) [2–4]) are arguably the most consuming and costly. In short, difficulties in fabricating
promising methods for producing bulk ultrafine-grained bulk nanostructured samples expediently are a major road-
metals while avoiding the consolidation step that often block for research and development of nanostructured
brings artifacts into the samples [5]. However, the grain metals and alloys [10].
sizes achieved so far through these routes for pure metals Recently, we have developed a new technique, dynamic
such as Cu are usually outside the true nanoregime, plastic deformation (DPD), i.e. plastic deformation at high
i.e. are of the order of several hundred nanometers. There strain rates [11]. After the DPD process, high-density
nanoscale deformation twins together with nano-sized
*
Corresponding authors. Tel.: +86 24 2390 6826; fax: +86 24 2399 8660. grains are introduced into copper. The nanostructured
E-mail addresses: nrtao@imr.ac.cn (N.R. Tao), lu@imr.ac.cn (K. Lu). copper has a yield strength of 600 MPa, while keeping

1359-6454/$30.00  2007 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2007.09.020
Y.S. Li et al. / Acta Materialia 56 (2008) 230–241 231

some plasticity. It is imperative to understand the micro- a high loading rate. The load and loading rate are con-
structure evolution and formation of nanostructures in trolled automatically by a computer. Multiple impacts were
copper during a DPD process from a scientific and techno- applied to deform a cylinder sample eventually to a disc of
logical point of view. about 1.5 mm thick. The strain rate applied in each impact
Strain-induced microstructural evolution has been exten- is estimated to be within 1 · 102–2 · 103 s1. The treated
sively investigated in numerous materials [6–10,12]. It has sample was cooled by liquid nitrogen before each impact
been well documented that for those materials with medium and the sample temperature was estimated to be below
to high stacking fault energies (SFEs), such as Fe, Cu and 100 C during each impact.
Al, coarse grains are refined upon continued straining by The deformation strain is defined as e = ln (L0/Lf),
various dislocation activities at low strain rates and at ambi- where L0 and Lf are the initial and final thickness of the
ent temperatures. On the other hand, deformation twinning deformed sample, respectively. Measurements showed that
is found to be an effective grain refinement mode in many the Cu samples after DPD treatments are identical in den-
low SFE face-centered cubic (fcc) materials [13–15]. Never- sity and purity to the original sample. No flaws (cracks or
theless, when pure Cu with a medium SFE is deformed at porosities) have been detected inside the DPD samples. The
cryogenic temperatures and/or at high strain rates, defor- microstructure is rather uniform in the DPD samples both
mation twinning may play an important role in accommo- in radius direction and along the thickness.
dating plastic strain as dislocation activity is suppressed
[16–19]. However, little work has been done to study the 2.3. Microstructure characterization
microstructural evolution in copper during high-strain-rate
deformation at cryogenic temperatures before. An optical microscope (OM, Leica MPS 30) was used to
The formation process of nanostructures in copper has observe the holistic microstructure of the deformed Cu
been investigated several times [9,20,21]. Nanocrystalline specimens. For OM observations, the samples were etched
structures can be introduced into copper by cold rolling in a solution containing HNO3, H3PO4 and H4C2O5 (with
to very large extensions at cryogenic temperatures, and for- a ratio of 1:1:1) for 5 s. The detailed microstructures and
mation of nanostructures is attributed to dynamic recrys- quantitative microstructure statistics of the samples were
tallization [9]. Extremely fine grains (<100 nm) can be characterized by means of transmission electron micros-
achieved in copper by means of either BM [20] or by copy (TEM) and high-resolution transmission electron
SMAT [21]. Formation of deformation twins and shear microscopy (HRTEM), performed on a JEOL 2010 high-
bands is responsible for the structure refinement in the resolution transmission electron microscope at an operat-
BM process, and the high density of nanoscale twins is ing voltage of 200 kV. The cross-sectional thin foils for
believed to play a crucial role in the grain refinement in TEM observations were prepared by means of double-jet
the SMAT process. electrolytic polishing in an electrolyte consisting of 25%
The objective of the present work was to study the (volume) alcohol, 25% phosphorus acid and 50% deionized
microstructure evolution and nanostructure development water at about 10 C.
in pure Cu subjected to DPD at cryogenic temperatures,
which is of significance for both the fundamental under- 3. Results and discussion
standing and technological applications of the DPD
processes. Fig. 1b–d shows cross-sectional optical micrographs of
the DPD samples with different levels of strain. Appar-
2. Experimental ently, the equiaxed coarse grains in the original as-annealed
sample were pancaked into flat grains perpendicular to the
2.1. Sample loading direction. With an increasing strain, grains become
more flat and refined. Statistical measurements of the lon-
Copper cylinders (9 mm in diameter and 12 mm in thick- gitudinal (DL, perpendicular to loading direction) and
ness) with a purity of 99.995 wt.% were used as raw mate- transverse grain sizes (DT, parallel to loading direction)
rial for DPD treatment. Prior to the DPD treatment, the indicated that the average DL/DT ratio increases with
Cu cylinders were annealed in vacuum at 973 K for 2 h strain, from 1 (original) to about 3.5 for e = 1.1, as shown
to diminish the effect of mechanical processing and obtain in Fig. 1e. The plastic deformation, as evidenced by the dif-
homogeneous coarse grains. The grain sizes of the as- ference in contrast, seems to be inhomogeneous in different
annealed sample were within the range of 100–250 lm grains and even in different regions of the same grain.
and annealing twins were found in some grains (Fig. 1a). Detailed microscopic characterization of the deformed
structures is given below.
2.2. DPD treatment
3.1. DPD Cu sample with e = 0.4
The DPD treatment was performed on a dynamic com-
pression facility. A copper cylinder sample placed on a The optical micrograph image of the DPD sample with
lower anvil was compressed by an upper impact anvil at e = 0.4 is composed of two distinct types of regions with
232 Y.S. Li et al. / Acta Materialia 56 (2008) 230–241

a ε= 0 b ε = 0.4 12 e
G 8
ε = 0.4

Loading direction
4
W
0

Frequency (%)
12

200 μm 200 μm 8 ε = 0.7


4
c ε = 1.1 d ε = 2.1
0

Loading direction
Loading direction

12

8 ε = 1.1
D 4

0
200 μm 200 μm 1 2 3 4 5 6 7
D /D
L T

Fig. 1. (a–d) Cross-sectional metallographic observations of the DPD Cu samples with different strains: (a) e = 0 (as-annealed), (b) e = 0.4, (c) e = 1.1 and
(d) e = 2.1. The letters ‘‘W’’ ‘‘G’’ and ‘‘D’’ denote the white region (hard-to-etch), the gray region (easy-to-etch) and the dark-gray region (very-easy-to-
etch), respectively. (e) Distribution of aspect ratios of grain sizes (DL/DT) in the three samples with different strains.

different contrasts: ‘‘white’’ (i.e. hard-to-etch) regions sim- be well developed and their sizes are of several hundred
ilar to that in the original sample (e.g. a region labeled with nanometers. The cell walls are rather chaotic regions with
‘‘W’’ in Fig. 1b) and ‘‘gray’’ (easy-to-etch) regions (labeled a high density of dislocations, while the cell interior has rel-
as ‘‘G’’ in Fig. 1b). The total area fraction of the ‘‘white’’ atively fewer dislocations. The corresponding selected area
regions is apparently larger than that of the ‘‘gray’’ regions electron diffraction (SAED) pattern, similar to that of a
in this sample. The gray regions are not distributed homo- copper single crystal, indicates that the misorientation
geneously throughout the whole sample. across the DC walls is negligible. This morphology is anal-
Close observation showed that no obvious structural ogous to that observed in conventionally deformed Cu that
feature is seen in the W region (Fig. 2a). In the G region, resulted from cross-slips of dislocations [22].
some parallel bands are detected in a deformed grain In addition to DCs, numerous deformation twins were
(Fig. 2b). These bands seem to initiate from the grain found in the form of bundles embedded in some deformed
boundary (GB) on the left-hand side and terminate inside grains, as evidenced in Fig. 3b. An SAED pattern along a
the grain (as indicated by black arrows) or at the GB on [0 1 1] zone axis from the twin bundles (Fig. 3c) showed the
the other side of the grain (indicated by white arrows). superposition of a couple of Æ0 1 1æ diffraction patterns
Occasionally, terminations at annealing twinning bound- which are symmetrical to each other with respect to the
aries are also visible in some grains. These band structures {1 1 1} plane, indicative of a typical twin relationship
indicate inhomogeneous deformation inside the grains. ({1 1 1}/[1 1 2] type) among the lamellae. These twin/matrix
Cross-sectional TEM observations of this sample (T/M) lamellae form bundles with a total width of several
revealed that a high density of dislocations cells (DCs) hundreds of nanometers and a length of several microme-
exists in most grains (Fig. 3a). The DCs do not seem to ters to several tens of micrometers. The measured statistic

Fig. 2. High-magnification metallographic images of the W region (a) and the G region (b) as indicated in Fig. 1b (e = 0.4).
Y.S. Li et al. / Acta Materialia 56 (2008) 230–241 233

a b

400 nm 400 nm

20
d
c 15

Volume fraction (%)


222M,T
022T
200M 111M,T
022M
200T
10
000

0
0 40 80 120 160
T/M lamellar thickness (nm)

Fig. 3. (a, b) Typical cross-sectional bright field TEM images for the DPD Cu sample with e = 0.4. The corresponding SAED pattern for (a) is shown in
the inset. The SAED pattern and its index of the twin bundles in (b) are shown in (c). (d) The measured statistic distribution of the T/M lamellar thickness.

thickness distribution for the T/M lamellae from TEM stress, sufficient strain hardening (or dislocation density) is
observations (Fig. 3d) is characterized by a normal loga- needed [24,25]. The present TEM result (Fig. 3b) confirmed
rithmic distribution within a range from several to that appreciably plastic deformation by slip takes place
150 nm. The mean thickness is approximately 47 nm. It fol- prior to activating deformation twinning. It is also seen
lows from many of the TEM images that the volume frac- that severe deformation primarily occurs near GBs or
tion of twinning region is about 10% in the deformed annealing twinning boundaries (TBs), where deformation
sample (e = 0.4). twins are preferred to form when the critical stress for twin-
From the observed microstructure characteristics, we ning is exceeded. With an increasing strain, the critical
identified that the hard-to-etch (white) regions under opti- stress of twinning can be achieved in more and more
cal microscopy observations correspond to deformed strain-hardened grains. The population of deformation
grains with DCs and the easy-to-etch (gray) regions to twins thereby increases, as presented later.
the deformed grains with deformation twins. Actually the Formation of nanoscale deformation twins in the DPD
observed bands in Fig. 2b are from deformation twin Cu samples is distinct from that in the SPD Cu samples
bundles. deformed at low strain rates and ambient temperatures,
Dislocation slip and deformation twinning are two typi- in which few deformation twins are formed. Structure evo-
cal plastic deformation modes for fcc Cu. Dislocation activ- lution and refinement in SPD Cu samples are carried out
ities are thermally activated processes. Various dislocation primarily via dislocation activities with negligible twinning
configurations, including dislocation cells, walls, geometri- deformation. Obviously, the high strain rate and cryogenic
cally necessary boundaries (GNBs) and incidental disloca- deformation temperatures in the DPD process are key
tion boundaries (IDBs), are formed via manipulation and parameters for forming a high density of nanoscale defor-
rearrangement of dislocations when a polycrystalline Cu mation twins, both of which are essential for suppressing
is deformed at room temperature or above (e.g. in cold roll- dislocation activities. The large plastic strain is another fac-
ing [22] or ECAP [23]). However, when deformed at lower tor to enhance the twinning processes because of a substan-
temperatures, dislocation manipulation and rearrangement tial strain-hardening of grains. As will be demonstrated in
are suppressed, leading to formation of some poorly devel- the following sections, more deformation twins are formed
oped dislocation cells and a small number of microbands. in subsequent straining.
Another consequence of suppressed dislocation activi-
ties is that deformation twinning becomes more significant 3.2. DPD Cu sample with e = 1.1
in carrying plastic strain during DPD at cryogenic temper-
atures. The local stress required to nucleate deformation With increasing deformation strain, the gray regions
twinning in Cu is high. In order to reach such a high local expand at an expense of the white, indicating that more
234 Y.S. Li et al. / Acta Materialia 56 (2008) 230–241

T/M lamella bundles are formed. In the sample with


e = 1.1, some ‘‘dark gray’’ regions (labeled as ‘‘D’’ in a
Fig. 1c) appeared in addition to the white and gray regions,
as seen in Fig. 1b. The area fraction of white regions is con-

Loading direction
siderably decreased.
The presence of ‘‘dark gray’’ regions means that the
local structure is very easy to etch. A close look at the
‘‘dark gray’’ regions revealed that high density parallel
bands (analogous to that seen in gray regions, i.e. T/M
lamellae) exist, as seen in Fig. 4a. Within the parallel T/
M lamellae a few distinct bands are visible which intersect
the T/M lamellae with an angle of about 20–30. These
bands are separated by a spacing of about a few microme-
ters to a few tens of micrometers. Some bands ‘‘cut’’ 20 μm
through the whole twinned region, and some terminated
within the twin bundle. The morphology implies that shear
banding might have occurred within the twin bundles. This b
has been verified by cross-sectional TEM observations, as
shown in Fig. 4b, in which a shear band about 300 nm wide
is present in the nano-sized T/M bundle. Within the shear
band, nano-sized grains are detected. The T/M lamellae
become clearly curved in the vicinity of the shear band.
The thickness and orientation of the shear bands vary SB
considerably from grain to grain and even in a single grain.
The intersection angles between the shear bands and T/M
lamellae determined from metallographic observations
over a large volume of the sample, within a range of
15–35 (as shown in Fig. 4c), are consistent with those 400 nm
determined from TEM observations. TEM measurements
showed that the thickness of shear bands varies from 30
100 nm to several micrometers. The lengths of shear bands
are mostly dependent on the width of the twin bundles they
c
25
shear. Since the twin bundles are usually constricted within
one grain, the shear bands should be intragranular. 20
Frequency (%)

Detailed analysis of characteristics of shear bands are pre-


sented in Section 3.3.2.
15
Shear banding is a kind of instability which often
emerges when homogeneous plastic deformation cannot
10
sustain work hardening. This deformation mode frequently
appears in hard materials such as nanocrystalline materials
[26] and amorphous alloys [27]. In the nanoscale T/M 5

lamella bundles, where further work hardening becomes


very difficult, localized deformation in the form of shear 0
10 15 20 25 30 35 40
banding appears at stress concentrations within the bundle.
θ (degree)
Sharp boundaries are seen between the shear band and the
unsheared (T/M) region. In the vicinity of the shear band Fig. 4. (a) A magnified metallographic image of the D region outlined in
boundaries, T/M lamellae become curved to fit the shear- Fig. 1(c), showing several shear bands in the highly twinned region. (b) A
ing direction. typical TEM image showing a shear band that intersects with the T/M
The characteristics of shear bands are determined by lamellae. (c) Measured statistic distribution (from optical microscopy and
TEM observations) for the intersection angles between shear bands and T/
many factors, including the deformation temperature,
M lamellae.
strain rate, and the crystallographic structure etc. [28].
Due to the ample slip systems, dislocation slip is usually
the main deformation carrier in copper. Shear bands are [29]. In the present DPD process, dislocation activities are
found uncommonly ‘‘cutting’’ the microband matrix in severely suppressed, as discussed above. Moreover, the T/
heavily rolled copper. This type of shear band is usually M lamellae are so fine that it is difficult to activate disloca-
formed in metals with a high or medium SFE and is based tions or to refine the T/M lamellae by forming additional
on dislocation matrix; it is called a copper-type shear band microtwins. Consequently, further straining of the
Y.S. Li et al. / Acta Materialia 56 (2008) 230–241 235

nanoscale T/M lamellae is accommodated by shear band- that of low SFE fcc metals, following a sequence of dislo-
ing. This type of shear band is often formed in materials cation–twinning–shear banding [31].
with a medium to low SFE and is based on twinning
matrix; it is called a ‘‘brass-type’’ shear band [29]. 3.3. DPD Cu sample with e = 2.1
Formation of an S-band, a band of intense local crystal-
lographic glide, was found in rolled Ni and Al [30]. During For the sample with e = 2.1, the deformation structure is
traditional plastic deformation, S-bands usually appear rather blurred but more homogeneous when compared
earlier than shear bands and are more dense than them. with that of copper deformed to small or medium strain
During the DPD at cryogenic temperatures, dislocation (Fig. 1a–c), and microstructure identification becomes very
glide is suppressed, making it difficult for the traditional difficult under OM, as seen in Fig. 1d. To clarify the e = 2.1
S-band to form. The deformation was predominantly from sample’s microstructure, cross-sectional TEM observations
twinning and shear banding in DPD Cu at medium to high were performed.
strains. Fig. 5a and b are representative TEM micrographs from
It is generally believed that a decrease in deformation this sample. It is obvious that the microstructure is com-
temperature of fcc metals has a effect similar to decreasing posed of nanometer-thick T/M lamellae and nanometer-
SFE, which is verified by the present result. That is, during sized grains (crystallites). The nanoscale T/M lamellae,
deformation at high strain rates and at cryogenic tempera- which are perpendicular to the deformation direction, are
tures, the microstructure evolution in copper is similar to in the form of bundles in which several shear bands are

Fig. 5. (a, b) Typical cross-sectional bright field TEM images for the DPD Cu sample with e = 2.1. The microstructure consists of nano-sized T/M
lamellae and three types of nano-sized crystallites as indicated and described in the text.
236 Y.S. Li et al. / Acta Materialia 56 (2008) 230–241

formed (as indicated in Fig. 5a). The density of shear bands


in this sample is much higher than that in e = 1.1. Statistic a
TEM measurements showed that the volume fraction of T/
M lamellae is about 33%, with nano-sized grains constitut-
ing the remaining volume.
TEM characterization of the T/M lamellae (Fig. 6a and
b) indicated that there are high-density dislocations locat-
ing at the TBs, typical of deformation twins. This is dis-
tinct from the growth twin structure in the same material
in which no dislocation exists at TBs [32]. The TBs are
highly stressed due to presence of numerous dislocations,
as evidenced by the HRTEM images (Fig. 6b). SAED pat-
terns indicate the same twin relationship as obtained in the
sample with e = 0.4. The T/M lamellar thickness is identi- 50 nm
cal to that in the sample with e = 0.4, as shown by the sta-
tistical measurements from TEM observations (Fig. 6c),
with an average value of about 49 nm. This means that, b
with an increase in the plastic strain, no thickening process
of the deformation twins occurs in the subsequent DPD
process. In other words, the T/M lamellae thickness is
insensitive to the degree of strain under the present DPD
conditions.
Nano-sized grains are found in three different types of
regions, indicating their different formation paths. The for-
mation mechanisms for these three types of nano-sized
grains are analyzed in following sections.

3.3.1. Nano-sized grains type (I): fragmentation of T/M


lamellae 4 nm
Within the T/M lamella bundles, discontinuous and
elongated ‘‘bamboo-like’’ nano-sized crystallites are identi- 20
fied, distinct from the continuous T/M lamellae structure, c
as shown in Figs. 5a and 7. From the dark field TEM image
and electron diffraction one can see that the nano-sized 15
grains possess random orientations (Fig. 7b and c). The
Volume fraction (%)

statistical measurement of the grain sizes of the nanograins,


as shown in Fig. 7e, indicated that the average transverse 10
grain size is about 47 nm and the longitude size is a few
hundred nanometers, with an average value of 119 nm
(the average aspect ratio is about 2.5). It is worth noting
5
that the average transverse grain size is coincident with
the T/M lamellar thickness (49 nm).
Micro-area electron diffraction within a very small vol-
ume (small circle in Fig. 7a) indicated that between the 0
0 40 80 120 160
adjacent grains, the fcc twin orientation relationships still T/M lamellar thickness (nm)
exist. These results indicate clearly that the discontinuous
grains are evolved from continuous T/M lamellae which Fig. 6. (a) A close TEM observation of the T/M lamellae in the DPD Cu
with e = 2.1. (b) An HRTEM image from the region outlined in (a).
were subdivided during subsequent deformation. Plenty (c) Statistic distribution of the T/M lamellar thickness in the sample
of TEM observations showed that the total volume frac- determined from TEM observations.
tion of this type of nano-sized grains is roughly about
30% in the whole sample.
Formation of nano-sized grains in nanothick T/M also changed due to accumulation of increasing number of
lamellae, i.e. fragmentation of T/M lamellae, resulted from dislocations at boundaries. These blocks will gradually
dislocation–TB interactions. When plastic strain is applied evolve into randomly orientated grains. Such a grain refine-
to the T/M lamellae, dislocation walls/tangles are devel- ment process involving the interactions of twins and dislo-
oped within the lamellae, which subdivide the 2D T/M lay- cations has been evidenced in an Ni-based alloy [14] and
ers into discontinuous blocks, the orientations of which are pure copper [21] subjected to the SMAT processing.
Y.S. Li et al. / Acta Materialia 56 (2008) 230–241 237

a b

400 nm

32
25
c e

Frequency (%)
24
20
16

15 8

0
10 0 4 8
Volume fraction (%) dL/dT

5 Transverse (dT)

0
25

d 20

15
Longitudinal (dL)
10

0
0 100 200 300 400 500
Grain size (nm)

Fig. 7. TEM observations of nano-sized crystallites derived from T/M lamellae in the DPD Cu sample with e = 2.1. (a) A bright-field image; (b) a dark-
field image; (c) an SAED pattern corresponding to the large circle in (a); (d) an SAED pattern corresponding to the small circle in (a). (e) Statistical grain
size (both transverse and longitudinal) distribution for the type-I nano-sized grains. The distribution of aspect ratio for the nanograins is shown in the inset
of (e).

Nano-sized grains can be obtained in the top surface of Systematic TEM observations showed that sizes and ori-
SMAT Cu via fragmentations of deformation twins, which entations of the nanograins within shear bands vary with
shows that dislocation activities can operate within fine the band width. Fig. 9a1–3 shows TEM micrographs of
twins with thickness as small as 10 nm [21]. three typical shear bands with widths (w) of 0.5, 1.2 and
The aspect ratio of most grains is larger than 2.0, mean- 3.8 lm, respectively. The microstructures inside each shear
ing that most grains are not equiaxed. If further straining is band are characterized by nano-sized grains. From the cor-
applied to this sample, formation of more nanograins with responding SAED patterns inside the three shear bands, as
a smaller aspect ratio is anticipated. This has been verified shown in Fig. 9b1–3, one may see different scenes. For the
by our ongoing research [33]. thin shear band (a1) both random orientations and a typi-
cal fcc twin relationship can be detected. This means that
3.3.2. Nano-sized grains type (II): inside shear bands some of the grains still remain in the twinning connection
Nano-sized grains were found with each shear band, as while others are randomly orientated within the thin shear
shown in Figs. 5a and 8. Most grains are elongated along bands. For the thick bands, such as in a2 and a3, twinning
the shear direction and some are equiaxed. The transverse orientation is not obtainable and only random orientations
grain size ranges from 20 to 200 nm, with an average of are seen, i.e. the original twin relationship in T/M lamellae
75 nm. The longitudinal grain size is about twice that of has been destroyed in these thick shear bands.
the transverse size (Fig. 8). In most cases, random orienta- During shear banding, the original T/M lamellae might
tions have been detected among the nanograins. The mea- be swept into the localized shearing region (band) at high
sured volume fraction of the nano-sized grains within shear strain rates. The large degree of shear strain may induce
bands constitutes about 15%. fragmentation and rotation of the T/M lamellae, forming
238 Y.S. Li et al. / Acta Materialia 56 (2008) 230–241

14
b 25

Frequency (%)
12 20

a 10
15

10
8 5

6 0

Volume fraction (%)


0 2 4 6
dL/dT
4
2 Transverse (dT)
0
14
12
10 Longitudinal (dL)
8

200 nm 6
4
2
0
0 100 200 300 400 500
Grain size (nm)

Fig. 8. (a) A typical bright field TEM image showing nano-sized crystallites in a shear band in the DPD Cu sample with e = 2.1. The inset shows the
corresponding SAED pattern. (b) measured statistic distribution of grain sizes (transverse and longitudinal) for the nano-sized grains inside shear bands.
The distribution of aspect ratio for the nanograins is shown in the inset of (b).

elongated (or equiaxed) crystallites. This may explain our accompanied by softening. Thus, in subsequent straining,
observation that twinning connections remained partially deformation in existing shear bands may occur preferably
between some adjacent crystallites. For the crystallites that in the form of further shearing. This may lead to band thick-
have undergone large degrees of rotation, the twin relations ening and further grain coarsening following more mechan-
are destroyed. However, one may not exclude the possibil- ical and thermal activations. Eventually, the grain sizes tend
ity of secondary twinning within shear bands of which the to a saturated value, analogous to that in the SPD samples.
strain rate is very high. This possibility has been noticed in With an increasing number of shear bands, the T/M
other alloy systems before [34]. lamella bundles formed earlier are progressively eroded.
Another observation is that grains in thin shear bands Therefore, the volume fraction of T/M lamellae does not
are obviously smaller than those in thick bands. As shown increase monotonically with increasing strain. Instead, it
in Fig. 9c1–3 and d1–3, the average transverse grain size tends to saturate with further increasing strain (see Fig. 11a).
increases from 57 nm (w < 0.5 lm) to about 74 nm Shear bands intersecting pre-existing growth-in twins
(0.5 < w < 1.5 lm) and 79 nm (w > 1.5 lm). A similar trend have been reported in cold-rolling electrodeposited copper
is found for the longitudinal grain size. This observation [36]. The deformed structure inside the shear bands is
means that grain coarsening occurs simultaneously with dominated by a lamellar dislocation structure typical for
the thickening of shear bands. cold-rolled samples. In shock-loaded Cu samples, ultrafine
Usually, severe plastic deformation at high strain rates grain formation in adiabatic shear bands has also been
takes place within the adiabatic shear bands (narrow observed [37]. Ultrafine grains are thought to be formed
regions of highly localized deformation which experience via a dynamic recovery process, which helps formation
unusually high levels of strain and strain rate) during their of elongated subgrains with low angle boundaries at early
formation and thickening processes in which a substantial stages of shear banding, followed by breakdown and split-
transient temperature rise may also be induced. The R c temper- ting of the elongated crystallites into equiaxed crystallites
b
ature increase within a shear band is DT ¼ qc c0
s dc, in in the latter stage. Different from that in the shocked sam-
which b = 0.9 (assuming 90% of work of deformation was ples, shear bands in the DPD Cu sample are formed in
converted to heat), q is the density of the sample, c is the spe- nanoscale T/M lamellae. Hence, formation of nanograins
cific heat capacity and c0 is the shear strain for shear band in shear bands in the present DPD Cu may result primar-
initiation. From the equation, the temperature increment ily from the breakdown of nanoscale T/M lamellae and
DT is related solely to that of the shear strain in a shear rotations of the fragmented crystallites. This seems to be
band. The estimated transient temperature within shear consistent with our observation that the grain sizes in
bands in Cu during plastic deformation at ambient temper- shear bands are very close to the T/M lamellar thickness,
ature is about 500 K [35]. Such a high thermal pulse might which is much smaller than that in SPD samples (often
induce grain coarsening or dynamic relaxation (or recrystal- 200 nm).
lization) in pure Cu [35]. Therefore, coarsening of nano-
grains might be activated mechanically by the large plastic 3.3.3. Nano-sized grains type (III): outside T/M lamellae
straining at high rates and/or thermally by the transient In addition to type-I and type-II nanograins, some
temperature rises. Grain coarsening in shear bands is grains with sizes ranging from several tens of nanometers
Y.S. Li et al. / Acta Materialia 56 (2008) 230–241 239

a1 a2
DT
DT

SB SB
DT DT

500 nm 500 nm

DT
a3

SB

DT

500 nm

b1 b2 b3

24 c1 16 d1
18 12
w < 0.5 μm w < 0.5 μm
12 8

6 4

0 0
Volume fraction (%)
Volume fraction (%)

24 c2 16
d2
18 0.5 μm < w < 1.5 μm 12 0.5 μm < w < 1.5 μm
12 8

6 4

0 0
24 c3 16 d3
18 w > 1.5 μm 12 w > 1.5 μm
12 8

6 4

0 0
0 40 80 120 160 200 0 100 200 300 400
Grain size (perpendicular to SB, nm) Grain size (parallel to SB, nm)

Fig. 9. (a1–3) Typical TEM images of three shear bands with different widths (w) in the DPD Cu sample with e = 2.1. (b1–3) The corresponding SAED
patterns for the three shear bands. (c1–3, d1–3) Statistical distributions for the transverse and longitudinal sizes of the nano-sized grains inside shear bands
with different widths (c1 and d1: w < 0.5 lm, c2 and d2: 0.5 lm < w < 1.5 lm, c3 and d3: w > 1.5 lm), respectively, determined from a large number of
TEM images.
240 Y.S. Li et al. / Acta Materialia 56 (2008) 230–241

a b
15

Volume fraction (%)


10

400 nm 0
0 50 100 150 200 250 300
Grain size (transverse, nm)

Fig. 10. (a) A typical TEM image showing the nano-sized grains outside T/M lamellae in the DPD Cu sample with e = 2.1. Inset is the corresponding
SAED pattern. (b) Measured statistic distribution of transverse sizes for the type-III nano-sized grains.

to submicrometers are observed outside the T/M lamella cron-sized grains via dislocation manipulation and rear-
bundles. Most of these grains are elongated with curved rangement implies that deformation twinning in these
boundaries and some are equiaxed, as shown in Fig. 10a. regions has not been initiated, probably due to unfavorable
Misorientations across the boundaries vary from a few orientations of the original coarse grains or insufficient
degrees to a few tens of degrees. High-density dislocations local stress under the DPD conditions.
are always observed in association with these grains. These Fig. 11b depicts the variation of several characteristic
grains are obviously larger than type-I and type-II nano- sizes in the DPD Cu samples with different strains. We
grains. The average transverse size of these grains is about found that the T/M lamellar spacing does not change with
121 nm, as depicted in Fig. 10b. The longitude sizes range the plastic strain under the present experimental condi-
from several submicrometers to micrometers. The total vol- tions. The average sizes of three types of nanograins do
ume fraction of type-III grains is about 22%. not seem to depend on the strain either.
From the morphology and boundary misorientations, Table 1 summarized the structure characteristics of the
one may find that the refined grains are evolved from DPD Cu sample with e = 2.1. Obviously this sample is
DCs, which is analogous to those in SPD Cu and in deep composed of a mixed nanostructure: about 33 vol.% nano-
layers of SMAT Cu samples [21] in which dislocation activ- scale twins (or T/M lamellae) with three types of nano-
ities dominate the deformation and grain refinement pro- sized grains. The average size of all grains is about
cess. The sizes of this type of grain are comparable to 68 nm, which is much smaller than that in SPD Cu samples
those obtained in SPD Cu [38]. Formation of these submi- [23,38].
Tensile test results showed that the DPD Cu sample
exhibits a high tensile yield strength of about 600 MPa
and an ultimate tensile strength of 633 MPa [11]. Both
45 Twin volume fraction (%) a are much higher than those of SPD Cu samples (415
Twin volume (%)

30
and 455 MPa, respectively [23]). In addition, the tensile
elongation-to-failure of the DPD Cu is about 11%,
15 which is comparable to that of the SPD Cu [23]. These
results indicate the DPD Cu sample may possess an
0 enhanced fracture toughness in comparison with the
T/M lamellar thickness
Transverse size of nano-grain I
b SPD Cu samples. Detailed investigations on the
120
Transverse size of nano-grain II mechanical behaviors of the DPD Cu samples are in
Transverse size of nano-grain III
progress [33,39].
Average size (nm)

90

60
Table 1
Volume fraction and average thickness (kT/M) for the T/M lamellae and
30 average sizes (dT: transverse size, dL: longitudinal size) for three types of
nanograins in the DPD Cu sample with e = 2.1, determined from cross-
0 sectional TEM observations
0.0 0.5 1.0 1.5 2.0
T/M lamellae Nano-sized grains
Strain
Type I Type II Type III
Fig. 11. Variation of volume fraction of T/M lamella bundles (a) and
Volume (%) 33 30 15 22
variation of several characteristic sizes (b) as a function of deformation
Average size (nm) kT/M = 49 dT = 47 dT = 75 dT = 121
strain, including average T/M lamellar thickness, transverse sizes for three
dL = 119 dL = 160
types of nanograins derived from different origins.
Y.S. Li et al. / Acta Materialia 56 (2008) 230–241 241

4. Conclusions [3] Iwahashi Y, Horita Z, Nemoto M, Langdon TG. Acta Mater


1998;46:3317.
[4] Hebesberger T, Stüwe HP, Vorhauer A, Wetscher F, Pippan R. Acta
Dynamic plastic deformation of bulk pure Cu samples Mater 2005;53:393.
at cryogenic temperatures induces a significant microstruc- [5] Sanders PG, Eastman JA, Weertman JR. Acta Mater 1997;45:4019.
tural evolution and formation of nanostructures. The dom- [6] Koch CC. Nanostruct Mater 1993;2:109.
inant mechanisms governing the plastic deformation and [7] Hughes DA, Hansen N. Phys Rev Lett 2001;87:135503.
microstructural refinement include dislocation manipula- [8] Tao NR, Wang ZB, Tong WP, Sui ML, Lu J, Lu K. Acta Mater
2002;50:4603.
tion and rearrangement, deformation twinning forming [9] Wang YM, Chen MW, Sheng HW, Ma E. J Mater Res 2002;17:
nanoscale T/M lamella bundles, and shear banding in the 3004.
T/M lamellae. At low strains, dislocation activities and [10] Zhu YT, Jiang HG, Huang JY, Lowe TC. Metall Mater Trans A
twinning dominate the deformation process. With increas- 2001;32:1559.
ing plastic strain an obviously increasing tendency of defor- [11] Zhao WS, Tao NR, Guo JY, Lu QH, Lu K. Scripta Mater
2005;53:745.
mation twinning and shear banding was observed. [12] Bay B, Hansen N, Hughes DA, Kuhlmann-Wilsdorf D. Acta Metall
For strain e = 2.1, a mixed nanostructure is formed in Mater 1992;40:205.
the DPD Cu sample with about 33 vol.% of nanoscale T/ [13] Zhang HW, Hei ZK, Liu G, Lu J, Lu K. Acta Mater 2003;51:1871.
M lamellae and 67 vol.% of nano-sized grains. The nano- [14] Tao NR, Wu XL, Sui ML, Lu J, Lu K. J Mater Res 2004;19:1623.
grains can be classified into three types in terms of their ori- [15] Tao NR, Zhang HW, Lu J, Lu K. Mater Trans 2003;44:1919.
[16] Blewitt TH, Coltman RR, Redman JK. J Appl Phys 1957;28:651.
gin: (i) nanograins derived from fragmentation of [17] Niewczas M, Basinski ZS, Basinski SJ, Embury JD. Phil Mag A
nanoscale T/M lamellae with an average transverse size 2001;81:1121.
of about 47 nm; (ii) nanograins in shear bands with an [18] Gray III GT, Follansbee PS, Frantz CE. Mater Sci Eng A 1989;111:
average transverse size of about 75 nm; and (iii) nanograins 9.
derived from dislocation cells with an average transverse [19] Murr LE, Esquivel EV. J Mater Sci 2004;39:1153.
[20] Huang JY, Wu YK, Ye HQ. Acta Mater 1996;44:1211.
size of about 121 nm. Obviously, deformation twinning [21] Wang K, Tao NR, Liu G, Lu J, Lu K. Acta Mater 2006;54:5281.
induced by high strain rates and cryogenic temperatures [22] Hansen N. Mater Sci Technol 1990;6:1039.
in the DPD process plays a critical role in the microstruc- [23] Torre FD, Lapovok R, Sandlin J, Thomson PF, Davies CHJ. Acta
tural evolution and formation of nanograins, which is the Mater 2004;52:4819.
fundamental difference from that in the conventional [24] Christian JW, Mahajan S. Deformation twinning. Prog Mater Sci
1995;39:1.
SPD processes. [25] Meyers MA, Vöhringer O, Lubarada VA. Acta Mater 2001;49:4025.
[26] Wei Q, Jia D, Ma E, Ramesh KT. Appl Phys Lett 2002;81:1240.
Acknowledgements [27] Johnson WL. MRS Bull 1999;24:42.
[28] Hatherly M, Malin AS. Scripta Metall 1984;18:449.
Financial supports from the National Natural Science [29] Hatherly M, Malin AS. Met Technol 1979;6:308.
[30] Hughes DA, Hansen N. Acta Mater 2000;48:2985.
Foundation of China (Grant Nos. 50671106, 50621091 [31] Blicharski M, Gorczyca S. Met Sci 1978;12:303.
and 50431010) and the Ministry of Science and Technology [32] Lu L, Shen YF, Chen XH, Qian LH, Lu K. Science 2004;304:422.
of China (Grant No. 2005CB623604) are acknowledged. [33] Zhang Y, Tao NR, Lu K, in press.
The authors thank Q.H. Lu, B. Wu, C.S. Hong and Y. [34] Xue Q, Liao XZ, Zhu YT, Gray III GT. Mater Sci Eng A 2005;410–
Zhang for stimulating discussions. 411:252.
[35] Mishra A, Kad BK, Gregori F, Meyers MA. Acta Mater 2007;55:13.
[36] Huang X, Lu QH, Sui ML, Li DX, Hansen N. Mater Sci Forum
References 2007;539–543:5013.
[37] Andrade U, Meyers MA, Vecchio KS, Chokshi AH. Acta Metall
[1] Gleiter H. Acta Mater 2000;48:1. Mater 1994;42:3183.
[2] Valiev RZ, Islamgaliev RK, Alexandrov IV. Prog Mater Sci [38] Belyakov A, Sakai T, Miura H, Tsuzaki K. Phil Mag A 2001;81:2629.
2000;45:103. [39] Li YS, Tao NR, Lu K, in press.

You might also like