Vous êtes sur la page 1sur 9

Deformation Microstructure in (001) Single Crystal Strontium Titanate

by Vickers Indentation
Kai-Hsun Yang, New-Jin Ho, and Hong-Yang Lu*
,w
Department of Materials Science and Centre for Nanoscience, National Sun Yat-Sen University,
Kaohsiung 80424, Taiwan
Recent interests on the plastic deformation of strontium titanate
(SrTiO
3
) are derived from its unusual ductile-to-brittle-to-duc-
tile transition (DBDT). The transition is divided into three
regimes (A, B, and C) corresponding to the temperature range
of 1131053 K (1601 to 7801C), 1053 to B1503 K (7801 to
B12301C), and B15031873 K (B12301 to 16001C), discov-
ered by Sigle and colleagues in the MPI-Stuttgart. We report
the dislocation substructures in (001) single crystal SrTiO
3
deformed by Vickers indentation at room temperature, studied
by scanning and transmission electron microscopy. Dislocation
dipoles of screw and edge character are observed and conrmed
by insideoutside contrast using 7g-vector by weak-beam dark
eld imaging. They are formed by edge trapping, jog dragging,
and cross slip pinching-off. Similar to dipole breaking off in de-
formed sapphire (a-Al
2
O
3
) at 12001C and g-TiAl intermetallic
at room temperature, the dipoles pinch off at one end, and emit a
string of loops at trail. Two sets of slip systems {110}/1

10S
and {100}/011S are activated under both 100 g and 1 kg load.
The suggestion is that plastic deformation has reached the stage
II work hardening, which is characterized by multiplication of
dislocations through cross slip, interactions between disloca-
tions, and operating of multiple slip systems.
I. Introduction
M
OST ceramics are brittle and fractured by cleavage without
appreciable plastic deformation (of a failure strain
eo0.1%) at room temperature, but with increasing tempera-
ture they become more plastic at and above a transition tem-
perature (T
C
). The transition temperature varies, e.g., 41273 K
(10001C) for sapphire (single-crystal alumina (a-Al
2
O
3
)).
1
Such
mechanical behavior is commonly known as the brittle-to-
ductile transition (BDT).
14
Strontium titanate (SrTiO
3
) has the perovskite structure, and
retains the cubic symmetry until below 105.5 K (167.51C).
5
Not
only does it undergo a signicant plastic deformation by a com-
pressive strain e 57%9% at room temperature, but also exhib-
its an unusual ductile-to-brittle-to-ductile transition (DBDT).
68
Taking SrTiO
3
as a structural analogue,
9
the rheology of Earths
lower mantle,
10
consisting predominantly of silicate perovskites,
(Mg,Fe)SiO
3
, can be simulated and investigated.
Two types of BDT are commonly reported.
11,12
Gradual
transition from brittle to ductile is more frequently observed,
where the fracture stress increases over a temperature range of
4373 K (1001C), e.g., in steels, BCC metals, intermetallics, and
Al
2
O
3
. On the other hand, transitions in Si and sapphire are very
sharp and occur abruptly within 51101C. The transition is a
function of the strain rate (or the rate of the applied stress in-
tensity factor) and structure sensitive dependent on the variabil-
ity in the density of dislocation sources, and is temperature
dependent.
11,13
Sharp transition is due to triggering of disloca-
tion sources while gradual transition has dislocation sources
readily available at the crack tip. Not only does SrTiO
3
exhibit
unusual ductility at room temperature,
68
its BDT to high-
temperature ductile behavior over an extended temperature
range of B1050 K (7771C) and B1500 K (12271C) also sug-
gests the transition is the gradual type.
3,4
Deformed microstructures in SrTiO
3
have been reported.
68,14
They were often taken to support the proposed DBDT models,
e.g. dissociated collinear partials of smaller cores and higher
core energy rendering larger Peierls stresses were taken to suggest
temperature-dependent core structure, and decreasing length of
mobile dislocation segments. However, the room temperature
plasticity for a traditionally brittle ceramic is yet satisfactorily
explained.
The Vickers indentation technique is used to investigate
how dislocations are generated and migrated in SrTiO
3
in the
plastic regime (i.e., regime A in Brunner et al.
14
). We report
an analysis of dislocation substructure thus produced in (001)
single crystal to understand how the plasticity occurs at room
temperature.
II. Experimental Procedures
Single crystal SrTiO
3
containing 0.7 wt% Nb-doping (approx-
imately 0.02 mol/% Nb
2
O
5
), of the /001S orientation supplied
by MTI (Richmond, CA) were used in this study. The crystals
were microindented by two applied loads of 100 g and 1 kg using
a Vickers indenter (Nikon AVK-C2, Tokyo, Japan). Extrinsic
defects, e.g. charge-compensating cation vacancies of V
Sr
00
and
V
Ti
00 00
(using the Kro ger-Vink notation) for Nb
2
O
5
-donor oxide,
may have contributed to plastic deformation by dislocation pin-
ning, and kink migration. The indented samples were then
chemical-etched for 60 s in a solution of 70% HNO
3
added
with a few drops of HF. Microstructure was observed using the
scanning electron microscope (SEM, SEM6330, JEOLt, To-
kyo, Japan) equipped with a cathodoluminescence (CL) detector
(MonoCL2, Gatant, Pleasanton, CA) operating at 20 kV, and
the transmission electron microscope (TEM, AEM3010,
JEOLt) operating at 300 kV.
TEM foils were prepared either by the conventional tech-
nique of cutting, polishing, dimple-grinding mechanically, and
Ar-ion beam thinning (PIPSt, Gatan) to electron transparency,
or using a focus ion beam (FIB) apparatus (SMI3050, Seiko,
Tokyo, Japan).
III. Results
A representative indentation crack is shown in Fig. 1 (for load-
ing with 1 kg). A large c/d-ratio 42.5 (where d is the diagonal of
T. Mitchellcontributing editor
*Member, The American Ceramic Society.
This research is funded by the National Science Council of Taiwan through contract
NSC-95-2221-E110-033, and the Ministry of Education through the Centre of NanoScience
and Technology.
w
Author to whom correspondence should be addressed. e-mail: hyl@mail.nsysu.edu.tw
Manuscript No. 25998. Received March 11, 2009; approved May 2, 2009.
J
ournal
J. Am. Ceram. Soc., 92 [10] 23452353 (2009)
DOI: 10.1111/j.1551-2916.2009.03189.x
r 2009 The American Ceramic Society
2345
the Vickers impression and c the crack length measured from the
center of the impression) suggests
15
the crack is the half-penny
type. Fracture toughness K
IC
0.89 MPa m
1/2
, similar to those
reported by Bernard et al.,
15
was determined using the formula
K
IC
constantE=H
1=2
P=c
3=2
where constant is 0.0154 (dimensionless), E/H is the ratio of
elastic modulus (225 GPa) to hardness (9.5 GPa), and P the
applied load (9.8 N).
(1) Slip Lines and Dislocation Etch Pits
Figure 2(a) shows radial cracks emanating from the indent im-
pression (as indicated). Dislocation etch pits are aligned along
both the /010S and /110S directions, making a 451 angle to
each other.
15
The indication is that two slip planes intersect at
451 as depicted in the inset of Fig. 2(a). A schematic illustration
of etch pits constituting slip lines is shown in juxtaposition,
where the radial cracks lie parallel to /110S. Slip lines indicate
that the dislocations are generated predominantly around the
indent. Dislocations are viewed end-on, where they lie in slip
plane and constitute a slip band.
Dislocation multiplication by multiple cross glide is evidenced
from the extending and lateral growth of slip bands (Fig. 2(a))
along a major radial crack. A dislocation-free zone
16,17
of sev-
eral micrometers wide on both sides, straddled across the crack,
where no etch pits are visible (indicated in Fig. 2(a)), is detected
along the whole length of radial crack.
Strain eld and slip lines are clearly visible in CL images. The
areas of bright contrast indicate highly strained region produced
by indentation (Fig. 3(a)), and the criss-cross dark lines extend-
ing along the /110S and /100S directions are the slip lines.
These slip lines running along two directions intersect with
each other. While the strain contrast is revealed unambiguously
by CL, the slip lines are also signicantly better discerned
(Fig. 3(b)) as compared with SEM secondary-electron image
(SEI) in Fig. 2(a). This also conrms the observation shown in
Figs. 2(a) and (b), where dislocation-etch pits are analyzed by
SEM-SEI.
(2) Burgers Vectors and True Line Directions
Representative dislocations in the crack vicinity are shown
in Fig. 4(a) by (g,3g) weak-beam dark eld (WBDF) imag-
ing. All TEM images shown here are taken from along the [001]
zone axis.
Slip planes are determined by b u 5slip plane normal
from the dislocation Burgers vector (b) and the true disloca-
tion line direction (u). Both b and u are determined by conven-
tional contrast analysis; the former by adopting the invisibility
criteria of g b 50 or 2np from more than three noncoplanar
g-vectors, the latter by performing trace analysis for at least
three projected dislocation images from the corresponding
stereographic projection (using CaRIne
s
Crystallography,
version 3.1).
Tables I and II summarize an analysis of dislocations shown
in Figs. 4(a) and (b) for Burgers vectors and true line directions,
and the type of dislocations determined from angles between b
and u. Both the edge dislocation B
1
(with b
B
5[1

10]) in Fig. 4(a)


and the edge dislocation E
1
(with b
A
5[0

11]) in Fig. 4(b),


are produced by indentation. Accordingly, both the primary
{110}/1

10S
6
and secondary {001}/110S slip systems are acti-
vated by indentation with a 100 g load at room temperature.
The angle between the faces of the pyramidal Vickers indenter
with a square base is 1361. Although the Schmid factor vanishes
in pure compression along /001S, shear stresses generated
along the interface between the indenter and work piece
(i.e., /001S SrTiO
3
single crystal) would have made the Sch-
mid factor nonzero (m50.172) such that the slip system {001}
/110S is activated.
Edge dislocation glides in its slip plane, as indicated, is
evidenced from the upper left corner of Fig. 4(b). The intersec-
tion of screw and edge dislocations would have resulted in
forming jogs in both screw dislocation D
1
and edge dislocation
E
1
18
after they have moved passing each other. Figure 4(b)
reveals the mutual interaction occurring at the intersection, as
indicated in the inset. Sessile jogs can only move nonconser-
vatively by climb.
(3) Dislocation Intersection and Screw Dipoles
Dislocations in the framed region of Fig. 5(a) are further
analyzed. Intersecting screw (A
2
) and edge (B
2
) dislocations,
similar to those in Fig. 4(b) are again detected. Adopting the
insideoutside contrast technique, the features (as indicated by
arrows) are determined to be dislocation dipoles,
6
rather than
dissociated partials,
19
by imaging with 7g-vectors under
strong-beam diffraction conditions. Results of (g,3g) WBDF
images are shown in Figs. 5(b) and (c) under g 5

10(exhib-
iting outside contrast and wider separation between dipoles) and
1g 5110 (exhibiting inside contrast and narrower separation
between dipoles), respectively. Although collinear partials
in sintered SrTiO
3
ceramics
20
and in undoped bicrystals
21
were
reported, only dipoles
6
are detected here in the indented /001S
single crystal.
Both edge (D
1
) and screw (D
2
and D
3
) dipoles, and mixed
type (D
4
and D
5
), as indicated in Fig. 5(b), are found.
Screw dipole D
2
pinching-off at its end has cross slipped
and containing segments of a mixed nature. Most of the dipoles
are close ended in the view, one (D
1
) of the open-ended dipoles
is indicated in Fig. 5(c). Spherical loops pinched off from
the close end of a screw dipole are detected (indicated in
Fig. 5(a)).
Under the applied stress, an edge dislocation gliding in its
slip plane has drawn out two screw dislocation lines from
an edge jog, J
1
, as indicated in Fig. 5(d), and forms a screw
dipole (D
6
, also indicated). The two dislocation arms (desig-
nated D
6s
) joining at edge jog (J
1
) and forming a dipole, have
the same Burgers vector but mutually antiparallel line direc-
tions. An edge dipole bowing out at another sessile jog (J
2
)
generating an edge dipole trail is also observed, which is rela-
tively short in length (e.g., comparing with edge dipole D
1
and
screw dipole D
3
in Fig. 5(b)). Edge dipole trails are ubiquitous in
other areas of the deformed crystal, as indicated in Fig. 5(e).
Dislocation entanglement forming subgrain boundaries is
also evidenced.
(4) Dislocation Intersection and Edge Dipoles
Dislocations with Burgers vector b
B
5/

110S of a mixed char-


acter, indicated by lled arrowheads (Fig. 6(a)), are found to
cross slip in (110). Only the primary slip system {110}/

110S is
activated, representing the initial stage of plastic deformation.
Fig. 1. A typical indentation crack of the half-penny shape having c/d-
ratio 42.5 (SEM-SEI).
2346 Journal of the American Ceramic SocietyYang et al. Vol. 92, No. 10
The framed area imaged by WBDF is shown in Figs. 6(b) and
(c), using 7g-vectors. Two edge dipoles (indicated by D
7
and
D
8
) appear in the eld of view. At one end of dipole D
7
is its
bowed-out segments, of a screw character (D
7s
, as indicated in
Fig. 6(a)). A portion of dipole D
7
shows a distinctive contrast
from the rest of the dipole (Fig. 6(b)). The indication is that ei-
ther or both dipole arms, of an edge character, have slipped in
their own glide plane. At the other end of D
7
is a jog (designated
as J
3
). Nevertheless, the screw segment D
8s
(as indicated in
Fig. 6(c)) of edge dipole D
8
, indicated by lled arrowheads,
remain separated under the applied stress. Again, subgrain
boundaries are formed (Fig. 6(d)).
Dislocation dipoles are further decomposed into a string
of loops by pinching-off at dipole trail. Although appeared
in both samples, this is more pronounced in crystals indented
by the higher load of 1 kg. Loops are broken up from one end
of both the open-ended (Fig. 7(a)) and close-ended (Fig. 7(b))
dipoles.
Fig. 2. Dislocation etch pits (b) in the vicinity of radial cracks (a), running at 451 to each other (SEM-SEI).
October 2009 Deformation of SrTiO
3
by Vickers Indentation 2347
IV. Discussion
Dislocation substructure in Vickers-indented (001) single crystal
SrTiO
3
is represented by: (1) dislocations of edge, screw, and
mixed character, (2) sessile jogs, (3) dipole trails, edge, screw,
and mixed dipoles, (4) decomposition of such dipoles into loops,
and (5) formation of subgrain boundaries. Together with two
slip systems {110}/1

10S and {100}/011S activated by the in-


dentation stress, this suggests that plastic deformation in such
single crystals has the characteristic features of the stage II work
hardening (Figs. 4(a), 5(a) and 6(b), and Table II). Deformation
occurs when the easy glide system /1

10S{110} is operative,
as experimentally evidenced (Table II). It was activated at
B900 mN for the (001) single crystals,
22
but the stage I work
hardening would have been obscured at the indentation stresses
because the primary and secondary slip systems are both acti-
vated. Therefore, a signicant plastic deformation would have
taken place at room temperature before unstable crack growth
ensues to result in brittle fracture eventually.
(1) Mechanisms of Dipole Formation
For ductile materials, represented by FCC metals, dislocation
dipoles are a feature of the stage I hardening when dislocation
glide only occurs in a single slip system. They are formed by
edge trapping in this stage. Edge dislocations with Burgers
vectors of the opposite sign,
18,23,24
gliding passed each other
on parallel slip planes are trapped by their mutual (attractive)
stress eld and form edge dipoles.
23,24
Edge dipoles are also
produced by jog dragging when a gliding screw dislocation,
pinned by a sessile edge jog of small length (y),
18
draws out
two edge-dislocation lines of the same Burgers vector when
applied stresses (per unit length, tb) are unable to overcome
their mutual repulsion force (B0.25 Gb
2
/2p(1n)y, where
G: shear modulus, n: Poisson ratio). Jog dragging only occurs
after sessile jogs have been produced from the intersection
between dislocations.
Edge dipoles may have been generated in deformed SrTiO
3
single crystals by three mechanisms. (i) Firstly, similar to
sapphire,
23,24
edge dipoles in deformed SrTiO
3
(represented by
D
1
in Fig. 5(b)) are most likely produced from mutual trapping
initially in stage I hardening. Therefore, such edge dipoles of
opposite sign are open ended (e.g., D
1
).
Fig. 3. (a) Strain eld and (b) slip lines produced by indentation as re-
vealed by SEM-CL image.
Fig. 4. Dislocations in the crack vicinity shown by (g,3g) weak-beam
dark eld (WBDF) imaging whose Burgers vectors (b) and true line di-
rections (u) are determined [(a) in Table I and (b) in Table II], jogs (as
framed) are formed at the intersection between screw (D
1
) and edge (E
1
)
dislocations (TEM).
Table I. Determination of Burgers Vectors (b) Adopting
g b 50 Invisibility Criteria from WBDF Images at Three
Noncoplanar g-Vectors
Z
g b
b
001 1

13 102 1

12112
g 200

110 020 110 110 211 121 211 020 111 110
A
1
X O O O O X O O O X O 7[0

11]
B
1
X O O O O X O O O X O 7[0

11]
C
1
O O X O O O X O X X O 7[101]
Dislocations are invisible when g b 50. X, invisible; O, visible.
2348 Journal of the American Ceramic SocietyYang et al. Vol. 92, No. 10
(ii) Secondly, edge dipoles are formed by a mechanism
describable by cross slip and pinching-off proposed by Tetel-
man
25
for deformed single crystal Fe-3 wt%Si. Such mechanism
facilitated by a long jog produced by cross slip is best illus-
trated schematically in Figs. 8(a)(c). Double cross slip provides
sources for long jogs, which are the edge component laid
in the cross-slip plane. Two dislocations with Burgers vector b
of the same sign but inclined to each other are gliding on
two parallel planes (Fig. 8(a)). Part of their lengths reori-
ented by an applied stress, have resulted in two parallel edge
segments of opposite line directions (Fig. 8(b)). For small jog
length (y), only large stresses are able to separate the two
dislocations that they stay attracted mutually. If one end of
one of the dislocation segments can cross slip to join the other,
this creates a superjog (of a minimum length of the separation
between two parallel planes), and the joint pinches off readily
results in an edge dipole (Fig. 8(c)). Such dipoles generated
at jogs from cross slip are usually close ended and longer
in length, e.g., D
7
from long jog J
3
in Fig. 6(c). It is likely that
dipole D
8
is in an intermediate stage of cross slip and pinching-
off, where the two dislocations are cross gliding at one end
(Fig. 8(b)).
(iii) Thirdly, although edge dipoles can also be produced
from bowing out at an edge jog of a screw dislocation, they
are usually rather short in length and form dipole trails.
26,27
Sessile jogs are produced
18
when screw and edge dislocations,
or edge and edge dislocations intersect each other. When
two slip systems (Table II) are activated at room temperature,
edge dipoles (Figs. 5(a) and (b)) are formed by jog drag-
ging (e.g., at J
2
in Fig. 5(b)). Such sessile jogs are formed
in edge as well as screw dislocations with mutually perpendic-
ular line directions (i.e., screw dislocations D
1
and edge dislo-
cations E
1
shown in Fig. 4(b)). Nevertheless, two screw
dislocation arms tagging behind an edge dipole trail are easily
annihilated, as evidenced in single crystal MgO during in situ
tensile test.
27
Screw dislocations usually assumed to have a higher veloc-
ity
23
are easily annihilated by cross slip. Screw dipoles are there-
fore unlikely to be formed by trapping (i.e., mechanism (i)), this
leaves two possible mechanisms. They are either produced by
cross slip and pinching-off (mechanism (ii)
25
), or a direct result
of edge jogs on screw dislocations (mechanism (iii)). Such
dipoles are bowed out at a sessile jog from edge dislocations
(e.g., D
6
in Fig. 5(d)), which have been intersected by screw dis-
locations. Therefore, screw dipoles are usually the close-ended
type terminating at a sessile jog. On the other hand, edge dipoles
are both open ended (D
1
in Fig. 5(c)) when formed by edge
trapping in stage I and close ended at super jog J
3
(D
7
in
Fig. 6(b)) when formed by Tetelmans mechanism
25
(Figs.
8(a)(c)) in stage II hardening.
Table II. Slips Systems with Slip Plane Normal
Determined from Dislocation Burgers Vectors (b) and
True Line Directions (u)
Dislocation True direction u
t
Burger vector b Slip plane Dislocation type
A
1
u
A1
5[001] 7[ 0

11]

100 Mixed (451)


B
1
u
B1
5[1

44] 7[ 0

11] (011) Mixed (101)


C
1
u
C1
5[001] 7[101] 0

10 Mixed (451)
Fig. 5. (a) Dislocations in the framed region, forming jogs; weak-beam dark eld (WBDF) images using (b) g 5

10, (c) 1g 5110 shows inside and


outside contrast conrming that they are dislocation dipoles, other features include (d) screw dipole D
6
formed by a gliding edge dislocation in its slip
plane, and (e) edge dipole trails (TEM).
October 2009 Deformation of SrTiO
3
by Vickers Indentation 2349
Fig. 6. (a) Cross-slipped dislocations, (b) and (c) insideoutside contrast shown by weak-beam dark eld (WBDF) images of edge dipoles using
7g-vectors, respectively, which are produced by cross slip and pinching-off. (d) Sub-grain boundaries are formed (BF) image (TEM).
2350 Journal of the American Ceramic SocietyYang et al. Vol. 92, No. 10
(2) Formation of Dislocation Loops by Cross Slip
Similar loops were reported for sapphire
23,24
deformed by com-
pression at high temperatures (@12001C). It was modeled on the
basis of the Rayleigh instability,
28,29
describing the morpholog-
ical change of solid rods or pore channels. However, conserva-
tive climb in sapphire involving pipe diffusion
23,24
is unlikely to
occur at room temperature. This may only occur if atom diffu-
sion leading to unstable uctuation is signicantly enhanced in
SrTiO
3
by the applied stress.
Veyssie` re and Gre` gori
30
also proposed a mechanism of dipole
truncation by cross-slip annihilation for the prismatic loops
produced from unjogged dipoles in Al-rich g-TiAl deformed at
room temperature. A schematic illustration based on the orig-
inal drawings
30
is shown in Figs. 9(a)(d).
At a local position in open-ended dipoles (e.g., D
2
and D
3
in Fig. 5(b)), reorientation occurs to a screw character due to
the applied stress, which could then cross glide. The dipoles
are divided into two close-ended segments each terminated
at a sessile jog in the slip plane. When such cross glide
happens in several places along the dipole, an equal number
of loops are produced. It may occur to dipoles of an edge and a
mixed character. Edge dipoles may also acquire kinks to facil-
itate cross slip. Kinks are formed in edge dipoles when inter-
secting with edge dislocations, or nucleated upon moving across
Peierls barrier.
18
The process relies on cross slip of the screw portion when the
dipoles are anchored at both ends (by e.g., extrinsic defects or
impurities). Further, several cross-slip positions (i.e., the number
of loops formed as shown in Figs. 7(a) and (b)) must nucleate
simultaneously along the dipole. This necessitates that the uc-
tuation of dipole orientation make the dipole possess corre-
spondingly an equal number of positions with a screw character
for cross slip to ensue. Again, the driving force for dipole re-
orientation must come from the applied stress when deformed at
room temperature.
Price
31
suggested a mechanism for the prismatic loops
emitted from screw dipoles based on the observation of Zn
deformed at room temperature. Several segments along a screw
dipole must cross glide simultaneously under the applied stress.
The cross-slipped screw segments are annihilated and that
results in pinching-off to a string of prismatic loops. This pro-
cess is shown schematically in Figs. 10(a)(c), which eventually
leaves an edge dislocation with two large jogs (Fig. 10(d)) but it
is not observed here.
All mechanisms proposed before for sapphire, g-TiAl inter-
metallic, and Zn for pinching-off loops require stress uctuation
in a long-range stress eld produced by indentation along dipole
lines a priori for producing such loops, which is likely to come
from the applied stress.
(3) Plasticity at Room Temperature
The slip families and resulting numbers of independent slip sys-
tems from FCC, BCC, and HCP metals
32
and SrTiO
3
are listed
in Table III. It is derived experimentally from dislocation anal-
ysis that all together ve systems are activated in SrTiO
3
:
(010)[

101], (100)[0

11], (001)[110], (

101)[101], and (1

10)[110].
The von Mises criterion requires ve independent slip
systems
33
operative for signicant plasticity, but this is a neces-
sary not sufcient condition.
32
The ability of each grain in poly-
crystals deformed simultaneously by an arbitrary amount
of strain along the ve systems and the slip systems operative
cooperatively known as slip exibility,
32
is further required for
the ductility in polycrystalline solids to be realized. Although
only two or three of such slip systems were usually observed,
34,35
signicant plasticity in polycrystalline SrTiO
3
ceramics can be
expected.
The present results are consistent with Brunner and col-
leagues
6
in the sense that plastic deformation at room temper-
ature has also occurred in the indented samples. Although the
plastic strain is not determined experimentally, and dislocation
dissociation affecting its mobility is not detected here, the plas-
ticity is demonstrated unambiguously from the microstructure
features characteristic of the stage II work hardening (Figs. 4(a),
5(a) and 6(b), and Table II). Only that shear stresses introduced
by Vickers indentation has also induced the secondary slip
family {001}/110S. Its contribution to the deformation strain is
not known at the present time because the deformation strain
was not determined. Nevertheless, plasticity would be further
Fig. 7. Dislocation loops decomposed from (a) open-ended and
(b) closed-ended dipoles (TEM).
Fig. 8. Schematic illustration for a mechanism of edge dipole formation proposed by Tetelman
25
: (a) two dislocations of same b gliding in parallel
planes, (b) reorientation of a portion by applied stress leading to two parallel edge segments of opposite u, and (c) Cross slip resulting in joining of two
dislocations, pinching-off jog forms an edge dipole.
October 2009 Deformation of SrTiO
3
by Vickers Indentation 2351
enhanced with the additional, secondary slip systems induced by
the applied load.
V. Conclusions
Besides {110}/1

10S reported for compressed samples,


6
an
additional, secondary slip system of {001}/110S is identied
in (001) SrTiO
3
single crystal deformed at room temperature by
Vickers indentation. Both screw and edge dipoles are observed,
Fig. 10. Prices model
31
of loop formation from (a) a screw dipole, (b) cross slip of several segments in the dipole, (c) and (d) annihilation of cross-
slipped segments resulting in pinching-off a string of loops, and leaving the initial edge dislocation with two long jogs.
Fig. 9. A mechanism of dipole truncation by cross-slip annihilation producing prismatic loops from two unjogged dipoles based on Veyssie` re and Gre` gori
30
:
(a) dipole suffers local reorientation along screw components, (b) annihilation of screw portions in the cross-slip plane resulting in two closed-ended dipoles,
(c) a few segments along dipole experience reorientation similar to (a), (d) annihilation of screw components in the cross-slip plane producing an equal number
of prismatic loops.
Table III. Independent Slip Systems in FCC, BCC, and HCP
Metals, and Those Determined for (001) Single Crystal SrTiO
3
from this Study
Crystal structure Slip family Physically distinct Independent
FCC {111} /1

10S 12 5
BCC {1

10} /111S 12 5
HCP {0001} /11

20S 3 2
SrTiO
3
at room
temperature
{1

10} /110S
{001} /110S
12 5
2352 Journal of the American Ceramic SocietyYang et al. Vol. 92, No. 10
which decompose subsequently into loops under the indentation
stress. Screw dipoles stemmed from edge dislocations are gen-
erated by bowing out at edge jogs, resulting from dislocation
intersection. Edge dipoles are formed by edge trapping as in
high-temperature deformed sapphire, and by cross slip and
pinching-off as in deformed SiFe single crystal. They also ap-
pear in dipole trails, which are resulted from dragging of jogs
produced by dislocation intersection.
References
1
T. E. Mitchell and A. H. Heuer, Dislocations and Mechanical Properties of
Ceramics; pp. 341402 in Dislocations in Solids, Vol. 12, Edited by F. R. N.
Nabarro, and J. P. Hirth. Elsevier B.V., Amsterdam, the Netherlands, 2004.
2
J. R. Rice and R. Thomson, Ductile versus Brittle Behavior in Crystals,
Philos. Mag., 29 [1] 7397 (1974).
3
P. B. Hirsch and S. G. Roberts, The BrittleDuctile Transition in Silicon,
Philos. Mag., A64 [1] 5580 (1991).
4
A. S. Booth and S. G. Roberts, Dislocation Activity, Stable Crack Motion, and
the Warm-Prestressing Effect in MgO, J. Am. Ceram. Soc., 77 [6] 145766 (1994).
5
E. K. H. Salje, M. C. Gallardo, J. Jimenz, F. J. Romero, and J. Del Cerro, The
Cubic-Tetragonal Phase Transition in SrTiO
3
: Excess Specic Heat Measurements
and Evidence of a Non-Tricritical, Mean-Field Transition Mechanism, J. Phys.
Condens. Matter, 10 [25] 553543 (1998).
6
W. Sigle, C. Sarbu, D. Brunner, and M. Ru hle, Dislocations in Plastically
Deformed SrTiO
3
, Philos. Mag., 86 [2931] 480921 (2006).
7
P. Gumbsch, S. Taaeri-Baghbadrani, D. Brunner, W. Sigle, and M. Ru hle,
Plasticity and an Inverse Brittle-to-Ductile Transition in SrTiO
3
, Phys. Rev.
Lett., 87 [8] 085505-14 (2001).
8
S. Taeri, D. Brunner, W. Sigle, and M. Ru hle, Deformation Behavior
of SrTiO
3
between Room Temperature and 1800 K Under Ambient Pressure,
Z. Metallkd., 95 [6] 43346 (2004).
9
J. P. Poirier, S. Beauchesne, and F. Guyot, Deformation Mechanism of Crys-
tals with Perovskite Structure; pp. 11923 in Perovskite: A Structure of Great
Interest to Geophysics and Materials Science, Edited by A. Narvotsky, and D. J.
Weidner. American Geophysical Union, Washington, DC, 1989.
10
A. K. McNamara, S.-I. Karato, and P. E. van Keken, Localization of Dis-
location Creep in the Lower Mantle: Implications for the Origin of Seismic An-
isotropy, Earth Planet. Sci. Lett., 191 [1] 8599 (2001).
11
P. B. Hirsch and S. G. Roberts, Modeling Plastic Zone and the Brittle-Ductile
Transition, Philos. Trans.: Math. Phys. Eng. Sci., 355 [1731] 19912002 (1997).
12
S. G. Roberts, M. Ellis, and P. B. Hirsch, Dislocation Dynamics and Brittle-
to-Ductile Transitions, Mater. Sci. Eng., A164 [1] 13540 (1993).
13
P. B. Hirsch and S. G. Roberts, Comment on the Brittle-to-Ductile Transi-
tion: A Cooperative Dislocation Generation Instability; Dislocation Dynamics
and the Strain-Rate Dependence of the Transition Temperature, Acta Mater., 44
[6] 236171 (1996).
14
D. Brunner, S. Taaeri-Baghbadrani, W. Sigle, and M. Ru hle, Surprising Results
of a Study on the Plasticity in SrTiO
3
, J. Am. Ceram. Soc., 84 [5] 11613 (2001).
15
O. Bernard, M. Andrieux, S. Poissonet, and A. M. Huntz, Mechanical
Behavior of Ferroelectric Films on Perovskite Substrate, J. Eur. Ceram. Soc.,
24 [5] 76373 (2004).
16
S. J. Chang and S. M. Ohr, Dislocation-Free Zone Model of Fracture,
J. Appl. Phys., 52 [12] 717481 (1981).
17
S. M. Ohr and S. J. Chang, Dislocation-Free Zone Mode of Fracture Com-
parison with Experiments, J. Appl. Phys., 53 [8] 564551 (1982).
18
D. Hull and D. J. Bacon, Introduction to Dislocations, 3rd edition, Pergamon,
London, UK, 1984.
19
T. Matsunaga and H. Saka, TEM of Dislocations in SrTiO
3
, Philos. Mag.
Lett., 80 [9] 597604 (2000).
20
Z. Mao and K. M. Knowles, Dissociation of Lattice Dislocations in SrTiO
3
,
Philos. Mag., A73 [3] 699708 (1996).
21
Z. Zhang, W. Sigle, W. Kurtz, and M. Ru hle, Electronic and Atomic Struc-
ture of a Dissociated Dislocation in SrTiO
3
, Phys. Rev. B, 66, 214112-17 (2002).
22
P. Pauer, B. Bergk, M. Reibold, A. Belger, N. Patzke, and D. C. Meyer,
Why is SrTiO
3
Much Stronger at Nanometer Than at Centimeter Scale, Solid
State Sci., 8 [7] 78292 (2006).
23
D. S. Phillips, B. J. Pletka, A. H. Heuer, and T. E. Mitchell, An Improved
Model of Break-Up of Dislocation Dipoles into Loops: Application to Sapphire,
Acta Metall., 30 [2] 4918 (1982).
24
K. P. D. Lagerlo f, T. E. Mitchell, and A. H. Heuer, Energetics of the Break-Up
of Dislocation Dipoles into Prismatic Loops, Acta Metall., 37 [12] 331525 (1989).
25
A. S. Tetelman, Dislocation Dipole Formation in Deformed Crystals, Acta
Metall., 10 [9] 81320 (1962).
26
J. R. Low and A. M. Turkalo, Slip Band Structure and Dislocation
Multiplication in SiFe Crystals, Acta Metall., 10 [3] 21527 (1962).
27
F. Appel, H. Bethge, and U. Messerschmidt, Dislocation Motion and Mul-
tiplication at the Deformation of MgO Single Crystal in the High Voltage Electron
Microscope, Phys. Status Solidi A, 42 [1] 6171 (1977).
28
F. A. Nichols and M. W. Mullins, Morphological Changes of a Surface of
Revolution Due to Capillarity Induced Surface Diffusion, J. Appl. Phys., 36 [6]
182635 (1965).
29
J. Colin, J. Grilhe , and N. Junqua, Morphological Instability of a Stressed
Pore Channel, Acta Mater., 45 [9] 383541 (1997).
30
P. Veyssie` re and F. Gre` gori, Properties of /110]{111} Slip in Al-Rich g-TiAl
Deformed at Room Temperature II. The Formation of Strings of Prismatic
Loops, Philos. Mag., A82 [3] 56777 (2002).
31
P. B. Price, Pyramidal Glide and the Formation and Climb of Dislocation
Loops in Nearly Perfect Zinc Crystals, Phil. Mag., 5 [8] 87386 (1960).
32
A. Kelly, Strong Solids, 2nd edition, pp. 86117. Oxford University Press,
Oxford, 1973.
33
G. W. Groves and A. Kelly, Independent Slip Systems in Crystals, Philos.
Mag., 8 [89] 87787 (1963).
34
R. L. Fleischer, Number of Active Slip Systems in Polycrystalline Brass: Im-
plications for Ductility in Other Structures, Acta Metall., 35 [8] 212936 (1987).
35
R. W. K. Honeycombe, The Plastic Deformation of Metals, 2nd edition, pp.
81128. E Arnold, London, UK, 1984. &
October 2009 Deformation of SrTiO
3
by Vickers Indentation 2353

Vous aimerez peut-être aussi