Sie sind auf Seite 1von 8

Plasticity of indium antimonide between 176 and 400 C under

hydrostatic pressure. Part I: Macroscopic aspects of the deformation


B. Kedjar, L. Thilly
*
, J.-L. Demenet, J. Rabier
PHYMAT, University of Poitiers (UMR CNRS 6630), SP2MI, 86962 Futuroscope, France
Received 18 August 2009; received in revised form 26 October 2009; accepted 27 October 2009
Available online 2 December 2009
Abstract
Indium antimonide (InSb) single crystals have been plastically deformed between 176 and 400 C, i.e. below and above the brittle-to-
ductile transition temperature situated around 150160 C, via the use of microindentation below room temperature (RT) and the Pat-
erson press (compression under gaseous pressure) above RT. The evolution of the macroscopic mechanical data (hardness and critical
resolved shear stress) with temperature suggests the existence of three deformation regimes with transitions at T
tr1
= 150 C and
T
tr2
= 20 C. T
tr1
coincides with the brittle-to-ductile temperature, while T
tr2
may coincide with a transition in the nature of dislocations
with dislocations propagating in the glide set above T
tr2
while moving in the shue set below T
tr2
.
2009 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Semiconductor; Indium antimonide; Brittle-to-ductile transition; Compression; Indentation
1. Introduction
The development of technological applications of semi-
conductors (SCs) has created demand for electronic devices
of ever-increasing performance. This requirement is one of
the driving forces for the interest of the scientic commu-
nity in the properties of SCs. Although optoelectronic
properties have been widely studied and improved via the
control of crystal growth processes of elemental and com-
pound SCs, the defects that are created during fabrication
still appear highly detrimental to the devices functionality
and usually lead to their untimely degradation. Moreover,
the survival of devices is highly dependent on the function-
ing conditions since defect density may evolve with temper-
ature as dierential dilatation is produced in the devices,
which in turn creates stress concentrations. As a conse-
quence, knowledge of the mechanical behaviour of SCs
and their deformation mechanisms is still of great impor-
tance for the design of improved devices.
In addition to this technological aspect, SCs have long
been considered as model materials on which to study the
eects of temperature, strain rate and impurity content
on deformation mechanisms, and in particular on disloca-
tion nucleation and propagation. SCs serve as model mate-
rials because (i) the fabrication of large single crystals of
SCs has been undertaken for many decades and (ii) SCs
exhibit a brittle-to-ductile transition (BDT, usually situated
around 0.6T
m
, where T
m
is the melting temperature) that is
still the subject of investigation. Moreover, with the
increasing miniaturization of devices, new properties have
been reported: for instance, the mechanical properties of
nanocrystalline materials or micro- and nano-objects have
been improved (the HallPetch strengthening with grain
size reduction is an example of such eect). Therefore, dur-
ing the last decade, the study of small objects has devel-
oped, such as in microelectromechanical systems
(MEMS), which are fabricated from silicon wafers by etch-
ing techniques or focused ion beam (FIB) milling. In such
small-scale objects, complex stress tensors are developed
and the deformation response may dier from that found
in bulk systems. As an example, Michler and co-workers
have recently fabricated micrometer-sized columns (so-
1359-6454/$36.00 2009 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2009.10.050
*
Corresponding author. Tel.: +33 5 49 49 68 31; fax: +33 5 49 49 66 92.
E-mail address: ludovic.thilly@univ-poitiers.fr (L. Thilly).
www.elsevier.com/locate/actamat
Available online at www.sciencedirect.com
Acta Materialia 58 (2010) 14181425
called micropillars) by FIB milling a GaAs wafer and
compressed these objects in a scanning electron microscope
at room temperature (RT) [1]: they observed that pillars
with a diameter of 10 lm are brittle and fail in compression
without experiencing plastic deformation, while pillars with
a diameter of 1 lm exhibit plasticity (with nucleation of
Shockley partial dislocations and microtwinning). The sup-
pression of brittleness is not understood but could be
related to a complex stress tensor inhibiting crack propaga-
tion, thereby favouring plasticity. As a matter of fact, all
these elements show that the study of elemental deforma-
tion mechanisms in SCsespecially in high-stress
domainsis valuable from the theoretical and applied
points of view.
In this study, the IIIVSCcompound indiumantimonide
(InSb) has been chosen because of its low melting tempera-
ture (T
f
= 525 C = 798 K). This characteristic enables InSb
tocover a wide temperature range, andis especially useful for
studying the deformation microstructures above and below
the BDT temperature. InSb is narrow-band-gap SC with
E
g
0.2 eV and crystallizes, as most IIIV compounds, in
the zincblende structure (space group F 43m) based on a
face-centred cubic lattice with In at (0, 0, 0) and Sb at (1/
4, 1/4, 1/4). This structure canbe seenas the superimposition
of {1 1 1} planes made exclusively of Inor Sbatoms: InSbis a
polar material as illustratedby Fig. 1. Moreover, the distance
between adjacent {1 1 1} planes can be of two types: two
planes which are closely spaced dene the glide set
(Fig. 1b); otherwise they dene the shue set. InSb under-
goes several phase transitions under pressure: the rst tran-
sition pressure, P
s
= 2 GPa, corresponds to a SC-to-metal
transition (similar to b-Sn) [2]. The most complete phase-
temperature diagrams can be found in Refs. [3,4].
The macroscopic mechanical properties of InSb have
been studied since the 1960s using various deformation
techniques such as 3-point or 4-point bending [5,6], uniax-
ial compression at room pressure [712], uniaxial compres-
sion under solid conning pressure (in the Griggs
apparatus) [13,14] and microindentation [15,16]. It should
be pointed out that none of the previously cited studies
reports a systematic study of InSb mechanical properties
over the entire temperature range. Nevertheless, when
taken together, the macroscopic data show overall that:
(i) the yield stress of InSb increases rapidly below 300 C,
suggesting that dislocation glide is highly thermally acti-
vated [11] and (ii) when deformed at room pressure, InSb
becomes brittle below 150160 C [11]. With compression
under solid conning pressure, Suzuki and co-workers
deformed InSb to 130 C, and evidenced at a temperature
T
tr
, close to RT, a change in the yield stress evolution vs.
temperature: this so-called hump has been discussed in
terms of a change of deformation mechanism with disloca-
tions propagating in the glide set above RT while moving
in the shue set at very low temperature [14].
This study is divided into two parts: in the present paper
(part I), the macroscopic mechanical properties of InSb are
revisited from 400 down to 176 C by means of two com-
plementary deformation techniques: uniaxial compression
under gaseous conning pressure from 400 C down to
RT and microindentation below RT. From these experi-
ments, the evolution of the yield stress with temperature
is presented and compared to available literature data.
Observations of slip traces on the surfaces of the samples
are also reported. All these results are nally discussed to
propose a schematic view of the yield stress evolution with
temperature in the context of the appearance of the BDT.
Part II of this study is devoted to the microscopic char-
acteristics of the deformation of InSb, and the dislocation
microstructures are characterized from 400 down to
176 C.
2. Experimental
In this study, bulk S-doped InSb single crystal
(n = 4.88 10
16
cm
3
) has been chosen for the compres-
sion tests under hydrostatic pressure. The bulk single crys-
tal was cut into parallelepipeds with typical size
3 3 8 mm
3
. All faces were carefully mechanically pol-
Fig. 1. (a) Zincblende structure (space group F 43m) of InSb, based on a face-centred cubic lattice with In at (0, 0, 0) and Sb at (1/4, 1/4, 1/4). (b) Structure
observed along the 110 direction, showing the superimposition of {1 1 1} planes made exclusively of In or Sb atoms. The distance between adjacent
{1 1 1} planes can be of two types: two planes closely spaced are the glide set; otherwise they are the shue set.
B. Kedjar et al. / Acta Materialia 58 (2010) 14181425 1419
ished and subsequently chemically polished to remove any
surface defects that would act as surface dislocation
sources and hinder the observation of in-volume disloca-
tions via the slip traces that they create when emerging at
the sample surface. The samples were inserted into fully
annealed aluminium cylinders to t into the deformation
assembly of the compression apparatus. The compression
axis was chosen to be 213 to favour single slip on the
a/2011111 slip system (the lateral surfaces are com-
posed of 111, 111, 451 and 451 faces). Uniaxial
compression tests were performed in the Paterson appara-
tus [17,18] under an argon conning pressure of 300 MPa
at a strain rate of 2 10
5
s
1
, from RT up to 400 C.
After deformation, the samples were removed from the
Al jackets for optical observation of the slip traces at sam-
ples surface, and thin foils were cut parallel to the 111
primary glide plane for transmission electron microscopy
(TEM) observations of the deformation microstructure
(see part II). For the low-temperature mechanical charac-
terization, h1 1 1i-oriented Te-doped InSb wafer was
selected for microindentation tests; both 111 In-atom-
terminated face and 111 Sb-atom-terminated face have
been indented. For this purpose, a Shimadzu microindenter
was installed inside a controlled-atmosphere chamber lled
with argon to avoid ice formation at the sample surface
upon cooling, while a liquid nitrogen circulation was
installed under the sample holder to decrease the sample
temperature from RT down to 176 C. Several hundred
Vickers indentations were performed at dierent tempera-
tures under a maximum load of 50 g (with holding time
of 10 s) and the hardness was calculated from the in situ
measurement of the dimensions of the indents. After inden-
tation, the InSb wafer was optically observed to check for
the presence of cracks and subsequently thinned for TEM
observations of the deformation microstructure (see part
II). In both experiments, the temperature was monitored
during mechanical testing with accuracy better than 2 C.
3. Results
3.1. Compression tests
A number of tests were performed at 20, 100, 150, 300
and 400 C, i.e. below and above the BDT temperature
(about 150 C). Because the samples are inserted into Al
cylinders, the mechanical data that are recorded during
the tests comprise the results from both InSb and Al. To
remove the contribution of the Al jacket, a procedure
was developed that consists of preliminary tests on pure
Al samples under identical conditions to those used for
the InSb + Al tests. The obtained stressstrain Al curve
is then subtracted from the InSb + Al mechanical data fol-
lowing a rule of mixture (ROM) that takes into account the
InSb and Al volume fractions. The validity of the ROM
law is based on the fact that the two phases, InSb and
Al, are tested in parallel without interactions at interfaces
and they obey isostrain conditions: the validity of the
ROM has been veried by nite-element modelling
(FEM) of the compression tests under hydrostatic pressure
[19]. In addition, the stiness of the compression apparatus
and its temperature dependence have been taken into
account. As a consequence, the mechanical properties of
InSb are obtained in the form of stressstrain curves as
reported in Fig. 2, where the engineering stress r
InSb
is plot-
ted vs. engineering strain e
InSb
at 20, 100, 150, 300 and
400 C. The mechanical properties appear highly tempera-
ture dependent with increasing strength as the temperature
decreases. Furthermore, two behaviours can be distin-
guished, indicated by the presence or absence of a stress
peak before a hardening phase. In the rst case (at 20,
100, 150 and 300 C), the presence of this upper and lower
yield stress (UYS and LYS) is generally attributed to a low
initial dislocation density and the need for higher stress to
trigger dislocation sources followed by dislocation ava-
lanches that lead to a reduction in the macroscopic stress.
The absence of the stress peak at 400 C may be related
to the remaining surface defects acting as dislocation
sources that hinder the avalanche phenomenon. In addi-
tion, at this temperature, thermal activation could be high
enough to stimulate activity of dislocation sources emitting
dislocations in a sucient number to accommodate
mechanical distortion. To study the evolution of the elastic
limit vs. temperature, we compare the LYS (at 20, 100, 150
and 300 C) to the stress at intersection between the elastic
regime and the constant hardening regime (at 400 C). As a
result, the critical resolved shear stress (CRSS), s
c
, is calcu-
lated at each temperature as the resolved stress in the pri-
mary 111 slip plane, i.e. the LYS multiplied by the
Schmid factor, m, equal to 0.47 (for a a/2011111 sys-
tem and stress applied along 213).
The values of s
c
are plotted in Fig. 3 as a function of
the absolute temperature, T, and can be compared to lit-
erature data on InSb deformed by compression under
Fig. 2. Stressstrain curves obtained from the compression of InSb single
crystals at 20, 100, 150, 300 and 400 C, after correcting from the inuence
of the Al jacket and the temperature dependence of the stiness of the
compression apparatus.
1420 B. Kedjar et al. / Acta Materialia 58 (2010) 14181425
hydrostatic pressure in the Griggs apparatus (from 373
down to 143 K) [14], i.e. where the hydrostatic pressure
is obtained by compression of a solid medium (lead or
salt). Our data are systematically higher than the ones
obtained in the Griggs apparatus: rst, unlike the Pater-
son apparatus, where the load cell is installed inside the
pressure vessel, the data from the Griggs apparatus are
measured by an external load cell and are potentially
screened or smoothed by elevated solid/solid friction
between sample and conning medium; second and most
importantly, as shown by FEM [19], the friction between
sample and solid conning medium creates unavoidable
shear stresses at the sample surface (and corners) when
pressurized: the resulting Von Mises stresses lead to early
yielding of the sample during subsequent axial compres-
sion. The apparent yield stresses are then reduced com-
pared to the ones obtained in the Paterson apparatus
(gaseous conning pressure). This eect is more pro-
nounced in InSb at high temperature because of its
strong softening above the BDT temperature. Therefore,
the data obtained here are considered to be reliable.
After removal from the Al jacket, the deformed InSb
parallelepipeds were optically inspected on their lateral
faces to observe the slip traces. Beside the primary a/
2011111 slip system with a Schmid factor m = 0.47,
two secondary slip systems can be equivalently activated
during a [2 1 3] compression: a/2[1 1 0] 111 and a/
2011 (1 1 1) with m = 0.35. A ternary slip system may
as well be activated, a/2[1 0 1] 111 with m = 0.29.
Fig. 4 presents optical micrographs for each lateral face,
111, 111, 451 and 451, of InSb samples deformed
at 400, 300, 150, 100 and 20 C; in addition, theoretical
directions of slip traces are added. For the latter tempera-
ture, Fig. 5 exhibits a high magnication view of the 111
face. Table 1 summarizes the slip trace observations: after a
plastic deformation of 4% at 400 C, only traces of the pri-
mary slip system are observed to be homogeneously dis-
tributed on the samples surface, with no trace of cracks.
After a plastic deformation of 5% at 300 C, the primary
and one secondary (a/2011(1 1 1)) slip systems are
observed on all lateral faces and the ternary slip system is
additionally observed on the 111 face. Again no cracks
are observed. After a plastic deformation of 7% at
150 C, the 111 and 451 faces exhibit traces of the pri-
mary and one secondary (a/2011(1 1 1)) slip systems; the
111 face presents traces of the primary and the ternary
slip systems while only traces of the primary slip system
are observed on the 451 face. After a plastic deformation
of 6% at 100 C, all faces exhibit traces of the primary and
one secondary (a/2011(1 1 1)) slip systems; the ternary
slip system is additionally observed on the 111 face.
After a plastic deformation of 6% at 20 C, all faces pre-
sents traces of the primary and the a/2011(1 1 1) second-
ary slip systems. In addition, several slip traces from the
primary slip system on the 111 face exhibit double
cross-slip events (as indicated in Fig. 5), leading the average
direction of the slip lines to deviate from the 41 angle from
the [2 1 3] compression axis. It must be emphasized that
evidence of double cross-slip could only be found after
RT deformation.
3.2. Microindentation tests
A series of Vickers indentations was performed at RT
and 176 C on both 111 In-atom-terminated and
111 Sb-atom-terminated faces. At RT, the In-face
exhibits an average Vickers hardness HV = 2.2 0.1 GPa,
while for the Sb-face, HV = 2.3 0.1 GPa. At 176 C,
the average hardness for the In-face and Sb-face are respec-
tively HV = 3.2 0.1 GPa and HV = 3.3 0.1 GPa. The
observation of a softer Sb-face compared to the In-face
has been similarly observed at and above RT [16]. In the
following, the average hardnesses of the In- and Sb-faces
are considered.
After indentation, the InSb wafer was mechanically pol-
ished from the opposite surface before ion milling to nally
obtain electron transparency for TEM analysis of the
deformation microstructure. After mechanical polishing,
a very regular network of cracks could be observed to be
aligned with h2 1 1i in-plane directions, as shown on the
micrograph displayed in Fig. 6.
3.3. Evolution of mechanical properties between 176 and
400C
To obtain a better insight into the temperature depen-
dence of the macroscopic mechanical properties of InSb,
the compression and indentation data are plotted together
as ln(s
c
) or ln(HV) as a function of 1000/T (Fig. 7). This
representation provides an estimation of the apparent acti-
vation energy for dislocation nucleation, Q/n (where Q is
the activation energy and n is the stress exponent), follow-
ing the phenomenological relation:
Fig. 3. Evolution of s
c
vs. the absolute temperature, T, for the present
study (InSb deformed by compression under gaseous conning pressure)
and compared to literature data on InSb deformed by compression under
hydrostatic solid pressure in the Griggs apparatus [14].
B. Kedjar et al. / Acta Materialia 58 (2010) 14181425 1421
ln s
c
Q=nkT 1=nln _ c ln A 1
or equivalently:
ln HV Q=nkT 1=nln _ c ln B 2
where k is the Boltzmann constant, and A and B depend on
the shear strain rate _ c.
Q/n is then proportional to the slope of the curve
ln(s
c
) = f(1/T) or ln(HV) = f(1/T). Fig. 7 also shows the lit-
erature data on InSb deformed in the Griggs apparatus
(from 100 down to 130 C) [14] and InSb indented above
20 C [15] (following Tabors empirical law, the hardness is
proportional to the yield stress, with HV 3r
y
).
The dierence between the Griggs-apparatus data and
Paterson-apparatus data has been discussed earlier; the dif-
ference between the two indentation data sets at RT may
be attributed to the doping dierence (n-type in this study
vs. undoped in Ref. [14]) that is known to modify the
mechanical behaviour of SCs, and InSb in particular, with
p-doped stronger than undoped, which is in turn stronger
that n-doped [20]. Our value of s
c
at 150 C is unexpectedly
low: several compression tests have been performed at this
temperature and all yield the same value.
Table 2 summarizes the apparent activation energies Q/
n that can be deduced from the slopes of the curves
observed in Fig. 7; several other literature data have also
been analyzed but are not shown on this gure for the sake
Fig. 4. Overview of the optical micrographs for each lateral face, 111, 111, 451 and 451, of InSb samples deformed at 400, 300, 150, 100 and
20 C; theoretical directions of slip traces are added.
Fig. 5. High-magnication view of the 111 face after a plastic
deformation of 6% at 20 C, where several slip traces from the primary
slip system exhibit double cross-slip events, as encircled.
1422 B. Kedjar et al. / Acta Materialia 58 (2010) 14181425
of clarity. From Fig. 7 and Table 2, it appears that three
distinct regimes can be distinguished for all data sets, with
strong variations of apparent activation energy at 150 and
20 C, with Q/n reducing by a factor of 43 between 400 and
176 C. In the following, these three regimes will be
labelled as the high-temperature (HT) regime above
150 C, the low-temperature (LT) regime between 150
and 20 C, and the very-low-temperature (VLT) regime
below 20 C. In addition, the two transition temperatures
will be labelled as T
tr1
= 150 C = 423 K and
T
tr2
= 20 C = 293 K.
4. Discussion
4.1. Validity of data comparison
The use of a single deformation technique spanning the
VLT and HT ranges was not possible in the present study,
and therefore compression tests under hydrostatic pressure
above RT were complemented by indentation tests below
RT. Moreover, the literature data relative to InSb also
combine several deformation techniques ranging from
compression tests with or without conning pressure,
Table 1
Observation of slip traces on the lateral faces of samples after compression at dierent temperatures.
Temperature
(C)
Lateral
faces
Slip traces
Primary system [0 1 1] 111
(m = 0.47)
Secondary system [1 1 0]
111 (m = 0.35)
Secondary system 011
(1 1 1) (m = 0.35)
Ternary system [1 0 1] 111
(m = 0.29)
400 111
111
451
451
300 111
111
451
451
150 111
111
451
451
100 111
111
451
451
20 111
111
451
451
Fig. 7. Ln(s
c
) or ln(HV) as a function of 1000/T. Literature data on InSb
deformed in the Griggs apparatus (from 100 down to 130 C) [14] and
InSb indented above 20 C [15] have been added.
Fig. 6. InSb wafer after indentation and mechanical polishing from the
opposite surface: a regular network of cracks is observed to be aligned
with h2 1 1i in-plane directions.
B. Kedjar et al. / Acta Materialia 58 (2010) 14181425 1423
bending tests and indentation. Therefore, it is crucial to
investigate the possibility of comparing the dierent data
sets prior to deeper analysis. In particular, the complexity
of the applied stress tensor must be discussed since the
presence of hydrostatic and deviatoric components has a
strong impact on the sample as the rst component may
lead to phase transition if it reaches the materials rst tran-
sition pressure, P
s
(P
s
= 2 GPa for InSb [2]) and the second
is the cause of plastic deformation by shear.
Let us consider the most unfavourable case, i.e. indenta-
tion, where the hydrostatic and deviatoric components may
be simultaneously elevated. For this discussion, Si is chosen
as a model material since the literature data on the VLT to
HT regimes is vast. It has been claimed for many decades
that Si undergoes a phase transition when indented. This
assertion comes from the early work of Gridneva et al.
[21], which monitored the electrical resistance of Si upon
indentation as a function of applied load: as a resistance
decrease was observed with an increase in load, the authors
concluded that phase transformation occurs at high load
(i.e. high pressure) with the formation of a metallic phase
(P
s
(Si) = 12 GPa). Moreover, they observed a decrease in
the temperature dependence of the Si hardness below
400 C: it was then concluded that this hardness plateau
at LT was related to uncontrolled phase transformation.
Very recently, Khayyat [22] developed a similar electrical
monitoring during indentation and unambiguously showed
that the deformation aects the nature of electrical contacts
at samples surface, leading to the formation of ohmic con-
tacts that, as an artefact, induce a decrease in the measured
resistance. Combining this experiment with Raman spec-
troscopy, the same author showed that plastic deformation
and phase transformation coexist in Si indented at RT, but
no phase transformation is observed in Si when indented at
73 C. The latter result suggests that the rst transition
pressure of Si increases below RT. The coexistence of phase
transformation and plasticity and its impact on the hard-
ness of Si and Ge, between RT and 400 C, has been exam-
ined in detail by Vandeperre and co-workers via the
calculation of the stress tensor under the indentor [23]: if
the transition pressure is larger than two-thirds of the elas-
tic limit of the non-transformed material, plasticity occurs
simultaneously in the transformed and non-transformed
regions. Therefore, the hardness plateau of Si at LT cannot
be only imputed to a phase-transformation artefact. Addi-
tionally, we may compare the mechanical properties of Si
obtained with dierent techniques such as indentation
[15] and deformation under high pressure in a Griggs appa-
ratus [24] and in a multi-anvil setup [25]: the latter experi-
ments have been performed under a conning pressure of
respectively 1.5 and 5 GPa, i.e. well below the transition
pressure for Si. Fig. 8 is a plot of ln(HV) and ln(s
c
) as a
function of 1000/T for the three data sets: they present a
plateau below 200300 C that, again, cannot be only
imputed to phase-transformation.
As a consequence, it is possible to compare dierent
data sets and the observation of a hardness plateau at LT
and VLT is not indicative of phase transformation.
4.2. Comparison of InSb mechanical data with literature
Fig. 7 enables data obtained with dierent techniques to
be compared: independently from the data set, three dis-
tinct regimes can be distinguished with transition tempera-
tures at T
tr1
= 150 C = 423 K and T
tr2
= 20 C = 293 K.
Above T
tr1
= 150 C, the apparent activation energy is
estimated to be of the order of 0.1 eV whatever the defor-
mation technique used (Table 2) with a decrease by a factor
of 4 below T
tr1
down to T
tr2
= 20 C. Below T
tr2
, the
apparent activation energy decreases again by a factor of
more than 10. Following the work of Karmouda [11], the
BDT temperature of InSb is generally considered to take
place around 150160 C: the observation of a transition
of the mechanical properties at T
tr1
= 150 C may therefore
be correlated to the BDT. If so, the change of activation
energy at T
tr1
is indicative of a change in the deformation
mechanism at the BDT. It should be noted that a similar
decrease in the apparent activation energy was observed
Table 2
Estimated apparent activation energies Q/n calculated from the literature
and from the present study data sets.
Reference Q/n (eV)
T > 150 C 20 C < T < 150 C T < 20 C
Shimizu and Sumino
[6]
0.320.4
Karmouda [11] 0.20.31
Branchu [13] 0.140.17
Suzuki et al. [14] 0.190.32 0.0118
0.0193
Yonenaga and Suzuki
[15]
0.170.21 0.0480.08
Present study 0.140.25 0.0350.07 0.0045
Fig. 8. Ln(HV) and ln(s
c
) as a function of 1000/T for Si obtained with
dierent deformation techniques such as indentation [15] and deformation
under high pressure in a Griggs apparatus [24] and in a multi-anvil setup
[25].
1424 B. Kedjar et al. / Acta Materialia 58 (2010) 14181425
in GaAs between 315 and 350 C and correlated to a
change in dislocation nature at the BDT taking place at
this temperature [26]. A further decrease in the activation
energy at T
tr2
may also be the signature of another change
in the deformation mechanism at VLT. This temperature
domain is not very well documented because of the brittle-
ness of the SCs and the elevated stress level that is required
at such low temperatures. Only Suzuki et al. have also suc-
ceeded in deforming InSb below RT and they too observed
a sudden change in mechanical data evolution vs. temper-
ature that they qualied as a hump [14]. The authors sug-
gested that this transition could result from a change in
dislocation nature, with dislocations propagating in the
glide set above T
tr2
while moving in the shue set at
VLT. Nevertheless direct observation of the dislocations
at VLT cannot be found for InSb. The only observations
of this glide-to-shue transition are available for Si where
perfect non-dissociated screw dislocations were observed
after deformation at 20 C in a multi-anvil setup [2730].
Here, the observations of double cross-slip on the surface
of the InSb samples after RT deformation could be indica-
tive of propagation of non-dissociated perfect screw dislo-
cations at VLT.
5. Summary and conclusion
InSb single-crystalline samples have been successfully
plastically deformed from 400 down to 176 C by the
use of complementary deformation techniques, i.e. macro-
scopic compression of bulk samples under gaseous conn-
ing pressure above RT, and indentation below RT. The
study of the evolution of macroscopic mechanical data
vs. temperature leads to the following assertions:
1. A plot of ln(s
c
) or ln(HV) vs. 1/T exhibits three regimes
with transitions at T
tr1
= 150 C = 423 K and
T
tr2
= 20 C = 293 K.
2. The slope of ln(s
c
, HV) = f(1/T) in each regime is related
to the apparent activation energy for dislocation nucle-
ation: this energy decreases strongly at T
tr1
and T
tr2
,
indicating a change in deformation mechanism.
3. T
tr1
coincides with the brittle-to-ductile temperature that
is usually reported around 150160 C.
4. InSb samples deformed at RT exhibit double cross-slip
traces at their surfaces, suggesting the occurrence of easy
cross-slip that could be related either to available con-
strictions on dissociated screw dislocation or the pres-
ence of non-dissociated screws.
5. T
tr2
could therefore coincide with a transition in the nat-
ure of dislocations with dislocations propagating in the
glide set above T
tr2
while moving in the shue set at
VLT.
These assertions need careful analysis of the deformation
microstructure to be validated: this is the objective of the
part II of this study.
References
[1] Michler J, Wasmer K, Meier S, O

stlund F, Leifer K. Appl Phys Lett


2007;90:043123.
[2] Jayaraman A, Newton RC, Kennedy GC. Nature 1961;191:1288.
[3] Mezouar M, Besson JM, Syfosse G, Itie JP, Hau sermann D,
Hanand M. Phys Status Solidi B 1996;198:403.
[4] McMahon MI, Nelms RJ, Allan DR, Belmoonte SA, Bovornratan-
araks T. Phys Rev Lett 1998;80:5564.
[5] Peissker E, Haasen P, Alexander H. Philos Mag A 1962;80:1279.
[6] Shimizu H, Sumino K. Philos Mag A 1975;50:733.
[7] Schafer S, Alexander H, Haasen P. Phys Status Solidi 1964;5:247.
[8] Sazhin NP, Milvidskii MG, Osvenskii VB, Stolyarov OG. Sov Phys
Solid State 1966;8.
[9] Osvenski VB, Stolyarov OG, Milvidskii MG. Sov Phys Solid State
1969;2540.
[10] Gottschalk H, Patzer G, Alexander H. Phys Status Solidi A
1978;45:207.
[11] Karmouda M. PhD thesis, University of Lille, France; 1984.
[12] Ferre D. PhD thesis, University of Lille, France; 1987.
[13] Branchu S. PhD thesis, University of Poitiers, France; 1997.
[14] Suzuki T, Yasutumi T, Tokuoka T, Yonenaga I. Phys Status Solidi A
1999;171:47.
[15] Yonenaga I, Suzuki T. Philos Mag Lett 2002;82:535.
[16] Roberts SG. Philos Mag A 2000;54:37.
[17] Paterson MS. Int J Rock Mech Miner Sci 1970;7:517.
[18] Paterson MS, Olgaard DL. J Struct Geol 2000;22:1341.
[19] Mussi A, Thilly L, Rabier J, Demenet JL. Mater Sci Eng A
2008;478:140.
[20] Patel JR, Chaudhuri AR. Phys Rev 1966;143:2.
[21] Gridneva IV, Milman YV, Trelov VI. Phys Status Solidi A
1972;14:177.
[22] Khayyat MMO, Hasko DG, Chaudhri MM. J Appl Phys
2007;101:083515.
[23] Vandeperre LJ, Giuliani F, Lloyd SJ, Clegg WJ. Acta Mater
2007;55:6307.
[24] Demenet JL. PhD thesis, University of Poitiers, France; 1987.
[25] Rabier J, Renault PO, Eyidi D, Demenet JL, Chen J, Couvy H, et al.
Phys Status Solidi C 2007;15:3110.
[26] Wang S, Pirouz P. Acta Mater 2007;55:5500.
[27] Rabier J, Cordier P, Tondellier T, Demenet JL, Garem H. J Phys
Condens Matter 2002;12:10059.
[28] Rabier J, Demenet JL. Scr Mater 2001;45:1259.
[29] Rabier J, Denanot MF. Mater Sci Eng A 2004;387389:124.
[30] Rabier J, Demenet JL, Denanot MF, Milhet X. Mater Sci Eng A
2004;400401:97.
B. Kedjar et al. / Acta Materialia 58 (2010) 14181425 1425

Das könnte Ihnen auch gefallen