Sunteți pe pagina 1din 9

Acta Materialia 51 (2003) 4487–4495

www.actamat-journals.com

Enhanced growth of intermetallic phases in the Ni–Ti


system by current effects
J.E. Garay, U. Anselmi-Tamburini 1, Z.A. Munir ∗
Department of Chemical Engineering and Materials Science, University of California, One Shields Avenue, Davis,
CA 95616, USA

Received 13 December 2002; received in revised form 13 December 2002; accepted 9 May 2003

Abstract

The effect of direct current upon interfacial reactions in the Ni–Ti system was investigated. Isothermal diffusion
couple experiments were conducted under varying current densities to de-couple Joule heating from intrinsic effects
of the current flux. Current densities of up to 2546 A cm⫺2 were used in the temperature range of 625–850 °C. All
of the intermetallic compounds (NiTi, Ni3Ti and NiTi2) present in the equilibrium phase diagram were identified in
the product layer. In addition, β-Ti solid solutions formed in samples annealed above the α→β temperature, 765 °C.
The growth of all product layers was found to be parabolic and the applied current was found to significantly increase
the growth rate of the intermetallic layers. Using Wagner’s analysis the present results were compared to published
results on current-free diffusion couples. The intrinsic growth rate constant of the NiTi2 intermetallic was found to be
43 times higher under the influence of 2546 A cm⫺2 than that obtained without a current at 650 °C. The effective
activation energy for the formation of all phases was found to decrease with increasing current density. The effect was
strong for all phases but the decrease was most marked for Ni3Ti. In this case, the activation energy decreased from
292 kJ mol⫺1 under the influence of a current density of 1527 A cm⫺2 to 86 kJ mol⫺1 when the current density was
2036 A cm⫺2. The results are explained in terms of current induced changes in the growth mechanism arising from
changes in the concentration of point defects or their mobility.
 2003 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Electric current effects; Intermetallics; Growth rates; Titanium–nickel; Reactive diffusion

1. Introduction and consequently their synthesis by self-sustaining


combustion has been limited [1]. Activation by
With a few exceptions, the enthalpy of forma- thermal means (e.g., the preheating of the elemen-
tion of most intermetallic phases is relatively low tal reactants) and by electrical means (the impo-
sition of a field) has been utilized to overcome the
thermodynamic limitation [2]. In the former case,

Corresponding author. Tel.: +1-530-7520554; fax: +1- the product is often multi-phase, containing sig-
530-7528058.
E-mail address: zamunir@ucdavis.edu (Z.A. Munir). nificant amounts of phases not intended by the
1
Permanent address: Dipartimento di Chimica Fisica, Univ- starting stoichiometry. The use of an electric field
ersità di Pavia, V.le Taramelli, 16, 27100 Pavia, Italy. (in reality a current) has been shown not only to

1359-6454/03/$30.00  2003 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/S1359-6454(03)00284-2
4488 J.E. Garay et al. / Acta Materialia 51 (2003) 4487–4495

have a significant role in activating the process, but in the Ni–Ti system (Fig. 1) Ni3Ti, NiTi and NiTi2.
also in influencing the nature of the product. In an Ni3Ti is a hexagonal close-packed superlattice and
investigation on the synthesis of Ti3Al, it was NiTi2 has a complex face-centered cubic structure
found that the rate of the reaction increased with with 96 atoms per unit cell (space group Fd3m)
an increase in the applied field and that the nature [15,16]. The structure of the third equilibrium
of the product changed with this increase. Ulti- phase NiTi is interesting in that it has a high and
mately at a relatively high field, only the desired low temperature crystal structures due to a mar-
Ti3Al phase was present [3]. Similar results were tensitic-type transformation. This transformation
obtained in the case of another intermetallic, leads to a shape memory effect. The high tempera-
FeAl [4]. ture “austenitic” phase has a B2 (CsCl) structure
In these and other similar observations, the role (space group Pm3m), while the low temperature
of the current was initially assumed to be thermal, “martensite” has a more complex monoclinic
providing Joule heat to the reactants and product. B19⬘-type structure (space group P21/m) [17].
Modeling studies based on this assumption have
been made and the results show qualitative agree-
ments with the experimental observations [5–7]. 2. Experimental procedure
However, non-thermal current effects on mass
transport through electromigration have been A schematic of the apparatus we used in our
known for decades and experimental observations experimental investigation is shown in Fig. 2. It
demonstrating such effects have been provided consists of a vertical cylindrical vacuum chamber
[8,9]. In a recent investigation, we have shown that housing two electrodes surrounded by a furnace.
the imposition of a DC current across interfaces The electrodes carry current to the sample and pro-
between Al and Au metal layers has a significant vide a small mechanical load to ensure good con-
effect on the growth rate of the intermetallic layers tact between the reactants. The electrodes are mol-
and on the incubation time for the nucleation of ybdenum rods 28 mm in diameter with a terminal
these product phases [10,11]. In these results, no part reduced to 5 mm diameter, allowing the exter-
effect of the direction of the DC current on the nal DC power supply to produce a current density
growth of the products was observed, implying that up to 3310 A cm⫺2 through the sample. The fur-
electromigration is not the dominant effect of the nace consists of a cylindrical tungsten mesh 16.5
current. cm in diameter and 6.5 cm long powered by an
In this paper, we investigate the effect of the AC power supply. A 0.5-mm diameter type K
current on the phase formation kinetics of products
resulting from the interaction of Ti and Ni and we
compare our results with those reported earlier on
diffusion couples without a current [12].
The Ni–Ti system has received much attention
primarily for its commercially important intermet-
allic phase, NiTi. This material has relatively high
ductility and corrosion resistance in addition to dis-
playing the shape memory effect. Other alloys in
the system have received attention, including a
quasi-crystalline phase which can be formed near
the NiTi2 compound, and also because Ti-rich Ni–
Ti alloys can form metallic glasses [13]. Also of
interest is replacing Al with Ti in Ni-based superal-
loys, because Ti is believed to enhance the high
temperature strength [14].
There are three equilibrium intermetallic phases Fig. 1. Ni–Ti phase diagram.
J.E. Garay et al. / Acta Materialia 51 (2003) 4487–4495 4489

sition of the products in the reaction layer was


determined through electron microprobe analyses
(EPMA).

3. Results

In the experimental conditions we have investi-


gated, the interaction between Ni and Ti always
produced well-defined product layers that grew
parabolically. The composition of these layers cor-
responded to all the intermetallic phases present in
the phase diagram (Fig. 1). Samples which reacted
above 765 °C also contained a β-Ti solid solution
layer. An example of a product layer morphology
is shown by the SEM micrograph in Fig. 3. An
important observation concerning the formation of
Fig. 2. Schematic of experimental apparatus. these layers was that it was independent of the
direction of the current, i.e., identical layer growth
(chromel–alumel) thermocouple placed in contact was observed on both interfaces. Fig. 4 shows a
with both the sample and one electrode was used typical trend in the growth of the individual phase
to monitor temperature and provide feedback to the layers (NiTi2, NiTi, and Ni3Ti) and the total pro-
PID furnace controller. The experiments were per- duct at 700 °C when the reaction was carried out
formed under high vacuum (ⱕ5 × 10⫺7 Torr). without an electric current. These results show a
The samples consisted of three metallic layers parabolic growth for all three phases. Furthermore,
with an Ni foil sandwiched between two Ti foils. there is a clear indication that an incubation time
The foils were circular in shape, 5 mm in diameter is required for the nucleation of these phases. With
and of 127 µm thick. All foils were reported to be two exceptions, these observations are in partial
99.99% pure (ACI Alloys, San Jose, CA). The agreement with the results of the only reported dif-
three-layer assembly provides both a Ti–Ni and fusion couple study on Ti–Ni system [12]. The
Ni–Ti interface with respect to the direction of the exceptions relate to the incubation time and to the
DC current, thus providing the opportunity to kinetics of the growth of one of the phases. In the
investigate the influence of current direction on
interfacial reactions. In addition, molybdenum foils
were used as protection layers between the elec-
trodes and the sample ensemble. Prior to each
experiment, the foils were polished using 600 and
1200 grit SiC metallographic paper and cleaned
with acetone.
Experiments were conducted under varying time
(0.5–30 h), current density (0–2546 A cm⫺2), and
temperature (625–850 °C). At the end of each
experiment, the cooled samples were mounted in
conductive epoxy resin (Buehler Konductomet),
sectioned perpendicularly to the metal interfaces,
and prepared metallographically. The thickness of Fig. 3. Back-scattered electron micrograph showing typical
each of the intermetallic layers was measured using product layers. Sample reacted at 650 °C for 8 h with current
scanning electron microscopy (SEM). The compo- density of 1527 A cm⫺2.
4490 J.E. Garay et al. / Acta Materialia 51 (2003) 4487–4495

A cm⫺2 at 650 °C. In contrast to the case where


the growth took place in the absence of a current,
Fig. 4, here no induction time in layer growth is
apparent. A more significant difference, however,
is the marked increase in growth rate. In the
absence of a current and at a higher temperature
(700 °C) no product was observed after 5 h of
annealing, within the limits of resolution of our
measurements (0.1 µm). In contrast, at 650 °C, the
total product layer was about 9.5 µm thick after 5
h of annealing. The dramatic effect of the current
is shown more explicitly in Fig. 6. This shows the
Fig. 4. Product layer thickness squared (⌬x)2 vs. time for effect of the current density on the thickness of the
samples reacted at 700 °C with no current (0 A cm⫺2). NiTi layer resulting from reacting the system for
4 h at 650 °C. As the current density is increased
previous work by Bastin and Rieck [12], no incu- from 1018 to 2546 A cm⫺2 the thickness of the
bation period was observed and the growth of the NiTi layer increased by nearly a factor of seven
Ni3Ti phase was not parabolic as was the case for under the same conditions. Qualitatively similar
the other two phases. With regards to the incu- results were obtained for the other intermetallic
bation period, the difference between our work and phases at this and other temperatures. However,
that reported by Bastin and Rieck [12] may be the although the current had an effect on the growth
consequence of the sample preparation method of the layers in all conditions, the effect was more
employed by these authors. They prepared their pronounced at lower temperatures.
diffusion couples by initially welding the reacting In the work of Bastin and Rieck [12,18],
layers, a circumstance that could produce some incremental product formation was investigated by
degree of intermixing of the reactants at the inter- using two compounds as reactants to obtain kinetic
face prior to the initiation of the diffusion experi- data on the formation of an intermediate phase.
ments. Although we used Ti and Ni as starting reactants,
With the application of a current, significant our results can provide similar kinetic information
changes in the reactivity between Ti and Ni were using the recent analysis of Buscaglia and
observed. Fig. 5 shows the growth behavior of the Anselmi-Tamburini [19] based on the derivations
layers in the presence of a current density of 2036 of Wagner [20]. This analysis is based on the

Fig. 5. Product layer thickness squared (⌬x)2 vs. time for Fig. 6. NiTi layer thickness (⌬x) vs. current density for
samples reacted at 650 °C with current density of 2036 A cm⫺2. samples reacted at 650 °C for 4 h.
J.E. Garay et al. / Acta Materialia 51 (2003) 4487–4495 4491

assumptions that: the growth of all phases is para- where Ci = √kIi, Vi the molar volumes, and ni the
bolic, local equilibrium is established at each inter- stoichiometric coefficients of each phase nor-
face, the phases have narrow homogeneity ranges, malized to 1 mol of component B, AniB. Values of
and the extent of solubility in the terminal phases, kiII represent the intrinsic property of each phase,
i.e., Ti and Ni, in this case, is limited. These depending only on the interdiffusion coefficient
assumptions can be reasonably applied to the Ni– and the free energy of formation of that phase.
Ti system in the temperature range of this investi- The intrinsic rate constants obtained through this
gation. The homogeneity range of the phases NiTi method for all the experiments performed at tem-
and NiTi2 below 750 °C is relatively small and the peratures (650–750 °C) below the α→β transition
fairly large solubility of Ti in Ni can be ignored temperature for Ti are reported together with the
since the diffusivity of Ti in Ni is extremely slow first kind constants in Table 1. It should be noted
even at fairly high temperatures [18]. that no data are reported in this table for the highest
Using the nomenclature of Wagner, we calculate current density, 2456 A cm⫺2, at 750 °C. Under
the rate constant of the first kind, kI, from the these conditions, the growth rate is so high that the
slopes of lines such as those shown in Fig. 7 which middle, Ni, layer of the starting material will be
is for the growth of the NiTi layer at 750 °C under completely consumed after a relatively short time.
different current densities. Thus When this occurs, the equilibrium conditions will
be different. Table 2 reports the first kind rate con-
(⌬x)2 ⫽ kIi t (1)
stants for experiments performed at temperatures
where ⌬x is the thickness of the layer of the inter- (800–850 °C) above the α→β transition tempera-
metallic and t is time. When only one phase forms, ture for Ti. Intrinsic (second kind) rate constants
then the constant is the true rate constant and is cannot be calculated for the higher temperatures,
referred to as the rate constant of the second kind, since above 765 °C, the extensive solubility of Ni
kiII. The relationship between kiI and kiII has been in Ti to form β-Ti violates the assumption in the
provided by Buscaglia and Anselmi-Tamburini as derivation of this kinetic parameter, as indicated

冋冘
above.
(ni⫺1⫺ni+1)Vi
i⫺1
Cj ni
kIIi ⫽ C2 n ⫹ (2)
(ni⫺1⫺ni)(ni⫺ni+1) i ij ⫽ 1VjCi Vi

冘 册
n
4. Discussion
njCj
⫹ i ⫽ 1,…,n
V C From the calculations described above, we are
j ⫽ i⫹1 j i
able to compare our results with those of Bastin
and Rieck [12,18]. The rate constant, kII, determ-
ined by these authors for the growth of the NiTi
phase at 700 °C (grown from the couple Ni3Ti and
NiTi2) is reported as 1.5 × 10⫺11 cm·s⫺1. Under
the same conditions (i.e., at 700 °C and without
using a current) but starting with the couple Ni and
Ti, we calculate a value for the intrinsic constant
of 2.43 × 10⫺11 cm·s⫺1. The agreement between
the two values is reasonable.
However, the imposition of a current during the
reaction shows a significant change in this con-
stant. Table 3 shows the ratios of the kII values
calculated in this work under various current den-
sities to those calculated by Bastin and Rieck (with
Fig. 7. NiTi layer thickness squared (⌬x)2 vs. time for experi- no current) at the temperatures indicated for the
ments reacted at 650 °C for different current densities. phases NiTi and NiTi2. No comparison is made for
4492 J.E. Garay et al. / Acta Materialia 51 (2003) 4487–4495

Table 1
Calculated growth rate constants for intermetallic phases in the Ni–Ti system. T below α→β transition of Ti

T (°C) Current density (A cm⫺2) Phase kI × 1012 (cm2 s⫺1) kII × 1012 (cm2 s⫺1)

650 1527 NiTi 1.89 8.02


NiTi2 0.25 2.21
Ni3Ti 1.32 3.64
2036 NiTi 6.98 34.0
NiTi2 0.62 7.48
Ni3Ti 12.90 28.90
2456 NiTi 27.80 117.00
NiTi2 2.60 26.70
Ni3Ti 21.10 56.40
700 1527 NiTi 9.02 38.80
NiTi2 0.63 7.52
Ni3Ti 8.59 21.70
2036 NiTi 14.50 62.00
NiTi2 1.28 13.80
Ni3Ti 14.00 36.40
2456 NiTi 58.10 208.00
NiTi2 4.61 44.90
Ni3Ti 21.42 58.00
750 1527 NiTi 23.00 90.90
NiTi2 1.75 18.60
Ni3Ti 12.00 34.80
2036 NiTi 34.30 140.00
NiTi2 3.42 33.10
Ni3Ti 19.90 56.70

Table 2
Table 3
Calculated growth rate constants for intermetallic phases in the
Ratios of intrinsic rate constant of this work (kII) with current
system Ni–Ti. T above α→β transition of Ti; current density
to values calculated from Bastin and Rieck, kIIBR, without cur-
= 1527 A cm⫺2
rent [12]

T (°C) Phase kI × 1012 (cm2 s⫺1)


Phase T (°C) Current density kII/kIIBR
(A cm⫺2)
800 NiTi 34.40
NiTi2 4.36
NiTi 650 1527 2.7
Ni3Ti 19.50
2036 11.3
850 NiTi 121.00
2546 39.0
NiTi2 21.70
700 1527 2.6
Ni3Ti 23.50
2036 4.1
2546 14.2
750 1527 1.8
the Ni3Ti since in the work of Bastin and Rieck, 2036 2.8
NiTi2 650 1527 3.6
the growth of this phase was not parabolic. Table 2036 12.2
1 shows a considerable increase in the rate constant 2546 43.7
as a result of the application of a current. The 700 1527 2.2
increase depends on the magnitude of the current 2036 4.0
density in all cases, but is most pronounced at the 2546 13.2
750 1527 2.1
lowest temperature, 650 °C. At this temperature, 2036 3.7
the ratio of the rate constant for NiTi increases
J.E. Garay et al. / Acta Materialia 51 (2003) 4487–4495 4493

from 2.675 to 39.028 as the current density is


increased from 1527 to 2546 A cm⫺2. An approxi-
mately similar increase was observed for the NiTi2
phase at this temperature. Due to the high reac-
tivity at higher temperatures, no data were obtained
for the highest current density, 2546 A cm⫺2 for
both intermetallics. The effect of current density
on the rate constant, kII, for all three intermetallic
phases at 650 °C is shown in Fig. 8.
The observed effect of the current on the rate of
growth is now examined in terms of its effect on
the activation energy. Fig. 9(a)–(c) is Arrhenius
plots of the intrinsic constants for the intermetallic
phases NiTi, Ni3Ti, and NiTi2, respectively. In
each case plots for different current densities are
shown. For the cases of NiTi and NiTi2, the cur-
rent-free data of Bastin and Rieck [12] are also
plotted. The figures show a significant reduction in
the activation energy with an increase in current
density for all phases, although the dependence on
current density appears to be different for each
phase, with the largest decrease being observed in
the Ni3Ti case. The calculated values of the acti-
vation energy for different current densities and
temperature ranges are also reported in Table 4. It
should be noted that in the investigation on the
Au–Al system, the current had no apparent effect
on the activation energy.
As indicated above, the degree of influence of
the current on the activation energy depends on the
phase produced. Examining the ratio of the acti-
vation energy at the lowest and intermediate cur-

Fig. 9. (a) Temperature dependence of NiTi intrinsic growth


rate constant for samples reacted with current densities of 1527
and 2036 A cm⫺2. The results of Bastin and Rieck [12] with
no applied current (0 A cm⫺2) are plotted for comparison. (b)
Temperature dependence of Ni3Ti intrinsic growth rate constant
for samples reacted with current densities of 1527 and 2036 A
cm⫺2. (c) Temperature dependence of NiTi2 intrinsic growth
rate constant for samples reacted with current densities of 1527
and 2036 A cm⫺2. The results of Bastin and Rieck [12] with
no applied current (0 A cm⫺2) are plotted for comparison.
Fig. 8. Intrinsic growth rate constants (kII) vs. current density
for sample reacted at 650 °C.
4494 J.E. Garay et al. / Acta Materialia 51 (2003) 4487–4495

Table 4 study, the imposition of a current reduced the incu-


Effect of current density on activation energy for phase growth bation time to zero (compare Fig. 4 with Fig. 5).
in the system Ni–Ti. values for zero current density from Bastin
and Rieck [12] However, since the growth of the layers is consist-
ent with bulk diffusion, the effect of the current on
Phase Current Activation energy (kJ mol⫺1) nucleation is likely governed by diffusion also as
density has been pointed out by Thompson [21].
(A cm⫺2) Thus in the absence of a clear electromigration
effect, the role of the current in enhancing dif-
NiTi 0 363.1 [12] fusion may relate to changes in defect concen-
1527 313.9
2036 181.4 tration or their mobility. In a recent positron
NiTi2 0 346.4 [12] annihilation spectroscopy investigation by Asoka-
1527 275.1 Kumar et al. [22], it was found that the imposition
2036 190.9 of a current increased the concentration of vacanc-
Ni3Ti 0 n/a ies in metals. The observed enhancement of the
1527 292.7
2036 86.4 growth of the intermetallics by the current may,
therefore, relate to an increase in the defect con-
centration in these phases or on their mobility. No
rent densities (with those with no current) provides information on the effect of the current on defect
the values 0.30, 0.58, and 0.69 for the phases mobility is at hand, but on-going investigations are
Ni3Ti, NiTi, and NiTi2, respectively. The obvious aimed at determining this possible effect.
correlation is with the nickel content of the phase: Another important aspect of the diffusion study
the current being more effective in the formation in the Ni–Ti system is the role of the Ti phase
of the phase with the highest Ni content. In view transformation on the observed reactivity. Bastin
of the fact that the diffusivity of Ni in Ti is higher and Rieck [12] reported that the transition of Ti
than that of Ti in Ni [18], we would propose that from α to β produces a considerable decrease in
the current has a greater effect on the diffusion of intermetallic layer growth rate due to the higher
Ni in the intermetallic. However, the absence of solubility of Ni in β-Ti. The transition between the
diffusion data of Ni in these intermetallics makes two forms occurs at 765 °C, a temperature corre-
it impossible to ascertain the validity of this pro- sponding to the Ti-rich eutectoid shown in Fig. 1.
posal. Fig. 10 shows the temperature dependence of the
As indicated above with reference to Fig. 3, the
growth of all layers was symmetric with respect to
the direction of the DC current. This observation
is taken as an indication that electromigration is
not the dominant effect of the current. Similar
results were reported for the Au–Al system
[10,11]. In addition to electromigration (electron
wind effect), the current can possibly play other
roles and other explanations have also been offered
for the observation of current direction indepen-
dence. It was proposed in one study that the current
enhances the nucleation rate of the new phases [8].
In our previous work, we observed a significant
effect on nucleation time in the system Au–Al [11]
in which layer growth was also independent of cur- Fig. 10. Temperature dependence of NiTi growth rate con-
stant showing effect of β-Ti phase transformation, for samples
rent direction. In the previous study, however, a reacted with a current density of 1527 A cm⫺2. The results of
nucleation time was observed even under the high- Bastin and Rieck [12] with no applied current (0 A cm⫺2) are
est current density employed. In contrast, in this plotted for comparison.
J.E. Garay et al. / Acta Materialia 51 (2003) 4487–4495 4495

experimental growth constant for the NiTi phase measurements and positron annihilation
from the results of Bastin and Rieck and from spectroscopy.
those obtained in this study when a current density
of 1527 A cm⫺2 was employed. In both cases, a
discontinuity is observed in a temperature range Acknowledgments
very close to the transformation temperature (1 / T
= 9.63 × 10⫺4). In addition to confirming the effect The financial support of this work by the
of phase transformation on the kinetics, this obser- National Science Foundation is gratefully
vation is also a validation of the temperature acknowledged.
measurements in our system.

References
5. Conclusion
[1] Munir ZA, Anselmi-Tamburini U. Mater Sci Rep
The imposition of a DC current was found to 1989;3:277.
increase the reactivity of the Ni–Ti system. In iso- [2] Munir ZA. Zeit Physik Chem 1998;207:39.
thermal diffusion couple experiments, the current [3] Orru R, Cao G, Munir ZA. Metall Mater Trans
pronouncedly increased the growth rate and 1999;30A:1101.
[4] Kawase K, Munir ZA. Int J SHS 1998;7:95.
decreased the effective activation energy for [5] Feng A, Munir ZA. Metall Trans 1995;26B:581.
growth of all three intermetallic phases (Ni3Ti, [6] Feng A, Munir ZA. J Am Ceram Soc 1997;80:1222.
NiTi and NiTi2). The effect increases with an [7] Feng A, Graeve OA, Munir ZA. Comput Mater Sci
increase in the magnitude of the current density in 1998;12:137.
all cases at all temperatures in the study, but is [8] Chen CM, Chen SW. J Electron Mater 2000;29:1222.
[9] Chen SW, Chen CM, Liu WC. J Electron Mater
most marked at the lowest temperature, 650 °C. 1998;27:1193.
The effect on growth rate was strongest for the [10] Bertolino N, Garay J, Anselmi-Tamburini U, Munir ZA.
NiTi2 compound, whose intrinsic growth rate con- Scripta Mater 2001;44:737.
stant was 43 times higher under the influence of [11] Bertolino N, Garay J, Anselmi-Tamburini U, Munir ZA.
2546 A cm⫺2 compared with no current diffusion Phil Mag B 2002;82:969.
[12] Bastin GF, Rieck GD. Metall Trans 1974;5:1817.
couple experiments at the same temperature of 650 [13] Pasturel A, Colinet C, Nguyen Manh D, Paxton AT, van
°C. The effect was nearly as strong (39 times Schilfgaarde M. Phys Rev B 1995;52:15176.
higher) for the NiTi phase. By contrast, the applied [14] Komai N, Watanabe W, Horita Z, Sano T, Nemoto M.
current decreased the effective activation energy Acta Mater 1998;46:4443.
most dramatically in the Ni3Ti case falling from [15] Taylor A, Floyd W. Acta Crystallogr 1950;3:285.
[16] Yurkow GA, Barton JW, Parr JG. Acta Crystallogr
292 kJ mol⫺1 under the influence of 1527 A cm⫺2 1959;12:909.
to 86 kJ mol⫺1 for 2036 A cm⫺2. It is suggested [17] Fukuda T, Kakeshita T. Mater Sci Eng A 1999;166:273.
that the application of the current increases the [18] Bastin GF, Rieck GD. Metall Trans 1974;5:1827.
mobility of defects or increases their concentration [19] Buscaglia V, Anselmi-Tamburini U. Acta Mater
over the thermal equilibrium value, thereby 2002;50:525.
[20] Wagner C. Acta Metall 1969;17:99.
increasing the product growth rate. Studies are [21] Thompson CV. J Mater Res 1992;7:367.
underway to determine the validity of these [22] Asoka-Kumar P, O’brien K, Lynn KG, Simpson PJ,
proposed explanations using electrical resistivity Rodbell KP. Appl Phys Lett 1996;68:406.

S-ar putea să vă placă și