Вы находитесь на странице: 1из 7

Journal of Alloys and Compounds 477 (2009) 870876

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


journal homepage: www.elsevier.com/locate/jallcom

Discontinuous and continuous precipitation in


magnesiumaluminium type alloys

K.N. Braszczynska-Malik
Czestochowa University of Technology, Institute of Materials Engineering, Al. Armii Krajowej 19,
42-200 Czestochowa, Poland

a r t i c l e

i n f o

Article history:
Received 9 September 2008
Received in revised form 23 October 2008
Accepted 1 November 2008
Available online 20 December 2008
Keywords:
Magnesium alloy
Precipitation
Microstructure

a b s t r a c t
The microstructure investigations of the AZ91 alloy and binary Mg9 wt.% Al alloy after different heat
treatments were presented. Solution annealing at 693 K for 26 h (with water quenching), followed by
ageing at 423, 473, 543 and 623 K was carried out. After ageing at 423 K only discontinuous precipitates were observed whereas at 623 K only continuous ones were revealed in the microstructure of both
alloys. At intermediate ageing temperatures (473 and 543 K) both discontinuous and continuous precipitates occurred competitively. Additional analyses revealed that after cooling a solid solution from 693 K
(without quenching) only discontinuous precipitates were formed. On the other hand, after heating a
supersaturated solid solution from room temperature to 665 K both discontinuous and continuous precipitates were observed simultaneously. On the basis of the obtained results a model of precipitate types
dependent on heat treatments for Mg9 wt.% Al alloys was proposed.
2008 Elsevier B.V. All rights reserved.

1. Introduction
In lightweight magnesium alloys, aluminium constitutes the
main alloying element, chiey because of its low price, availability, low density and the advantageous effect on corrosion and
strength properties [15]. The microstructure of as-cast MgAl
alloys is generally characterized by: a solid solution of aluminium in
magnesium (an -Mg phase with a hexagonal close-packed structure) and an + eutectic (fully or partially divorced depending
on its solidication rate) [510]. The -phase (called also -phase
[1115]) is an intermetallic compound with a stoichiometric composition of Mg17 Al12 (at 43.95 wt.% Al) and an -Mn type cubic
unit cell. In comparison with binary MgAl alloys, new phases
do not appear in commercial ternary alloys with zinc (like AZ91)
when the Al to Zn ratio is larger than 3:1 [1314]. In this case,
zinc substitutes aluminium in the -Mg17 Al12 phase, creating a
ternary intermetallic compound Mg17 Al11.5 Zn0.5 or Mg17 (Al,Zn)12
type [4,15]. On the other hand, the presence of a small amount of
manganese in commercial magnesiumaluminium alloys additionally causes the formation of aluminiummanganese intermetallic
compounds Al8 Mn5 or Al11 Mn4 [16].
Magnesiumaluminium alloys are susceptible to heat treatment due to the variable solubility of their alloying elements in

Tel.: +48 34 3250 721; fax: +48 34 3250 721.


E-mail address: kacha@wip.pcz.pl.
0925-8388/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.jallcom.2008.11.008

a solid state with temperature [5]. The maximum solid solubility of aluminium in magnesium is reasonably high at 12.9 wt.%
Al at an eutectic temperature of 710 K whereas the equilibrium
concentration at 473 K is about 2.9 wt.% Al. During conventional
heat treatment, involving solution annealing at about 690 K for a
minimum of 24 h, followed by ageing at about 430 K for 16 h (T6
conditions), the precipitation process appears and the formation
of -phase precipitates occurs. During -phase precipitation, neither Gunier-Preston zones nor other metastable phases are formed.
However, precipitation occurring in supersaturated alloys can take
place either continuously or discontinuously. Discontinuous precipitation (DP) is the cellular growth of alternating plates of the
secondary phase and near-equilibrium matrix phase at high angle
boundaries. This heterogeneous reaction leads to the formation of
a lamellar structure behind a moving grain boundary. Continuous
precipitation (CP) forms in all the remaining regions of the supersaturated matrix. In most alloys, nevertheless, these two types of
precipitates can also occur simultaneously [5,17].
Recently, investigations concerning precipitation processes in
MgAl alloys have been reported [1332]. In most of them, however, discontinuous and continuous precipitation were analysed
separately, whereas these two types of precipitates can occur
simultaneously and compete in an intricate manner because they
nucleate and grow at different rates and with different mechanisms.
Duly et al. [17] proposed a possibility to superimpose a morphology map on the phase diagram but it has not found conrmation in
different experiments. Bradai et al. [25] reported that only discon-

K.N. Braszczy
nska-Malik / Journal of Alloys and Compounds 477 (2009) 870876

tinuous precipitates were formed after ageing at 498 K for Mg7%


to 11% Al alloys, whereas according to Dulys results, at this temperature there should also be continuous precipitation. Continuous
precipitates were revealed in the Mg6% Al alloy after ageing at
475 K [15], while to be agreement with ref. [17] there should only
be discontinuous precipitation. Similarly, only discontinuous precipitates for Mg10% Al after ageing at 495500 K were disclosed
[2527,29]. For the AZ91 alloy only continuously precipitates after
ageing at 373573 K were analysed in refs. [13,30], whereas only
discontinuous precipitates after ageing at 457 K were revealed in
refs. [19,15]. These results are not only in agreement with the Duly
model but also inverse.
It should also be noted, that both discontinuous and continuous precipitates in MgAl type alloys have a plate-like morphology
with an accurate orientation relationship (OR) with a matrix phase.
For both precipitates the predominant orientation relationship is
and [21 10]
|| [1 1 1]
the Burgers OR, namely: (0 0 0 1) || (010)
[13,21,22,30,31]. Additionally, other ORs were also reported in
MgAl based alloys, i.e. the Porter OR [13], the Gjmmes-strmoe
OR [30,31], the Crawley OR [13,30,31] or the Potter OR [30,32].
In the present work, the inuence of the ageing temperature and cooling/heating rate on the precipitation mechanism in
Mg9 wt% Al and AZ91 were analysed. Investigations of the individual microstructure development were realized using scanning
electron microscopy (SEM).

871

signicance of zinc addition on the formation of precipitate types some results


obtained from the experimental binary Mg9 wt.% Al alloy were also presented.
The analysed alloys were permanent moulds, producing rod samples of 40 mm
in diameter. Solution annealing was carried out at 693 K for 26 h in a protective
argon atmosphere for all the samples. The next steps of heat treatment included:
(i) water quenching (at approximately 287 K) and ageing at different temperatures:
423, 473, 543 and 623 K, (ii) water quenching and heating from room temperature to 665 K at a heating rate of 2 K/min., (iii) cooling down just after solution
annealing (without quenching) from 693 K to room temperature at a cooling rate of
5 K/min.
In order to determine the microstructure, light and scanning electron
microscopy (SEM) techniques were used in this study. A standard metallographic technique was used for sample preparation including wet prepolishing
and polishing with different diamond pastes without contact with water. To
reveal the microstructure, the samples were etched in a 1% solution of HNO3
in C2 H5 OH for about 60 s. Scanning electron microscopes, a JSM-5400 (Jeol,
Tokyo, Japan) and an XL30ESEM FEG (Philips, Eindhoven, The Netherlands) were
used.

2. Experimental procedures
The commercial AZ91 magnesium alloy with a nominal chemical composition
of 9 wt.% Al, 1 wt.% Zn, 0.5 wt.% Mn was used in this study. In order to determine the

Fig. 1. Microstructure of AZ91 alloy: (a) as-cast, (b) after solution annealing at 693 K
for 24 h; light microscopy.

Fig. 2. Discontinuous precipitates in (a) AZ91 and (b and c) Mg9%Al alloy after
ageing supersaturated solid solution at 423 K for 16 h.; SEM.

872

K.N. Braszczy
nska-Malik / Journal of Alloys and Compounds 477 (2009) 870876

3. Results
Fig. 1(a) shows a typical as-cast microstructure of the AZ91
alloy. Dendritic structure is typical for cast magnesium alloys,
which are characterized by very heavy segregation of alloying elements. MgAl alloys are prone to segregation due to relatively wide
temperature spans between the liquids and the solids curves. Nonequilibrium solidication conditions caused the formation of large
crystals of the -Mg phase (depleted in aluminium) and pushing
the Al admixture away into interdendrical spaces. The microstructure consisted of primary -Mg dendrites (impoverished in Al) and
the binary partially divorced eutectic + . In the AZ91 alloy an
Al8 Mn5 intermetallic phase is also present due to the manganese
addition to the commercial alloy.
Solution annealing of the AZ91 alloy at 693 K caused a total
dissolution of the + eutectic and homogenised the aluminium
throughout the matrix. The time necessary to obtain the homogeneous microstructure consisting of solid solution grains is very long
(minimum 24 h) due to very slow diffusion of aluminium in a magnesium solid state. It should be noted, that the Al8 Mn5 intermetallic
compound is not involved in heat treatment. The microstructure
obtained after solution annealing is shown in Fig. 1(b). The same
microstructure (without the Al8 Mn5 phase) was also observed for
binary MgAl alloys which was reported earlier [5,28].
The microstructure of the AZ91 and binary Mg9 wt.% Al alloys
after ageing of the supersaturated solid solution at 423 K for 16 h
(T6 conditions) was characterized by the presence of ne, plate-like,
discontinuous precipitates. Fig. 2 shows a typical microstructure
of the AZ91 and Mg9 wt.% Al alloys after this heat treatment. For

Fig. 4. Microstructure of AZ91 alloy after ageing at 543 K (a) for 1 h, discontinuous
and continuous precipitates, (b) for 2 h, continuous precipitates; SEM.

Fig. 3. Microstructure of AZ91 alloy after ageing at 473 K (a) for 1 h, discontinuous
precipitates, (b) for 2 h, continuous precipitates; SEM.

both alloys the -phase precipitates had a lamellar morphology


with a marked anisotropy of growth.
Figs. 36 show typical SEM images representing the variation
in precipitation morphology with the temperature. At the temperature of 473 K both continuous and discontinuous precipitates
were observed. After ageing for 1 h, only discontinuous precipitation of the -phase was revealed (Fig. 3(a)) but heat treatment for
2 h at 473 K caused the formation of distinct continuous precipitates (Fig. 3(b)). Similarly, after ageing at 543 K the same results
were obtained. Fig. 4 shows the precipitates morphology observed
after ageing for 1 and 2 h at 543 K. The light areas in Fig. 4(a)
represent typical colonies of discontinuous lamellar precipitates
growing from the grain boundaries. Additionally, after ageing at
this temperature, small continuous precipitates inside the grains
were observed already after 1 h. Fig. 4(b) illustrates ne continuous
precipitates observed after ageing for 2 h at 543 K.
A different situation was observed after ageing supersaturated
alloys at 623 K. Although in the samples aged for 1 h the presence of
precipitates at the grain boundaries was also revealed (Fig. 5(a)),
the process of its growth was quickly stopped. Ageing of the AZ91
alloy at 623 K for 2 h caused the occurrence only of continuous
precipitates also with a visible orientation relationship with the
matrix grains (Fig. 5(b)). It should be noted, that at this temperature (623 K) typical colonies of discontinuous precipitates did not
occur. The same results were observed for the binary Mg9 wt.% Al
alloy. Fig. 6 shows continuous precipitates present in the AZ91 and
Mg9 wt.% Al alloy after ageing at 623 K for 8 h.
After cooling a solid solution from 693 K to room temperature,
without quenching, only discontinuous precipitates were observed
in all of the samples. In this case continuous precipitates did not
appear. Fig. 7(a) presents colonies of discontinuous precipitates cre-

K.N. Braszczy
nska-Malik / Journal of Alloys and Compounds 477 (2009) 870876

Fig. 5. Microstructure of AZ91 alloy after ageing at 623 K (a) for 1 h precipitates at
grain boundary and continuous precipitates inside grains, (b) for 2 h, continuous
precipitates; SEM.

873

Fig. 7. Microstructure of AZ91 alloy (a) discontinuous precipitates after cooling from
693 K to room temperature with a rate of 5 K/min. (without quenching), (b) discontinuous and continuous precipitates after heating supersaturated solid solution to
665 K with a rate of 2 K/min; SEM.

ated during the cooling down of a solid solution from 693 K to room
temperature. Different results were obtained after heating a supersaturated solid solution from room temperature to 665 K with a
heating rate of 2 K/min. In this case, both discontinuous and continuous precipitates of the -phase were observed. Fig. 7(b) shows
the formation of discontinuous precipitates at the grain boundaries
and continuous ones in the grains.
4. Discussion

Fig. 6. Continuous precipitates in (a) AZ91 and (b) Mg9% Al alloy after ageing
supersaturated solid solution at 623 K for 8 h; SEM.

During both discontinuous and continuous precipitation


reactions, representing a solidsolid-phase transformation, a
supersaturated solid solution (0 ) decomposes into a new soluterich precipitate () and a less-saturated, near-equilibrium, initial
phase () with the same crystal structure as the 0 . The differences between discontinuous and continuous precipitation consist
in the nucleation places and growth. During discontinuous reaction, alternating layers of and phases grow behind a moving
grain boundary. Continuous precipitation proceeds by a different
mechanism, where precipitates of the -phase nucleate and grow
inside the 0 grains.
As was revealed, for the AZ91 and Mg9 wt.% Al alloys, discontinuous and continuous precipitates can occur simultaneously or
competitively, dependently on the ageing temperature. After ageing at 423 K typical colonies of discontinuous precipitates were
observed (Fig. 2) whereas at 623 K only continuous precipitates
were formed inside grains (Figs. 5 and 6). It proved that continuous
precipitation tends to be favoured at high temperatures (i.e. close

874

K.N. Braszczy
nska-Malik / Journal of Alloys and Compounds 477 (2009) 870876

Fig. 8. Schemes of -phase precipitation possibilities in MgAl alloys depending on heat treatment; (a) during ageing of supersaturated (and quenching in water) solid
solution, (b) during cooling of solid solution from solution annealing temperature (without quenching), (c) during heating of supersaturated solid solution (after quenching)
from room temperature.

to the solvus curve) whereas at low temperatures of ageing, discontinuous precipitation invades all the samples. These results are in
agreement with the ndings of Duly et al. [17]. The presented observations also indicate that at an intermediate temperature range
(473 or 543 K) both discontinuous and continuous precipitates can

be observed (Figs. 3 and 4). These results are in contradiction to


those described in ref. [17], where the existence of solely discontinuous precipitates in the range of middle ageing temperatures
was claimed. Fig. 8(a) shows all the revealing possibilities for precipitation from a supersaturated solid solution (after quenching

K.N. Braszczy
nska-Malik / Journal of Alloys and Compounds 477 (2009) 870876

in water) in MgAl alloys depending on different ageing temperatures. The same results were obtained for the commercial AZ91
alloy and experimentally binary Mg9 wt.% Al alloy. It conrmed,
that zinc present in the commercial alloy does not inuence the
precipitate type. It should also be noted, that the volume fraction of
continuous precipitates increased with ageing temperature. Additionally, as could be expected, the time necessary for nucleation
and growth of precipitates decreases with ageing temperature.
As well known, the volume fraction of regions transformed by
discontinuous precipitation essentially depend on reaction front
velocity, interlamellar spacing, average composition of the solutedepleted lamellae, temperature, solute content and conditions
of the grain boundaries [25,26,29,3337]. The phase transformation during continuous precipitation is mainly described as the
sequential or simultaneous processes of nucleation, growth and
impingement [13,15,17]. In both cases, reduction of the driving force
for propagation and impediment of nucleation may be the limiting
factors. However, a simple dependence describing the dominant
type of precipitation did not seem to exist. It should also be noted,
that at all temperatures, grain boundaries tend to be rapidly decorated by arrays of heterogeneous precipitates testifying that the
grain boundaries are the rst and privileged places of nucleation
for precipitates.
Additionally, the amount of continuous precipitates very
strongly depends on the number of crystal defects within the matrix
which may act as heterogeneous nucleation sides. It is well known,
that besides the grain boundaries, the various nucleation sizes are:
vacancies, dislocations, stacking faults and solid free surfaces. One
possibility is the heterogeneous nucleation of continuous precipitates on vacancies. On the other hand, vacancies also determine the
volume diffusion mechanism. Due to comparable Al and Mg atom
diameters, vacancy diffusion can be dominant in the investigated
alloys. Solution annealing and quenching result in a higher (nonequilibrium) vacancy concentration. If a concentration of vacancy
is lower (i.e. near-equilibrium) continuous precipitates of the phase
cannot occur. Although, homogeneous nucleation of continuous
precipitates of the -phase was concluded by Celotto [13], the
obtained results did not conrm this hypothesis. As was revealed,
during slow cooling of the samples just after solution annealing
(without quenching) continuous precipitation did not proceed, in
spite of the high temperature, because of the deciency of nucleation sides for these precipitates inside the grains (Fig. 7(b). On the
other hand, during heating from room temperature of a supersaturated alloy, both types of precipitates occur competitively. At the
beginning of the process, discontinuous precipitates started on the
grain boundaries, but they were stopped when the temperature was
allowed to dominate the volume diffusion and continuous precipitates could start (Fig. 7(b). These cases of precipitate formation are
presented schematically in Fig. 8(b) and (c).
As expected for a higher temperature the diffusion became
faster, resulting in an increase of volume precipitates. The precipitates also became larger and more massive. Nevertheless, the
mutual ratio of the discontinuous and continuous precipitates in
each sample is practically not evident, because it depends on the
local conditions for nucleation and growth in the grains and at grain
boundaries. Moreou et al. [38] have found that the volume diffusion
coefcient (DV ) for aluminium in magnesium can be determined
using:


DV = 12 104 exp

144 103
RT


m2 /s

(1)

Discontinuous precipitation is described by Sundquist, Cahn,


Hillert, Turnbull or Petermann-Hornbogen models [25,3336]
whereas continuous precipitation by the Austin and Ricket (AR)

875

Fig. 9. Volume (DV ) and grain boundary (Dgb ) diffusion coefcient for aluminium in
magnesium based on [19,25,27] data.

equation or the well-known Avrami equation, which is also


known as the JohnsonMehlAvramiKolmogorov (JMAK) equation [13,18]. For consistency, the grain boundary diffusivities for
discontinuous precipitation can be evaluated using the PetermannHornbogen equation [25,37]:
sDgb =

RT
2 
8G

(2)

where Dgb is the grain boundary diffusion coefcient, the grain


boundary width, s the segregation factor,  the velocity for
the process, G the total driving force, R the gas constant, T
the absolute temperature of the process and  is the interlamellar
spacing.
Bradai et al. [25] has determined the Arrhenius parameters as
the pre-exponential factor (sDgb )0 = 1.15 106 m3 /s and the activation energy for grain boundary diffusion: Qgb = 105.3 kJ/mol. A
different value of Qgb = 128 kJ/mol was reported in ref. [19]. On the
other hand, the grain boundary diffusion coefcient, Dgb , can be
calculated from (Eq. (2)) assuming s = 1 and = 0.5 nm, according
to [25]. In Fig. 9 the diffusion coefcients, Dv (from Eq. (1)) and
Dgb (from (Eq. (2)) where the pre-exponential factor (sDgb ) was
evaluated from the Arrhenius plot were presented. The obtained
values show that the volume diffusion coefcient is approximately
67 orders of magnitude less than the diffusion coefcient for
the grain boundary. At lower temperatures the process of secondary precipitation is controlled by grain boundary diffusion.
It causes discontinuous precipitates to form earlier at the grain
boundaries in characteristically shaped colonies of lamellae. As
the temperature increases, bulk diffusion becomes faster, which
tends to favour continuous precipitation. Continuous precipitation
reduces the amount of chemical driving energy available for both
the initiation and propagation of discontinuous precipitates. So if
the continuous precipitation process started inside the grains then
discontinuous precipitates are stopped.
5. Conclusions
1. In Mg9 wt.% Al type alloys only discontinuous precipitates are
obtained after ageing at 423 K or after cooling a solid solution
from 693 K to room temperature.

876

K.N. Braszczy
nska-Malik / Journal of Alloys and Compounds 477 (2009) 870876

2. For AZ91 and binary Mg9 wt.% Al alloys only continuous precipitates are formed after ageing of a supersaturated solid solution
at 623 K.
3. Both discontinuous and continuous precipitates can occur simultaneously after ageing of a supersaturated solid solution at an
intermediate temperature (473, 523 K) and also after heating a
supersaturated solid solution from room temperature.
4. Generally, precipitation in MgAl alloys is determined by two
main factors (i) ageing temperature and (ii) concentration of
structural defects (i.e. vacancies). Discontinuous precipitates are
favoured when the grain boundary diffusion process is dominant
whereas continuous precipitates are formed from a quenched
solid solution when volume diffusion becomes faster.
References
[1] H. Friedrich, S. Schumann, J. Mater. Process. Technol. 117 (2001) 76281.
[2] B.L. Mordike, T. Ebert, Mater. Sci. Eng. A 302 (2001) 3745.
[3] B. Smola, I. Stulikova, F. von Buch, B.L. Mordike, Mater. Sci. Eng. A 324 (2002)
113117.
[4] S. Kleiner, E. Ogris, O. Beffort, P.J. Uggowitzer, Adv. Eng. Mater. 9 (2003) 653658.

[5] K.N. Braszczynska-Malik,


The Study on the Shaping of the Microstructure of

MagnesiumAluminium Alloys, WIPMiFS PCz., Czestochowa,


2005 (in Polish).
[6] C.H. Caceres, C.J. Davidson, J.R. Grifths, C.L. Newton, Mater. Sci. Eng. A 325
(2002) 344355.
[7] A. Lindemann, J. Schmidt, M. Todte, T. Zeuner, Thermochim. Acta 382 (2002)
269275.
[8] G.L. Hao, F.S. Han, Q.Z. Wang, J. Wu, Physica B 391 (2007) 186192.
[9] A. Srinivasan, U.T.S. Pillai, B.C. Pai, Mater. Sci. Eng. A 452453 (2007) 8792.

[10] A.K. Dahle, Y.C. Lee, M.D. Nave, P.L. Schaffer, D.H. StJohn, J. Light Met. 1 (2001)
6172.
[11] R. Gonzlez-Martnez, J. Gken, D. Letzig, K. Steinhoff, K.U. Kainer, J. Alloys
Compd. 437 (2007) 127132.
[12] J. Bursik, M. Svoboda, Microchim. Acta 139 (2002) 3942.
[13] S. Celotto, Acta Mater. 48 (2000) 17751787.
[14] M.A. Gharghouri, G.C. Weatherly, D.J. Embury, Philos. Mag. 78 (1998) 11371149.
[15] S. Celotto, T.J. Bastow, Acta Mater. 49 (2001) 4145.
[16] D. Ohno, R. Mirkovic, R. Schmid-Fetzer, Acta Mater. 54 (2005) 38833891.
[17] D. Duly, J.P. Simon, Y. Brechet, Acta Metall. Mater. 43 (1995) 101106.
[18] J.P. Zhou, D.S. Zhao, R.H. Wang, Z.F. Sun, J.B. Wang, J.N. Gui, O. Zheng, Mater.
Lett. 61 (2007) 47074710.
[19] Q.M. Amir, S.P. Gupta, Can. Metall. Q. 34 (1995) 4350.
[20] T.J. Bastow, S. Celotto, Mater. Sci. Eng. C 23 (2003) 757762.
[21] T.J. Brastow, M.E. Smith, J. Phys. Condens. Matter 7 (1995) 49294937.
[22] E. Cerri, S. Barbagallo, Mater. Lett. 56 (2002) 716720.
[23] J.F. Nie, Scripta Mater. 48 (2003) 10091015.
[24] J.F. Nie, Scripta Mater. 48 (2003) 981984.

[25] D. Bradai, M. Kadi-Hani, P. Zieba,


W.M. Kuschke, W.J. Gust, Mater. Sci. 34 (1999)
331336.
[26] D. Bradai, P. Zieba, E. Bischoff, W. Gust, Mater. Chem. Phys. 72 (2001) 401404.
[27] C.J. Bettles, Mater. Sci. Eng. A 348 (2003) 280288.

[28] K.N. Braszczynska,


Z. Metal. 93 (2002) 845850.

[29] D. Bradai, P. Zieba,


E. Bischoff, W. Gust, Mater. Chem. Phys. 78 (2002) 222226.
[30] M.X. Zhang, P.M. Kelly, Scripta Mater. 48 (2003) 647652.
[31] J.F. Nie, X.L. Xiao, C.P. Luo, B.C. Muddle, Micron 32 (2001) 857863.
[32] D. Duly, M.C. Cheynet, Y. Brechet, Acta Metall. Mater. 42 (1994) 38433847.
[33] E. Rabkin, W. Gust, Y. Estrin, Scripta Mater. 37 (1997) 119124.
[34] L. Klinger, Y. Brechet, D. Duly, Scripta Mater. 37 (1997) 12371242.
[35] L.M. Kingler, Y.M. Brechet, G.P. Prudy, Acta Mater. 45 (1997) 50055013.
[36] S. Matrejean, M. Verona, Y. Brchet, G.R. Prudy, Scripta Mater. 41 (1999)
12351240.

[37] P. Zieba,
Mater. Chem. Phys. 62 (2000) 183313.
[38] G. Moreau, J. Cornet, D. Calais, J. Nucl. Mater. 38 (1971) 197202.

Вам также может понравиться