You are on page 1of 7

Available online at www.sciencedirect.

com

Journal of Magnesium and Alloys 8 (2020) 414–420


www.elsevier.com/locate/jma

Full Length Article


Creep behavior of AJ62 Magnesium–Aluminum–Strontium alloy
F. Dobeš∗, P. Dymáček
CEITEC IPM, Institute of Physics of Materials CAS, Žižkova 22, 616 62 Brno, Czech Republic
Received 17 April 2019; received in revised form 30 January 2020; accepted 30 March 2020
Available online 6 May 2020

Abstract
Creep tests were conducted in uniaxial compression to evaluate the creep behavior of magnesium–aluminum–strontium alloy at temperatures
from 373 to 673 K. Stress dependencies of the creep rate over the whole interval of temperatures and stresses can be well described
phenomenologically by the Garofalo sine hyperbolic equation modified by the inclusion of a threshold stress. The threshold stress increases
with decreasing temperature. Creep data normalized by a diffusion coefficient and shear modulus clearly reveal the existence of two different
regions. Possible mechanisms by which plastic deformation takes place have been identified in both regions. The critical stress at which
dislocations break away from the cloud of foreign atoms agrees well with the value determined by data normalization. At low stresses, a
value of the stress exponent of n∼ =3 is consistent with the model of deformation that takes place through dislocation glide controlled by
dragging of solute atoms. At high stresses, the multiple regression yields the activation energy which agrees with that for prismatic glide.
© 2020 Published by Elsevier B.V. on behalf of Chongqing University.
This is an open access article under the CC BY-NC-ND license. (http://creativecommons.org/licenses/by-nc-nd/4.0/)
Peer review under responsibility of Chongqing University

Keywords: Dislocations; Diffusion; Mechanical properties; Creep.

1. Introduction likin [13]. The elevated temperature creep of the alloys has
received considerable attention, but the reported experimen-
Magnesium-based alloys have been extensively developed tal data focused mainly on a limited temperature range 423–
for use as structural materials in the automotive, aircraft and 473 K [10,11,14–16] with sporadic excursions to lower (408 K
aerospace industries [1–3]. Automotive powertrain compo- [17], 398 K [18,19]) or higher temperatures (523 K [20]). To
nents (e.g., transmission cases and engine blocks) require ma- understand the creep mechanisms, this work aims to carry
terials that are resistant to the creep deformation that occurs out an experimental investigation using a broader range of
as a result of long-term exposure to loading at elevated tem- temperatures and stresses.
peratures. To meet this requirement, it is necessary to care-
fully select alloy compositions that have acceptable mechan-
ical properties and conform to high-productivity operations 2. Experimental procedure
(i.e. die casting). Prospective creep-resistant alloys are based
on ternary additions of Si, rare earth elements, Ca and Sr The AJ62 magnesium alloy (nominal compositions in
to Mg-Al binary alloys [4–8]. Several creep-resistant alloys wt.%: 6Al–2Sr–balance Mg) used in this study was prepared
based on Sr additions to the Mg-Al system were developed by the squeeze-casting technique. Chemical analysis was per-
by Pekguleryuz et al. at the beginning of new millennium formed using glow discharge optical emission spectroscopy
[9–12]. and energy-dispersive X-ray spectroscopy. Fig. 1 shows an
The available creep investigations of these alloys are de- optical micrograph of the as cast alloy. The primary solidified
scribed in detail in the review paper by Pekguleryuz and Ce- Mg dendrites are surrounded by interdendritic ternary phase.
Its chemical composition corresponds to Mg9 Al3 Sr. The re-
∗ Corresponding author. sults of constant strain-rate testing of the same experimental
E-mail address: dobes@ipm.cz (F. Dobeš). heat can be found in previous paper [21].
https://doi.org/10.1016/j.jma.2020.03.001
2213-9567/© 2020 Published by Elsevier B.V. on behalf of Chongqing University. This is an open access article under the CC BY-NC-ND license.
(http://creativecommons.org/licenses/by-nc-nd/4.0/) Peer review under responsibility of Chongqing University
F. Dobeš and P. Dymáček / Journal of Magnesium and Alloys 8 (2020) 414–420 415

Fig. 1. Light optical microscope image of microstructure of the as-received


alloy AJ62.

Fig. 2. Example of creep curve at 623 K. Note a compressed time axis on


The creep tests were performed in uniaxial compression
the right-hand side of the figure.
on samples with a gauge length of 12 mm and a diameter of
6 mm. The compression testing allows studying creep defor-
mation not affected by the onset of fracture processes. On compressive test. Differences in the creep rates obtained by
the other hand, this mode of testing does not make it pos- means of both techniques were negligible in all cases.
sible to study superplasticity or fracture-related phenomena
[22,23]. The test arrangement, which uses specific compres- 3. Results
sion cage (reversible grips), also does not allow rapid cooling
of the specimen under stress. Effective observation of the mi- The stress dependencies of the creep rate ε˙ at tempera-
crostructure after creep is, therefore, limited. The tests were tures used in the experiments are shown in Fig. 3 in double
performed under a constant load in a protective atmosphere logarithmic coordinates. Analysis of these dependencies in a
of dry purified argon at temperatures ranging from 373 to broad temperature interval showed that they can be well de-
673 K. The test temperature was held constant within ± 1 K scribed by the sine hyperbolic equation [25] modified by the
for each individual test. Changes in specimen length were inclusion (addition) of a threshold stress, σ th ,
measured using a linear variable displacement transducer. The ε˙ = A(sinh [B (σ − σth )] )n (4)
samples were subjected to stepwise loading, where the load
where A, B and n are stress-independent parameters. Optimal
was changed after the steady-state creep rate was established
values of these parameters and of the threshold stress were
for a given load.
found by nonlinear regression with the criterion of the least
Examples of the time dependence of true compressive
squares of relative deviations.
strain [24],
The value of the stress exponent n is close to 3 and the in-
ε = − ln (l/l0 ), (1) teger value n = 3 is used for the computation of the regression
where l and l0 are the instantaneous specimen height and the curves drawn in the Fig. 3. The computed values of param-
initial specimen height, respectively, are given in Fig. 2. The eters A and B and of the threshold stress σ th are given in
terminal values of the true stress [24], Figs. 4, 5 and 6 as a function of temperature (reciprocal tem-
perature, respectively). Parameter A increases with increasing
Fi l
σi = , (2) temperature. The Arrhenius type of dependence can be for-
S0 l0 mally applied for this parameter (cf. Fig. 4)
where Fi is the force applied in the ith step and S0 is the initial  
Q
cross-section area, and of the creep rate (i.e., true compressive A ≈ exp − , (5)
RT
strain rate)
where R is the universal gas constant (R = 8.314 J/mole) and
ε˙i = dεi /dt = −(1/l )∗dl /dt , (3)
T is the absolute temperature in K. The activation energy Q
were evaluated for each step. The tests were conducted until is equal to 41 kJ/mole at low temperatures (T ≤ 523K) and
the strain reached a value of ε = 0.15. The stress dependence 330 kJ/mole at high temperatures (T ≥ 523K). Parameter B
of the creep rate ε˙ at a single temperature was usually ob- ranges from 0.01 to 0.06 MPa and is likely only slightly de-
tained from two or more tests. The results of the described pendent on temperature (Fig. 5). The threshold stress σ th de-
stepwise procedure were verified in several cases by a com- creases with increasing temperature (Fig. 6). To demonstrate
parison with results obtained from a conventional single-stress that its temperature dependence is stronger than that of the
416 F. Dobeš and P. Dymáček / Journal of Magnesium and Alloys 8 (2020) 414–420

Fig. 3. Dependence of the creep rate on applied stress at different temperatures.

Fig. 4. Dependence of parameter A on reciprocal temperature. Fig. 5. Temperature dependence of the parameter B.

shear modulus G, the threshold stress is normalized to the [28]: D = 0.0001 · exp [ − 135000/(RT)] m2 s−1 , G = 16.6 ·
shear modulus. The values of the threshold stress are of the [1 − 0.00053 · (T − 300)] GPa and b = 0.32nm. A simple vi-
same order as those observed previously in a Mg-4Al-1Ca sual inspection of the normalized data clearly reveals the exis-
(AX41) alloy but substantially less than that in the magne- tence of two different regions: the border is at approximately
sium alloys strengthened by Saffil fibres [26,27]. They are 0.0035σ /G. It is convenient to examine these regions sepa-
also close to the value of 30 MPa reported for the AJ62 alloy rately.
at 423 K determined by a different technique [19].
4.1. Low stress region

4. Discussion The creep results at low normalized stresses suggest that


the stress exponent is close to ∼3 and the creep rate can be
In Fig. 7, the creep rates normalized by DGb/(kT), where described by a Dorn type equation
D is the lattice diffusion coefficient, G is the shear mod-  
A1 DGb σ − σth 3
ulus, b is the length of the Burgers vector and k is the ε˙ = AB (σ − σth ) =
3 3
, (6)
Boltzmann constant, are plotted against the applied stresses kT G
σ normalized by the shear modulus G. For normalization, the where A1 is the material constant. The threshold stress σ th
values of material constants for pure magnesium were used can then be estimated using the linear extrapolation technique
F. Dobeš and P. Dymáček / Journal of Magnesium and Alloys 8 (2020) 414–420 417

Fig. 6. Temperature dependence of the threshold stress normalized to the


shear modulus. Fig. 8. Relationship between ε 1/3 and applied stress in double linear coordi-
nates.

Fig. 7. Relationship between normalized minimum creep rate and normalized


Fig. 9. Dependence of slope a (Fig. 8) on reciprocal temperature.
stress.

The dependence of the parameter a on the reciprocal tem-


(e.g. [29]) based on plotting ε˙1/3 versus applied stress perature is given in Fig. 9. Mean value of QC is equal
σ (Fig. 8). The procedure results in a simple linear function, to 123 kJ/mol. Nevertheless, it is evident that the activation
energy decreases with decreasing temperature: from 148 to
ε˙1/3 = a ∗ σ + b, (7) 105 kJ/mol.
The activation enthalpy H of the diffusion coefficient D,
where a is the temperature-dependent parameter and the D = D0 ∗ exp ( − H/RT), which controls the creep process is
threshold stress is given by then
σth = −b/a. (8) 2RT 2 dG
H = QC + + RT . (10)
G dT
The apparent activation energy of creep at constant effec-
tive stress σ − σ th can be estimated as The last two terms on the right-hand side have only neg-
  ligible influence on the value of the activation enthalpy H:
∂ ln ε˙ d ln a they increase the apparent activation energy by 1 to 2 kJ/mol
QC = =3∗ . (9)
∂ (−1/RT ) σ −σth d (−1/RT ) in the studied temperature range.
418 F. Dobeš and P. Dymáček / Journal of Magnesium and Alloys 8 (2020) 414–420

The creep results with a stress exponent of n ≈ 3 at the


lowest stresses are consistent with the model of deformation
that takes place through dislocation glide controlled by drag-
ging of solute atoms. The diffusion coefficient can be taken
as that of chemical interdiffusivity suggested either by Darken
[30] or by Fuentes-Samaniego et al. [31,32]
 
 ∗ ∗
 ∂ ln FA
DD = xA DB + xB DA 1 + , (11)
∂ ln xA
  
DA∗ DB∗ ∂ ln FA
DF−S = 1 + , (12)
xA DA∗ + xB DB∗ ∂ ln xA
where DA∗ and DB∗ are the tracer diffusion coefficients for the
A and B atoms in the AB alloy, respectively; xA and xB are
the atomic fractions of A and B in the alloy, respectively; and
FA is the activity coefficient for the A atoms. The respective
activation enthalpies are [33]
xA DB∗ QB∗ + xB DA∗ QA∗
QD = + Qm , (13)
xA DB∗ + xB DA∗
Fig. 10. Temperature dependence of calculated activation energies according
xA DA∗ QB∗ + xB DB∗ QA∗ to Darken [30] and Fuentes-Samaniego et al. [31,32].
QF−S = + Qm , (14)
xA DA∗ + xB DB∗
where we obtain a normalized breakaway stress σ B /G from 3 to
∂ ln m ∂ ln FA 5 × 10−3 , which is in very good agreement with the above
Qm = R T 2 , m=1+ . suggested division of normalized creep data in Fig. 7. The
∂T ∂ ln xA
temperature dependence of the breakaway stresses is evident
The microprobe analysis indicated negligible amount of from the dashed line added to Fig. 3.
strontium in the matrix (∼0.06 at.%), which agrees with the Above the breakaway stresses, the stress power n increases
previous investigations [19]. Consequently, only the diffusion to 7.7. It is more convenient to describe the creep rate at these
coefficients of Mg and Al were included in the estimations stresses by means of the exponential function
of activation energies QD and QF-S . The same measurement  
revealed aluminum content of ∼2.5 at.% in the matrix. The U0 − A∗ bσ
ε˙ = A2 exp − , (16)
diffusion of aluminum in magnesium can be best approxi- kT
mated as D = 0.000574∗ exp(−152310/(RT)) (m2 /s) [34]. The
where U0 is the Gibbs free enthalpy necessary for overcoming
contribution of the temperature dependence of the thermody-
a short range obstacle without the stress, A∗ is the activation
namic factor was neglected. The results of calculations of the
area and A2 is material constant. The multiple regression of
activation energies of diffusion that control dislocation glide
the high-stress data yields the activation energy U0 equal to
are given in Fig. 10. The energy from the Fuentes-Samaniego
111 kJ/mol. This is in fairly good agreement with the acti-
et al. proposal is weakly dependent on the concentration of
vation energy for prismatic glide (1.2 eV = 115.8 kJ/mol) pre-
aluminum and temperature.
dicted by Caillard and Martin [36] and compatible with the
values for creep in magnesium at intermediate temperatures
4.2. High-stress region
quoted by Couret and Caillard [37]. Additionally, the esti-
mated activation areas (9.5 b2 - 16 b2 ) correspond to those
With increasing effective stress, the solute atoms can no
reported by Caillard and Martin (9 b2 at room temperature).
longer hinder the dislocation motion. At a critical stress, the
dislocation breaks away from the cloud of foreign atoms and
moves in a different mode characterized by a high stress 5. Conclusions
power dependence. The critical stress depends on the con-
centration of solute atoms c and the binding energy between The creep of Mg-Al-Sr alloy was studied in the tempera-
the solute atom and the dislocation, which is proportional to ture range from 373 K to 673 K. The alloy was prepared by
the absolute value of the difference in volume between the the squeeze-casting technique. The following conclusions re-
solute and solvent atoms V [35] garding the creep resistance can be drawn:
 2  
1 1 + ν 2 M G2 cV 2
σB = , (15) • Stress dependences of the creep rate over the whole in-
2π 1−ν 10b3 kT terval of temperatures and stresses can be well described
where M = 3.06 is the Taylor factor and ν = 0.34 is the Pois- phenomenologically by the Garofalo sine hyperbolic equa-
son ratio. Taking V = 8.2 × 10−30 m3 and b = 3.2 × 10−10 m, tion modified by the addition of a threshold stress.
F. Dobeš and P. Dymáček / Journal of Magnesium and Alloys 8 (2020) 414–420 419

• Temperature dependence of the creep rate at lower stresses [12] A. Sadeghi, H. Mortezapour, J. Samei, M. Pekguleryuz, D. Wilkinson,
is characterized by an activation energy that decreases with Anisotropy of mechanical properties and crystallographic texture in hot
decreasing temperature: from 148 to 105 kJ/mol. Analysis rolled AZ31+XSr sheets, J. Magnes. Alloys 7 (2019) 466–473, doi:10.
1016/j.jma.2019.04.005.
of the creep rate at higher stresses yields the activation [13] M. Pekguleryuz, M. Celikin, Creep resistance in magne-
energy equal to 111 kJ/mol. sium alloys, Int. Mater. Rev. 55 (2010) 197–217, doi:10.1179/
• Creep data normalized by the diffusion coefficient and 095066010X12646898728327.
shear modulus clearly reveal the existence of two different [14] B. Jing, S. Yangshan, X. Feng, X. Shan, Q. Jing, T. Weijian, Effect of
extrusion on microstructures, and mechanical and creep properties of
regions.
• At low stresses, the value of the stress exponent of n∼
Mg–Al–Sr and Mg–Al–Sr–Ca alloys, Scr. Mater. 55 (2006) 1163–1166,
=3 is doi:10.1016/j.scriptamat.2006.08.020.
consistent with the model of deformation that takes place [15] B. Jing, S. Yangshan, X. Shan, X. Feng, Z. Tianbai, Microstructure
through dislocation glide controlled by dragging of solute and tensile creep behavior of Mg–4Al based magnesium alloys with
atoms. alkaline-earth elements Sr and Ca additions, Mater. Sci. Eng. A 419
(2006) 181–188, doi:10.1016/j.msea.2005.12.017.
• At high stresses, multiple regression yields the activation
[16] P. Zhao, Q. Wang, C. Zhai, Y. Zhu, Effects of strontium and titanium
energy, which agrees with that for prismatic glide predicted on the microstructure, tensile properties and creep behavior of AM50
by Caillard and Martin. alloys, Mater. Sci. Eng. A 444 (2007) 318–326, doi:10.1016/j.msea.
2006.08.111.
Acknowledgements [17] W. Blum, Y.J. Li, X.H. Zeng, P. Zhang, B. von Großmann, C. Haberling,
Creep deformation mechanisms in high-pressure die-cast magnesium-
aluminum–base alloys, Metall. Mater. Trans. 36A (2005) 1721–1728,
This research has been financially supported by the Min- doi:10.1007/s11661- 005- 0036- 0.
istry of Education, Youth and Sports of the Czech Republic [18] S. Xu, M.A. Gharghouri, M. Sahoo, Tensile-Compressive creep asym-
under the project CEITEC 2020 (LQ1601). metry of recent die cast magnesium alloys, Adv. Eng. Mater. 9 (2007)
807–812, doi:10.1002/adem.200700163.
References [19] M. Kunst, A. Fischersworring-Bunk, G. L’Esperance, P. Plamondon,
U. Glatzel, Microstructure and dislocation analysis after creep defor-
[1] S. You, Y. Huang, K.U. Kainer, N. Hort, Recent research and devel- mation of die-cast Mg–Al–Sr (AJ) alloy, Mater. Sci. Eng. A 510–511
opments on wrought magnesium alloys, J. Magnes. Alloys 5 (2017) (2009) 387–392, doi:10.1016/j.msea.2008.07.078.
239–253, doi:10.1016/j.jma.2017.09.001. [20] J. Kubásek, D. Vojtěch, M. Martínek, Structural characteristics and ele-
[2] V.V. Ramalingam, P. Ramasamy, M.D. Kovukkal, G. Myilsamy, Re- vated temperature mechanical properties of AJ62 Mg alloy, Mater. Char-
search and Development in Magnesium Alloys for Industrial and act. 86 (2013) 270–282, doi:10.1016/j.matchar.2013.09.018.
Biomedical Applications: A Review, Met. Mater. Int. (2019) 1–22, [21] Z. Trojanová, P. Lukáč, Deformation Behaviour of AX91 and AJ62 Mg
doi:10.1007/s12540- 019- 00346- 8. Alloys, Procedia Eng 10 (2011) 2318–2323, doi:10.1016/j.proeng.2011.
[3] E. Karakulak, A review: Past, present and future of grain refining of 04.382.
magnesium castings, J. Magnes. Alloys 7 (2019) 355–369, doi:10.1016/ [22] X. Liu, H. Zhan, S. Gu, Z. Qu, R. Wu, M. Zhang, Superplasticity in a
j.jma.2019.05.001. two-phase Mg–8Li–2Zn alloy processed by two-pass extrusion, Mater.
[4] N. Mo, Q. Tan, M. Bermingham, Y. Huang, H. Dieringa, N. Hort, M.- Sci. Eng. A 528 (2011) 6157–6162, doi:10.1016/j.msea.2011.04.073.
X. Zhang, Current development of creep-resistant magnesium cast al- [23] X. Liu, R. Wu, Z. Niu, J. Zhang, M. Zhang, Superplasticity at elevated
loys: A review, Mater. Des. 155 (2018) 422–442, doi:10.1016/j.matdes. temperature of an Mg–8%Li–2%Zn alloy, J. AlloysCompd. 541 (2012)
2018.06.032. 372–375 http://dx.doi.org/, doi:10.1016/j.jallcom.2012.06.067.
[5] M. Celikin, M. Pekguleryuz, Creep resistant Mg–Mn based alloys for [24] D.W.A. Rees, Basic Engineering Plasticity: An Introduction with Engi-
automotive powertrain applications, in: D. Orlov, V. Joshi, K. Solanki, neering and Manufacturing Applications, Elsevier, Boston, MA, 2006.
N. Neelameggham (Eds.), Magnesium Technology, Springer, Cham., [25] F. Garofalo, An empirical relation defining the stress dependence of min-
2018, pp. 337–342. TMS 2018. Part F7The Minerals, Metals & Ma- imum creep rate in metals, Trans. Met. Soc. AIME 227 (1963) 351–356.
terials Series, doi:10.1007/978- 3- 319- 72332- 7_51. [26] K. Milička, Creep threshold of an Mg-4Al-1Ca alloy, Kovove Mater. 46
[6] K.P. Rao, Y.V.R.K. Prasad, C. Dharmendra, K. Suresh, N. Hort, (2008) 323–329.
H. Dieringa, Review on hot working behavior and strength of calcium- [27] K. Milička, F. Dobeš, Thresholds in creep of magnesium alloy AX41
containing magnesium alloys, Adv. Eng. Mater. 20 (2018) art. no. composites reinforced with short fibers, Int. J. Mater. Res. 100 (2009)
1701102, doi:10.1002/adem.201701102. 407–412, doi:10.3139/146.110051.
[7] J. Majhi, A.K. Mondal, Microstructure and impression creep character- [28] H.J. Frost, M.F. Ashby, Deformation-Mechanism Maps - The Plasticity
istics of squeeze-cast AZ91 magnesium alloy containing Ca and/or Bi, and Creep of Metals and Ceramics, Pergamon Press, Oxford, United
Mater. Sci. Eng. A 744 (2019) 691–703, doi:10.1016/j.msea.2018.12. Kingdom, 1982.
067. [29] P. Zhang, Creep behavior of the die-cast Mg–Al alloy AS21, Scr. Mater.
[8] P. Qin, Q. Yang, K. Guan, F. Meng, S. Lv, B. Li, D. Zhang, N. Wang, 52 (2005) 277–282, doi:10.1016/j.scriptamat.2004.10.017.
J. Zhang, J. Meng, Microstructures and mechanical properties of a high [30] L.S. Darken, Diffusion, mobility and their interrelation through free
pressure die-cast Mg–4Al−4Gd−0.3Mn alloy, Mater. Sci. Eng. A 764 energy in binary metallic systems, Trans. AIME 175 (1948) 184–201.
(2019) (2019) art. no. 138254, doi:10.1016/j.msea.2019.138254. [31] R. Fuentes-Samaniego, W.D. Nix, G.M. Pound, Vacancy and substi-
[9] M. Pekguleryuz, P. Labelle, Magnesium-based casting alloys having im- tutional solute distribution around an edge dislocation in equilibrium
proved elevated temperature performance, International patent no. WO and in steady-state glide motion, Philos. Mag. A 42 (1980) 591–600,
01/44529, 27 November 2001. doi:10.1080/01418618008241838.
[10] M.O. Pekguleryuz, E. Baril, Creep resistant magnesium diecasting alloys [32] R. Fuentes-Samaniego, W.D. Nix, Appropriate diffusion coefficients for
based on alkaline earth elements, Mater. Trans. 42 (2001) 1258–1267, describing creep processes in solid solution alloys, Scr. Metal 15 (1981)
doi:10.2320/matertrans.42.1258. 15–20, doi:10.1016/0036- 9748(81)90129- 0.
[11] E. Baril, P. Labelle, M.O. Pekguleryuz, Elevated temperature Mg–Al– [33] M.S. Soliman, I. El-Galali, Appropriate diffusion coefficients for dis-
Sr: creep resistance, mechanical properties, and microstructure, JOM 55 location creep in solid-solution alloys, J. Mater. Sci. Lett. 7 (1998)
(2003) 34–39, doi:10.1007/s11837- 003- 0207- 7. 1027–1030, doi:10.1007/BF00720814.
420 F. Dobeš and P. Dymáček / Journal of Magnesium and Alloys 8 (2020) 414–420

[34] Z.L. Bryan, P. Alieninov, I.S. Berglund, M.V. Manuel, A diffusion mo- [37] A. Couret, D. Caillard, An in situ study of prismatic glide in magnesium
bility database for magnesium alloy development, CALPHAD 48 (2015) - I. The rate controlling mechanism, Acta Metall. 33 (1985) 1447–1454
123–130, doi:10.1016/j.calphad.2014.12.001. 1455-1462, doi:10.1016/0001- 6160(85)90045- 810.1016/0001- 6160(85)
[35] J. Friedel, Dislocations, Pergamon Press, Oxford, United Kingdom, 90046-X.
1964.
[36] D. Caillard, J.L. Martin, Some aspects of cross-slip mechanisms in met-
als and alloys, J. Phys. France 50 (1989) 2455–2473, doi:10.1051/jphys:
0198900500180245500.

You might also like