Sie sind auf Seite 1von 5

Materials Science and Engineering A 410411 (2005) 408412

Creep processes in pure aluminium processed


by equal-channel angular pressing
V. Sklenicka , J. Dvorak, P. Kral, Z. Stonawska, M. Svoboda
Institute of Physics of Materials, Academy of Sciences of the Czech Republic, Zizkova 22, CZ-616 62 Brno, Czech Republic
Received in revised form 14 April 2005

Abstract
It is suggested that creep of pure aluminium processed by equal-channel angular pressing (ECAP) occurs by diffusion-controlled movement of
intragranular dislocations and by grain boundary sliding. The contribution of grain boundary sliding is increasingly higher with increasing number
of ECAP passes. The coexistence of a dislocation process and grain boundary sliding may explain the observed decrease of the creep resistance
with increasing number of ECAP passes. This is because high-angle boundaries are necessary in order to achieve grain boundary sliding. Therefore,
an increase in the fraction of high-angle boundaries with increasing number of ECAP passes will essentially lead to increasing contribution of
grain boundary sliding to the total creep strain, as it was observed experimentally.
2005 Elsevier B.V. All rights reserved.
Keywords: Equal-channel angular pressing (ECAP); Aluminium; Ultrafine grained material; Creep; Grain boundary sliding

1. Introduction
The mechanisms controlling the creep properties of pure metals have been usually identified from the dependence of the
minimum and/or steady-state creep rate m on stress , absolute
temperature T and grain size d, using a power-law expression of
the form:
 p


Qc
n 1
m = A
exp
,
(1)
d
RT
where Qc is the activation energy for creep. With this approach,
the fact that n, p and Qc are themselves functions of stress, temperature and grain size is conventionally explained by assuming
that different mechanisms, each associated with different values of n, p and Qc control the creep characteristics in different
stress/temperature regimes. In turn, the dominant mechanisms
under specific test conditions are then generally determined by
comparing experimentally determined values of n, p and Qc
with the values predicted theoretically for different creep mechanisms.
Many investigations concerned with the identification of
creep mechanism have been undertaken using coarse grained

Corresponding author. Tel.: +420 5 4121 2290; fax: +420 5 4121 2301.
E-mail address: sklen@ipm.cz (V. Sklenicka).

0921-5093/$ see front matter 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2005.08.099

pure aluminium [1]. However, it is logical to expect that


the mechanism of hardening/softening observed in the equalchannel angular pressing (ECAP) aluminium may be different
from that observed in the coarse grained material. Consequently,
it cannot be excluded that creep in the ECAP aluminium is controlled by different creep mechanism(s) than that in the coarse
grained material. The present paper reports a series of creep
experiments that were conducted on samples of pure aluminium
processed by equal-channel angular pressing [25].
2. Experimental material and procedures
The starting material used in this investigation was an
extremely coarse grained (grain size 5 mm) high purity
(99.99%) aluminium. Equal-channel angular pressing was conducted at room temperature with a die that had an internal angle
of 90 between the two parts of the channel. The subsequent
extrusion passes were performed by route Bc [2,6] up to 12
passes. For this ECAP route, the rotation of the billet is always
90 in the same sense between successive passes.
Creep tests were performed on the as-pressed samples both in
tension and compression in the temperature interval 423523 K
under an applied stress range between 10 and 25 MPa. For
comparison purposes, some creep tests were performed also on
the unpressed aluminium. Following ECAP and creep testing,

V. Sklenicka et al. / Materials Science and Engineering A 410411 (2005) 408412

409

Fig. 1. Example of grain boundary sliding in the ECAP aluminium (route Bc ,


eight passes) after creep testing at 473 K and 15 MPa. Tensile stress axis is
horizontal.

samples were prepared for examination by means of transmission electron microscope (TEM) and scanning electron
microscope (SEM) equipped with an electron back scattering
diffraction (EBSD) unit. The amount of grain boundary sliding
(GBS) was determined by measuring the surface offsets produced at the intersections of grain boundaries with marker lines
transverse to the stress axis [6,7]. Fig. 1 shows one clear example of the occurrence of grain boundary sliding in creep of the
ECAP aluminium. Longitudinal displacements of the marker
lines, u, due to GBS, together with the fraction of boundaries,
s , with observable GBS, were measured using SEM.
3. Experimental results
3.1. Creep behaviour
Representative creep curves are shown in Figs. 2 and 3. All of
these plots were obtained at a temperature of 473 K (0.5 Tm )
under an initial applied uniaxial tensile or compression stress
of 15 MPa. The creep tests in tension were run up to the final
fracture of the creep specimens, whereas the creep tests in compression were interrupted at a true strain of about 0.35.
Standard versus t creep curves in Figs. 2a and 3a can be
easily replotted in the form of the instantaneous strain rate d/dt
versus strain as shown in Figs. 2b and 3b. As demonstrated by
figures, significant differences were found in the creep behaviour
of the ECAP material when compared to its coarse-grained counterpart. First, the ECAP materials exhibits markedly longer creep
life (Fig. 2a) or markedly longer duration of creep exposure to
obtain a strain of 0.35 (Fig. 3a) than coarse grained aluminium.
Second, the minimum creep rate for the ECAP material is about
one to two orders of magnitude less than that of coarse-grained
material. Third, the shapes of tensile creep curves for the ECAP
material after high number of pressings differ considerably from
the tests conducted at small number of the ECAP passes by the
extent of individual stages of creep.
The difference in the minimum creep rate for the ECAP material and unpressed state consistently decreases with increasing

Fig. 2. Standard creep and creep rate vs. strain curves for unpressed state and
various number of ECAP passes via route Bc (creep in tension up to fracture).

number of ECAP passes (Figs. 2b and 3b). An additional difference is illustrated by Fig. 4, which shows the variation of
the minimum creep rate with the applied stress for the ECAP
specimens after eight passes. The results demonstrate that at
high stresses the minimum compressive creep rate of the ECAP
material may be up to one order of magnitude lower than that of
the unpressed material, although this difference decreases with
decreasing applied stress and becomes negligible at 10 MPa. The
observed values of the stress exponent n = ( ln / ln )T are
6.6 for the unpressed material, 4.8 (creep in compression)
and 5.7 (creep in tension) for the ECAP Al, respectively.
To determine the apparent activation energy for creep Qc,
the minimum creep rate was measured in the temperature interval from 423 to 523 K and at two tensile applied stresses 15
and 20 MPa, respectively. The activation energy for creep Qc is
defined as


ln min
Qc =
.
(2)
(1/kT )

410

V. Sklenicka et al. / Materials Science and Engineering A 410411 (2005) 408412

Fig. 5. Relations between creep rate and temperature for two different stresses.

Thus, the activation energy Qc can be derived from the slope


of log d/dt versus 1/T plots shown in Fig. 5. A value of the apparent activation energy, Qc, was determined by the least square
methods. The Qc is stress dependent and equals to 129.7 16
and 110.9 9 kJ/mol for stresses 20 and 15 MPa, respectively.
3.2. Grain boundary sliding (GBS)

Fig. 3. Standard creep and creep rate vs. strain curves for unpressed state and
various number ECAP passes via route Bc (creep in compression up to strain
0.35).

Grain boundary sliding was measured on the surfaces of the


tensile specimens crept up to a predetermined strain 0.15.
Scanning electron microscopy made it possible to detect GBS
characterized by u 0.1 m. However, GBS was not observed at
all grain boundaries; that is why the relative frequency of sliding
boundaries s was determined. Then the strain component gb
due to GBS is expressed as [7]:
gb =

(1 + )u
s

(3)

was determined by the linear interwhere the mean grain size L


cept method and the overall contribution of GBS to the total
creep strain in the specimen, , was estimated as = gb /. The
results of GBS measurements are summarized in Table 1. It is
evident that the fraction of boundaries s increases as the number
of ECAP passes increases. This result supports the idea that GBS
is connected with microstructural changes of grain boundaries
[5]. It is to note that in the best case (12 passes) the contribution
of GBS to creep strain is only 33%.
Table 1
Summary of GBS measurements (
= 0.15)

Fig. 4. Stress dependences of creep rate for unpressed state and eight ECAP
passes.

No. of passes

u (m)

(m)
L

gb 102

gb/ 102

1
2
4
8
12

0.51
0.48
0.55
0.49
0.52

0.80
0.83
0.93
0.91
0.92

14.9
12.7
12.2
10.8
11.0

3.15
3.60
4.80
4.70
5.00

21.0
24.0
32.0
31.0
33.0

V. Sklenicka et al. / Materials Science and Engineering A 410411 (2005) 408412

411

3.3. Microstructural evolution during creep


Transmission electron microscopy results have shown that
one ECAP pass at room temperature leads to a substantial reduction in the grain size (1.4 m) and the microstructure now
consists of parallel bands of grains in the shearing direction.
The microstructure is very inhomogeneous and the grain size
varies location to location. The inhomogeneous nature of the
microstructure may reflect the coarse grain size (5 mm) prior to
ECAP. The grains subsequently evolve with further pressings at
room temperature into a reasonably equiaxed and homogeneous
microstructure with an average grain size of 1 m. Fig. 6a
shows a TEM micrograph from the longitudinal section of an
Al sample after eight ECAP passes.
To explore the thermal stability of ECAP processed aluminium annealing was conducted at a temperature of 473 K (i.e.
at the temperature of the intended creep tests) for different periods of time under no external load. It was found that a significant
growth of the ECAP grain size occurred at the very beginning of
annealing (up to 30 min). Thus, static annealing at 473 K introduces a significant grain growth leading to a stable average grain
size of around 12 m.
TEM observations were used also to establish details of
microstructure evolution during creep. The micrographs in
Fig. 6b and c illustrate a dislocation substructure inside the
grains. The dislocation lines are wavy and occasionally tangled
with each other. It is known that large grains in ultrafine grained
materials contain dislocations while grains smaller than a certain
size are dislocation free [4,8].
EBSD measurements were taken to determine the grain
boundary misorientations after creep. The value of the relative fraction of a high-angle ( > 15 ) grain boundary population
increases rapidly during the first four passes and slightly during
subsequent ECAP pressing (Fig. 7).
4. Discussion
According to Eq. (1) the stress dependence of the minimum
creep rate (Fig. 4) at constant temperature and grain size can be
described by the power-law relationship of the form:
= A n .
The observed values of n are 4.8 and/or 5.7 and 6.6 for the
ECAP samples and unpressed material, respectively. A similar
value of n was reported by Tobolova and Cadek [9] for creep
in polycrystalline aluminium of 99.99% purity with a standard
grain size about 400 m. For n 4 creep is known to occur
by diffusion-controlled movement of dislocations within grains
(slip creep) and by grain boundary sliding. The stress dependence of the grain boundary sliding rate can be described by a
similar way: gb
= A n , where the stress exponent of grain
boundary sliding n
= 0.8n [10]. Thus, the slightly lower values
of n for the ECAP material may reflect the effect of more intensive grain boundary sliding in the creep of the ECAP material
in comparison to its coarse-grained counterpart.
The mechanism which most probably plays the dominant
role in the power-law creep of aluminium is dislocation climb
controlled by lattice self-diffusion [1,10]. Therefore, the activation energy for creep Qc should be the same as the value of the

Fig. 6. TEM micrographs from the longitudinal section of an aluminium processed by ECAP route Bc : (a) after eight ECAP passes, (b) after one ECAP
pass and creep, and (c) after eight ECAP passes and creep. Creep at 473 K and
15 MPa.

activation enthalpy of lattice self-diffusion HSD in aluminium


(HSD
= 127143 kJ/mol [10]). The slightly higher experimental value of the activation energy for creep Qc
= 151 kJ/mol was
found by Tobolova and Cadek [9] for polycrystalline aluminium.
The values obtained for Qc (Fig. 5) in this work are somewhat
lower than that of HSD . Supposing that grain boundary sliding is controlled by grain boundary diffusion, which usually

412

V. Sklenicka et al. / Materials Science and Engineering A 410411 (2005) 408412

aluminium occurs by lattice diffusion-controlled movement of


dislocations being grain boundary sliding increasingly important with increasing the number of ECAP passes.
It was found, that the creep resistance of high purity aluminium is increased considerably already after one ECAP
pass. However, successive ECAP pressings lead to a noticeable
decrease in the creep properties. The results from earlier investigations [3,4,11] indicate that an inhomogeneity of the ECAP
microstructure in mesoscopic scale may influence the creep
behaviour of the pressed material. Accordingly, more detailed
and complex work on grain anisotropy and homogenization of
grain structure is needed to understand the effect of the ECAP
pressing on creep behaviour.
Acknowledgements

Fig. 7. The fraction of high-angle grain boundaries in the crept specimens as a


function of the number of ECAP passes.

is assumed to be about 0.7 times that for lattice self-diffusion,


our results give support to the assumption that grain boundary
sliding is increasingly important in the ECAP aluminium. Further, the contribution of grain boundary sliding to the total strain
increases at the same temperature with a decrease in the applied
stress [10]. This can explain the lower value of Qc determined for
= 15 MPa in comparison with the value of Qc for = 20 MPa.
Finally, since high-angle grain boundaries are necessary in
order to achieve GBS, an increase in the fraction of high-angle
boundaries with increasing number of ECAP passes (Fig. 7) will
essentially lead to an increasing contribution of GBS to the total
creep strain as it was observed experimentally (Table 1).
5. Summary and conclusions
Extremely coarse grained aluminium (99.99%) was subjected
to severe plastic deformation (ECAP) at room temperature using
multiple passes (up to 12) and processing route Bc . Creep tensile
and compression tests were conducted on the ECAP material and, for comparison purposes, on unpressed aluminium.
Based on the results, it is suggested that creep in the ECAP

Financial support for this work was provided by the Grant


Agency of the Academy of Sciences of the Czech Republic under
Grant A204 1301.
References
[1] B. Wilshire, C. Palmer, in: R.S. Mishra, J.C. Earthman, S.V. Raj (Eds.),
Creep Deformation. Fundamentals and Application, TMS, Warrendale,
2002, p. 51.
[2] M. Furukawa, Y. Iwahashi, Z. Horita, M. Nemoto, T.G. Langdon, Mater.
Sci. Eng. A 256 (1998) 328.
[3] S.D. Terhune, D.L. Swisher, K. Oh-Ishi, Z. Horita, T.G. Langdon, Metall.
Mater. Trans. A 33 (2002) 2173.
[4] R.Z. Valiev, R.K. Islamgaliev, I.V. Alexandrov, Progr. Mater. Sci. 45
(2000) 103.
[5] V. Sklenicka, J. Dvorak, M. Svoboda, Mater. Sci. Eng. A 387389 (2004)
696.
[6] V. Sklenicka, J. Dvorak, M. Svoboda, in: Y.T. Zhu, T.G. Langdon, R.Z.
Valiev, S.L. Semiatin, T.C. Lowe (Eds.), Ultrafine Grained Materials III,
TMS, Warrendale, PA, 2004, p. 647.
[7] V. Sklenicka, J. Cadek, Z. Metallkd. 61 (1970) 575.
[8] G.B. Raab, R.Z. Valiev, T.C. Lowe, Y.T. Zhu, Mater. Sci. Eng. A 382
(2004) 30.
[9] Z. Tobolova, J. Cadek, Phil. Mag. 26 (1972) 1419.
[10] J. Cadek, Creep in Metallic Materials, Elsevier Science Publishers, Amsterdam, 1988.
[11] L. Ilucova, P. Kral, M. Svoboda, I. Saxl, V. Sklenicka, in: C. Gundlach,
et al. (Eds.), Evolution of Deformation Microstructures in 3D, Riso
National Laboratory, Roskilde, 2004, p. 363.

Das könnte Ihnen auch gefallen