Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Materials Science & Engineering A 711 (2018) 343–349

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Basic creep modelling of aluminium T


a,⁎ b
S. Spigarelli , R. Sandström
a
DIISM, Università Politecnica delle Marche, via Brecce Bianche, 60131 Ancona, Italy
b
Materials Science and Engineering, KTH Royal Institute of Technology, Brinellvägen 23, S-10044 Stockholm, Sweden

A R T I C L E I N F O A B S T R A C T

Keywords: In recent years a basic creep model that does not involve adjustable parameters has been developed. The main
Creep feature of this model is that it is fully predictable and the assumptions at its basis can be easily verified once the
Aluminium output is compared with experimental data. This model, initially developed for pure copper, has been here
Plasticity applied to pure aluminium. A critical issue has been identified with the controlling mechanisms during power-
Modelling
law breakdown. The increase in the creep rate at high stresses and low temperatures can be quantitatively
Climb
explained from the raised climb rate due to the deformation-induced increase in concentration of vacancies. The
Glide
model can also account for the fairly wide range of stresses where aluminium follows power-law creep with a
creep exponent of 4–5. At slightly lower stresses, the creep exponent increases somewhat due to the presence of
an internal stress. Since no adjustable parameters have been required, the model represents a notable en-
hancement over the conventional approach, which is based on the use of the power-law equation and requires
fitting of experimental data to determine the values of the material parameters.

1. Introduction applied stress in double-log coordinates increases progressively with


applied stress level. At room temperature, the stress exponent reaches
The creep response of pure aluminium has attracted a great deal of up to 28 [4]. The behaviour associated to a stress exponent of 4–5 and
academic interest, leading to the publication of many papers, if one to an activation energy equivalent to that for self-diffusion identifies the
considers their limited, to be fair, industrial relevance as creep-resistant so-called “class M” (Metal) materials. The typical class-M response of
alloy. A non-exhaustive picture of the extensive coverage of this subject pure coarse grained Al is shown in Fig. 1, which plots the steady state
can be obtained from the relevant chapters in [1–3]. The cause for such creep rate as a function of stress from the well-known Servi-Grant da-
a debate can be traced back to the very nature of this material, whose taset for Al 99.995% (grain size 2 mm) [5] and other results on Al
simple microstructure was an ideal case-study to identify the main 99.999% obtained by Mecking at al (grain size 220 µm) [6].
dislocation creep mechanisms and separate their relative contributions Several theories, reviewed in [7,8], have been later developed to
to creep strength. justify the power law creep behaviour of pure metals and the resulting
In most cases, the secondary creep rate (ε )̇ dependence on applied values of stress exponent and activation energy. Although Nabarro
stress (σ) and temperature (T) has been described by the conventional considered these theories “unconvincing” [8], they could nevertheless
power-law and Arrhenius equations, in the form represent a basis to develop new creep models for more complex ma-
terials. Yet, the creep community was almost invariably satisfied with
D0sd Gb σ n Q
ε̇ = A ⎛ ⎞ exp ⎛− sd ⎞ using the power law in its phenomenological form to describe the re-
kT ⎝ G ⎠ ⎝ RT ⎠ (1)
sponse of innumerable materials, even with extremely complex mi-
where A is a material parameter, k is the Boltzmann constant, G is the crostructures (see [9–16] for just few very recent examples). In these
shear modulus (G = 30220-16T MPa for Al), b is the length of the cases, n, Q and A are calculated by a best fitting procedure of the ex-
Burgers vector (b = 2.86 × 10−10 m for Al) and R is the gas constant. perimental data and cease to have a direct correlation with micro-
D0sd in Eq. (1) is the pre-exponential factor in the Arrhenius equation structural phenomena. Even in those cases where authors attempt to
for self-diffusion and Qsd is the activation energy for self-diffusion in the develop constitutive models based on the physics of phenomena, a
metal. The stress exponent n in aluminium is about 4–5 at temperatures number of material parameters need to be calculated by fitting the
above 250 °C [4,5]. Above a certain stress, power-law breakdown oc- experimental data (see, for example [17]). The fact is that the power
curs and the slope of the curve describing the strain rate dependence on law remains phenomenological in nature, and for this reason,


Corresponding author.
E-mail address: s.spigarelli@univpm.it (S. Spigarelli).

https://doi.org/10.1016/j.msea.2017.11.053
Received 27 September 2017; Received in revised form 10 November 2017; Accepted 14 November 2017
Available online 16 November 2017
0921-5093/ © 2017 Elsevier B.V. All rights reserved.
S. Spigarelli, R. Sandström Materials Science & Engineering A 711 (2018) 343–349

Nomenclature Qsd activation energy for vacancy diffusion (self diffusion)


[J mol−1]
Symbols R universal gas constant [J mol−1 K−1]
Rmax maximum back stress [Pa]
b Burgers vector [m] T absolute temperature [K]
c0 equilibrium vacancy concentration α material constant in Taylor equations
CL work hardening constant ε strain
d grain size [m] ε̇ strain rate [s−1]
dsub sub-grain size [m] λ spacing between vacancy sinks
D0sd pre-exponential factor in Arrhenius equation for self dif- ν Poisson’s ratio
fusion [m2 s−1] ρ free dislocation density [m−2]
G shear modulus at the testing temperature [Pa] ρa free dislocation density in annealed state [m−2]
k Boltzmann constant [J K−1] σ applied stress (creep) or flow stress (constant strain rate
kd Hall-Petch constant [Pa m0.5] experiments) [Pa]
L mean dislocation free path [m] σi internal stress [Pa]
m Taylor factor σy yield strength [Pa]
Mc climb mobility of dislocations τl dislocation line tension [N m−1]
Mcg climb and glide mobility of dislocations ω recovery constant
n stress exponent in power-law equation

Fig. 1. Experimental steady state creep rate as a


function of stress for Al99.995% [5] and Al99.999%
[6]. In the latter case, the data were obtained from
constant stress creep experiments and constant strain
rate tests.

alternative approaches have been sought to introduce constitutive subgrain size (dsub). When grain and subgrain sizes become equivalent,
equations directly derived from the operating mechanisms. In this line deformation mechanisms change and grain boundaries get a more
of thought, Sandström has proposed models for plastic deformation of dominant influence [24]. Materials can be classified according to their
fcc metals based on physical mechanisms [18,19]. These models have grain size. For example, Mohamed gives 20 µm as the lower limit for
been successfully applied to copper and austenitic stainless steels, see large grain materials. Micrograined materials fall in the range 1–10 µm,
for example [20,21], and can handle not only power-law creep but also ultrafine-grained (UFG) materials have grain sizes between 300 and
power-law breakdown. For example, the equations proposed can cor- 900 nm and nanocrystalline (nanograined) materials between 20 and
rectly predict a creep exponent as high as 65 for copper at 75 °C [21]. 200 nm [25]. A group of ultrafine nanocrystalline materials with grain
These models do not require any best-fitting procedures, being based sizes below 10 nm can also be introduced. Although fine-grained ma-
only on a number of physical and microstructural pre-determined terials will not be discussed in this paper, it might be of interest to give
parameters, properly combined in a set of easy-to-handle constitutive some brief general comments, since deformation mechanisms in fine
equations [22,23]. Since no adjustable parameters are involved, the grained materials are extensively covered in literature and in particular
models are fully predictable. Thus, to predict the creep rate, the only nanocrystalline materials.
data required are the experimental ones (stress and temperature, in the When grain size is reduced, deformation processes at the grain
case of a pure metal, where composition is not an issue). The boundaries start to influence the creep rate. There are numerous models
straightforward development of the model is its application to the other for describing how this can take place. Most authors agree that the
metal which, with Cu, constitutes the classical case study of creep of dominating mechanism is grain boundary sliding (GBS), i.e. deforma-
class-M materials, i.e. pure Al [8]. Only once the suitability of the basic tion takes place by neighbouring grains sliding against each other. GBS
model to describe the creep response of the high-purity metal is at- cannot occur without an accommodating mechanism around the grain
tested, subsequent developments to describe the effects of the atoms in boundaries. A number of empirical models for GBS have been for-
solid solution and of secondary-phase particles can be planned. mulated for fine-grained materials, but no basic model seems to be
The model presented in [18–23] is applicable to fcc metals with available (for a review, see [26]). In agreement with experiments, these
large grain size. This means that grain size (d) should be larger than models suggest a stress exponent of 2 and a creep rate that is inversely

344
S. Spigarelli, R. Sandström Materials Science & Engineering A 711 (2018) 343–349

proportional to the square of the grain size for micro-grained materials


[26,27]. This dependence is also frequently observed for superplastic
alloys. For both micro-grained and ultrafine-grained materials, the ac-
commodation process is believed to occur by ordinary dislocation slip.
It is therefore quite possible that the behaviour in the micro-grained
range can be extended into the ultrafine-grained range, but little ex-
perimental data is available to confirm it. Nanocrystalline materials
have a reduced dislocation density. In spite of this, accommodation
takes place by ordinary dislocation slip [28] and by emission and ab-
sorption of partial dislocations at the grain boundaries [29]. The partial
dislocations lead to the formation of deformation twins that can im-
prove both strength and ductility by increasing the dislocation and
thereby work hardening. The presence of deformation twins has been
confirmed both by means of transmission electron microscopy and
molecular dynamics (MD) simulations. This is also observed for high
stacking fault materials such as aluminium [30]. For ultrafine nano-
crystalline materials, the accommodation mechanism is assumed to be
based on diffusion creep. This result is primarily obtained through MD
simulations [31]. In summary, GBS is believed to be of importance at all
grain sizes, but the accommodation processes vary with grain size. Fig. 3. Diffusion coefficient for self diffusion (see ref [47] for the sources of the original
The brief analysis above clearly shows that very fine-grained Al experimental data). The figure also plots the curves describing the most widely used
alloys behave quite differently with respect to conventional coarse- relationships (Qsd = 143) and the dependence of the coefficient of diffusion on tem-
grained alloys, where creep is controlled by dislocation climb and glide perature recalculated in this study (solid lines, Qsd = 122 kJ mol−1).

and class-M behaviour can be easily identified. For this reason, they will
be not analysed in this paper, which will rather deal with dislocation dislocations in the sub-boundaries and those in the subgrain interiors.
creep of pure coarse-grained aluminium. The basic model introduced in This applies for example to cold worked materials [32]. However, this
[18–23] will be applied. In particular, the model results in an enhanced distinction is not important for pure aluminium at high temperature,
creep rate at high stresses and low temperatures, which, as mentioned where free dislocation density plays a major role, and will therefore be
above, originates from a phenomenological expression. Although ver- ignored here. The free dislocation density (ρ) is usually related to the
ified experimentally a number of times, it can still be questioned. Thus, stress by the well-known Taylor equation, written in the form
a theoretical justification of the enhanced dislocation mobility in the
region of “power law breakdown” will be also presented, by introducing σ = σi + σd = σi + αmGb ρ (2)
a new physical parameter, i.e. the increase in vacancy concentration.
where m is the Taylor factor (m = 3.06 for fcc metals) and
σd = αmGbρ1/2 is the dislocation hardening term. The internal stress σi
2. The model represents the strength of the pure annealed large-grained metal, i.e.
the stress required to move a dislocation in the absence of other dis-
2.1. Basic dislocation model locations, and α is a constant (α = 0.2–0.4; the intermediate value 0.3
will be considered in the following).
The model, originally developed for Cu, is based on physically de- The evolution of the dislocation density during straining can be
rived equations. It will now be applied to pure aluminium. expressed as [18,32]
The total dislocation population is formed by dislocations stored in
subgrain boundaries and free dislocations forming a network inside the dρ m 2
= − ωρ − Mτl ρ2
subgrains. In special cases, it is essential to distinguish between dε bL ε̇ (3)

Fig. 2. Comparison of climb glide factor fclgl in Eq.


(18) with increase in vacancy concentration due to
plastic deformation gclimb in Eq. (22) as a function of
stress, using experimental values for stress and creep
rate from [6]. A detail of the high-T trends is given
on the right.

345
S. Spigarelli, R. Sandström Materials Science & Engineering A 711 (2018) 343–349

Fig. 4. Description of the steady state creep rate as a


function of the applied stress described by the p = 2
model, and comparison with the experimental data
presented in Fig. 1. The curvature in the very-low
stress/ high-temperature regime is probably over-
estimated, due to a parallel overestimation of the
internal stress in these conditions.

Fig. 5. Comparison of climb glide factor fclgl in Eq. (18) with the increase in vacancy
concentration due to plastic deformation gclimb, Eq. (22) as a function of temperature.
Fig. 6. Free dislocation density as a function of modulus compensated stress for pure Al
Results are shown for five strain rates.
[39,48] and Al-Zn (from [40]). The curves were obtained by Eq. (2) with α = 0.3, being σi
= 0 (broken line) or σi given by Eq. (16) (solid line). Also in this case, the effect of a slight
where ω is a constant (ω = 14.7 in pure Cu), τl is the dislocation line overestimation of the internal stress in the very-low stress regime is noticeable.
tension (τl = 0.5Gb2), M is the dislocation mobility and L is the dis-
location mean free path, i.e the distance travelled by a dislocation be- aware of. In this context, for consistency reasons, in this work the ter-
fore it undergoes a reaction, customary expressed as minology used in [18,19,22,23] was adopted.
The dislocation climb mobility, according to Hirth and Lothe [33],
CL
L= is
ρ (4)
D0sd b σ b3 Q
Mc = exp ⎛ d ⎞ exp ⎛− sd ⎞
⎜ ⎟

CL being a strain-hardening constant. Eq. (4) remains valid as long kT ⎝ kT ⎠ ⎝ RT ⎠ (5)


as it gives a value of the dislocation mean free path smaller than the
Creep results obtained in the low-temperature/high strain rate re-
grain size. This implies that, since the dislocation mean free path cannot
gime are characterised by higher stress exponents and lower values of
exceed the grain size, for vey fine-grained materials, under very low
the activation energy for creep. It is often suggested that this requires
applied stresses, i.e. for very low dislocation densities, L should equal d.
that glide is taken into account. Following [34], the glide of dislocations
This is not obviously the case for the coarse-grained materials
through an obstacle field was described by a phenomenological equa-
(d > 100 µm) here considered.
tion in the form
The first term on the right-hand side of Eq. (3) represents the strain
hardening effect due to dislocation multiplication, which is more rapid p q
dε Q ⎡ σ
when L and, consequently, CL assume low values and/or the dislocation = fσ 2 exp ⎧ − 1−⎛ d ⎞ ⎤ ⎫ ⎜ ⎟

dt ⎨ RT ⎢⎣ ⎝ Rmax ⎠ ⎥
⎦ ⎬ (6)
density is high. The second term on the right-hand side of Eq. (3) de- ⎩ ⎭
scribes the effect of dynamic recovery, i.e. strain dependent recovery. where f, Q, p, q and Rmax are unknown parameters. Rmax was specified
The third term takes into account static recovery, which is time de- as the maximum back stress and was chosen as the true tensile strength,
pendent recovery. Unfortunately, the different types of recovery are not corrected to take into account the effect of necking in very ductile
named consistently in literature, which is something one should be metals. To find the values of f and Q, a unified model for climb and

346
S. Spigarelli, R. Sandström Materials Science & Engineering A 711 (2018) 343–349

glide was formulated according to a procedure proposed by Nes et al m ρT 0.5 m α m2G


CL = = =
[35]. The resulting expression for the combined glide and climb mo- 0.5
bωρT ρRT 0.5 bωρRT 0.5 ω (Rmax − σy ) (14)
bility takes the form [23]
where the suffixes T and RT indicate high-T and room-T respectively,
p q
D b σ b3 Q σ i.e. the same result as in Eq. (11). In the derivation, it was assumed that
Mcg = 0sd exp ⎛ d ⎞ exp ⎧ − sd ⎡1 − ⎛ d ⎞ ⎤ ⎫
⎜ ⎟ ⎜ ⎟

kT ⎨ RT ⎢ ⎥
max ⎠ ⎦ ⎬
there is no temperature dependence of CL, an assumption that is con-
⎝ kT ⎠ ⎩ ⎣ ⎝ R ⎭ (7)
sistent with the result of Eq. (14). With Rmax ≅ 75 MPa (1.5RUTS) and a
Originally, the ranges 0 < p ≤ 1 and 1 < q ≤ 2 were suggested yield strength close to 20 MPa, CL = 85 is obtained. For a dislocation
[34]. Other studies showed that Eq. (6) works very well for copper and density as low as 5 × 1011 m−2, this value gives a dislocation mean free
austenitic stainless steels with p = 2 and q = 1 with Rmax equivalent to path close to 120 µm, which is well below the grain size of the materials
the true stress corresponding to the ultimate tensile strength of the considered in Fig. 1.
material [21]. This has also been demonstrated in other works [3,36].
2.3. Internal stress
2.2. Creep parameters The determination of the internal stress, which is temperature and
At steady state, with the climb glide mobility according to Eq. (7), strain rate dependent, is here based on the assumption that the an-
Eq. (3) gives nealed dislocation density (ρa) and the σi values account for the an-
nealed yield strength of the pure metal. The yield stress is thus given by
2Mcg τl bCL σ 3
Eq. (2) [40], where ρa is virtually nihil. The term σ0 includes the de-
εsṡ = ⎛ d ⎞
m − ωCL (σd/ αmG ) ⎝ αmGb ⎠ (8) pendence on grain size, that is

The term in the denominator that involves ω is a consequence of the kd


σi = σi0 +
dynamic recovery. It is especially important when describing tertiary d (15)
creep [37]. The model based on the combination of Eqs. (7) and (8) where σi0, the internal stress for a material with an infinitely large grain
requires the determination of two constants (CL and ω) and of σi. Roters size, depends on the impurity content of the metal, while kd is the Hall-
et al. gave a derivation of ω [38] that has been extensible applied to Petch constant. The variation of kd as a function of temperature, re-
copper [18] ported by Blum et al. [4], was used to estimate the effect of grain size on
the yield stress of high-purity Al as a function of temperature (grain size
m d int ⎛ 1 ⎞
ω= ⎜2 − ⎟ d1 = 1 mm, strain rate 5 × 10−4 s−1 [40].
b ⎝ nslip ⎠ (9) A way to describe the temperature dependence of the yield strength
is to assume that it is proportional to the creep strength in the creep
being dint the interaction distance where dislocations of opposite sign
range, whereas below the creep range the yield strength is proportional
can annihilate each other and nslip = 12 is the number of slip systems;
to the shear modulus. Pure Al data cover all the interval between these
dint can be taken as dislocation core diameter. With ab initio calcula-
two extreme cases, and for this reason the following expression was
tions, a core diameter for aluminium of 2.5 b was obtained [39], which
used
gives ω = 15.
In Eq. (3) the dynamic and static recovery terms give a maximum σi = Ay σcreep G (16)
dislocation density. For the dynamic recovery term, this maximum
dislocation density is where σcreep is the creep stress which corresponds to a given steady state
creep rate. For a strain rate of 5 × 10−4 s−1 as in [40], the Servi and
2
m ⎞ Grant dataset for pure-Al with a 2 mm grain size [5] was used to obtain
DR
ρmax =⎛ ⎜ ⎟

⎝ Lω ⎠
bC (10) the relevant creep stress at different temperatures. The Ay constant was
determined to obtain a reliable estimate of the yield strength of high-
At low temperatures, the dynamic recovery and static recovery purity coarse-grained Al. The results presented in [40] for an 99.999 Al
terms in Eq. (3) have the same size as for copper when the maximum with d = 1 mm can be used to recalculate the yield stress (σyi) of a
dislocation densities are inserted [6]. By analysing creep data, it can be material with a different grain size (di)
shown that the same applies to aluminium at room temperature. This
kd k
means that Eq. (10) can be used to determine CL at room temperature σyi = σy1 − + d
d di (17)
m α m2G
CL = DR 0.5
= which, combined with Eq. (2), gives, for di = 2 mm and ρa = 1 ×
bω (ρmax ) ω (Rmax − σy ) (11)
1010 m−2, Ay = 4.2 × 10−3.
In the second equality, the Taylor Eq. (2) is used. The maximum
dislocation stress is estimated as the difference between the tensile 2.4. Rate enhancement at low temperatures and high stresses
strength Rmax and the yield strength σy. Comparing Eqs. (5) and (7) one finds that the dislocation mobility is
At higher temperatures, the static recovery term in Eq. (3) is larger enhanced by the following factor
than the dynamic recovery term. A simplified version of Eq. (8) is then 2
Qsd ⎛ σd ⎞ ⎤
obtained fc lg l = exp ⎡ ⎜
⎢ RT Rmax ⎥

⎣ ⎝ ⎠ ⎦ (18)
2Mcg τl bCL
ε̇ = ρ1.5 assuming p = 2 and q = 1. This expression will now be justified.
m (12)
During plastic deformation the number of vacancies increases. The non-
The ratio between the dynamic and static recovery in Eq. (3) is conservative motion of dislocations generates and absorbs vacancies.
ω ε̇ ω bCL 0.5 Mecking and Estrin formulated a model, which demonstrated that there
= ρ is a net increase in the vacancy concentration due to plastic deforma-
2Mcg τl ρ m (13)
tion [41]. Their result can be expressed as
where Eq. (12) has been inserted. As mentioned above, this term is
Δc 2 λ2ε ̇ σd
close to unity at room temperature, so the change in the dislocation = 0.5
c0 Dsd G (19)
density must be taken into account

347
S. Spigarelli, R. Sandström Materials Science & Engineering A 711 (2018) 343–349

c0 is the thermal equilibrium vacancy concentration, Δc = c – c0 is temperature dependence is reasonably well described.
the excess concentration, λ is the spacing between vacancy sinks. In Fig. 5 plots the fclgl and gclimb factors, now calculated from the model
[41], a factor 0.1 was used in Eq. (19). However, a detailed derivation values of the strain rate. The excellent agreement already observed in
shows that it should be replaced by 0.5 and a factor 2 should be in- Fig. 2 is confirmed, which indicates that a detailed verification of Eq.
troduced. Following [41], λ is taken as the subgrain size dsub. It is well (18) has been derived.
known that the subgrain size can be related to the applied stress σ An interesting implication of the model used in this study is that it
Ksub bG results in a marked curvature of the strain rate vs stress plots for stresses
λ = dsub = below 1 MPa. This effect is simply caused by the presence of an internal
σ (20)
stress that, once calculated by Eq. (16), becomes non-negligible in this
Ksub is a constant that is about 18 for aluminium [42]. Inserting Eq. region when compared to the stress applied. Although Eq. (16) can be
(20) into Eq. (19) gives supposed to overestimate the internal stress in the very-low stress/very-
Δc 2 Ksub2ε ̇ b2 σd G high temperature region, the same experimental behaviour was indeed
= 0.5 observed also by other authors, at higher temperature [48].
c0 Dsd σ2 (21)
Fig. 6 plots a collection of dislocation density values as a function of
This increase in the vacancy concentration raises the climb rate in stress for Al [40,49] and for Al-Zn, which is thought to behave as the
proportion. Thus, the resulting enhancement factor for the climb rate pure metal on this regard [40]. The same Fig. 6 shows the curves ob-
gclimb is tained by Eq. (2) with α = 0.3 and σi = 0, and with σi given by Eq. (16).
Δc The description of the experimental data is excellent, except in the low
gc lim b = 1 + stress region, where, as mentioned above, Eq. (16) is thought to over-
c0 (22)
estimate the internal stress. Incidentally, it is interesting to observe that
The expression (18) is compared to the vacancy concentration en- the data reported in the Figure for pure-Al, with the exception of the
hancement (Eq. (22)) in Fig. 2. In the Figure, experimental values for value corresponding to the lowest stress, were obtained by relating the
the creep rate are used and a good general agreement between Eqs. (18) RT yield stress after various levels of severe plastic deformation by
and (22) is obtained. The temperature, stress and strain rate de- ECAP (Equal Channel Angular Pressing) and the corresponding dis-
pendencies are well covered. Quite a surprising result was obtained. location densities measured by X-ray analysis [49]. Data in Fig. 6,
The enhancement of the creep rate in relation to the high temperature which thus represent the total dislocation content at room temperature,
climb model in Eq. (7) is well justified for copper and austenitic closely align on the same curve identified by steady state flow stress at
stainless steels [20]. It was inspired by an empirical model for glide. high-T and the corresponding free dislocation contents [40]. At high
Instead, it turns out that it can be fully explained by the enhancement of temperature, torsion experiments in pure-Al demonstrated that the
the climb rate at low temperatures due to the increased vacancy con- steady-state value of the flow stress is unaffected by substantial changes
centration. This should not be considered as the final proof that creep at in subgrain size and sub-boundary orientation [50]. On the other hand,
ambient temperatures is fully controlled by climb. However, it makes it cell formation in Cu was observed to significantly affect the creep re-
much simpler to explain why full stationary creep is observed in alu- sponse at low temperatures [32], an effect that was attributed to the
minium and copper at near ambient temperatures. Yet, taking glide and high density of dislocation forming cell-walls. A similar mechanism can
cross slip into account results in a continuous build-up of a forest of be expected to be operative also in pure Al, and, as a matter of fact, the
edge dislocations, which would make the creep rate slow down and model presented in Section 2, which considers only the free disloca-
eventually stop, as it is observed for logarithmic creep [43]. tions, somewhat overestimates the steady state creep rate at 300 and
366 K (Fig. 4). Gubicza et al. [49] reasoned that the fact that the yield
3. Description of high purity aluminium strength obeys the Taylor equation using the dislocation density values
determined by X-ray line profile analysis attests that the main
The two datasets shown in Fig. 1 are substantially consistent with strengthening mechanism is still the interaction between dislocations,
each other, except at intermediate temperatures, where the Servi-Grant also in these ultrafine-grained material, irrespective whether they are
creep rates seem to be somewhat lower when compared with the data stored in cell walls or are free in cell interior.
from [6].
The first point to be addressed is the selection of the proper value of 4. Conclusions
the self-diffusion coefficient. As a matter of fact, this apparently trivial
problem requires an analysis of the most recent findings on this subject. A basic model for the description of creep in fcc metals has been
The “traditional” values of the diffusion coefficient parameters for self- applied to pure Aluminium.
diffusion in Al are D0sd = 1.86 × 10−4 m2 s−1, Qsd = 143.4 kJ mol−1 The set of physically based constitutive equations proposed contains
[44] or D0sd = 1.72 × 10−4 m2 s−1 and Qsd = 142.1 kJ mol−1 [45,46]. two important parameters: the internal stress, representing the stress
The accuracy of these estimates was challenged in [47], which reported required to move a dislocation in the matrix and the strain hardening
a wide collection of literature results (Fig. 3). The Figure shows how the constant CL, which, in combination with the free dislocation density,
value of 143 kJ mol−1 can describe the high-temperature literature determines the dislocation mean free path.
data very well, but strongly underestimates the value of the diffusion A comparison between model prediction and data suggests that
coefficient at lower temperatures. An excellent description of the data power-law breakdown, i.e. the increase in strain rate at low tempera-
collected in [47] is rather obtained with Qsd = 122 kJ mol−1 and D0sd tures, can be quantitatively explained from the raised climb rate due to
= 8.3 × 10−6 m2 s−1. the deformation-induced increase in concentration of vacancies. The
Since all the parameters have now been quantified, the equations model proposed, which does not require any specific variation in con-
can be used to model the steady state creep rate dependence on applied stitutive equations to describe the experimental range of steady state
stress, by simply substituting σ and T into Eqs. (7), (8) and (16), with CL creep rate, can also account for the fairly wide range of stresses where
= 85, ω = 15, Rmax = 75 MPa, Ay = 4.2 × 10−3, Qsd = 122 kJ mol−1 aluminium follows power-law creep with a creep exponent of 4–5.
and D0sd = 8.3 × 10−6 m2 s−1 (Fig. 4), without any data-fitting. The Since the model does not include any adjustable parameters, no data
correlation between the theoretically and calculated curves and the fitting is required to obtain an excellent description of the secondary
experimental data from the Mecking et al dataset is excellent; a larger creep rate with stress. In this sense, the model represents a notable
deviation is indeed observed in the case of the Servi-Grant results, at enhancement over the conventional approach, which is based on the
intermediate (533 K) temperatures, although on a whole the use of the power-law equation, thus constituting an excellent base for

348
S. Spigarelli, R. Sandström Materials Science & Engineering A 711 (2018) 343–349

further development aimed at the description of single phase solid so- superplasticity, Acta Metall. Mater. 42 (1994) 2437–2443.
lution alloys (class A metals) and age hardening aluminium alloys. [25] F.A. Mohamed, Deformation mechanism maps for micro-grained, ultrafine-grained,
and nano-grained materials, Mater. Sci. Eng. A528 (2011) 1431–1435.
[26] F.A. Mohamed, H. Yang, Deformation mechanisms in nanocrystalline materials,
References Metall. Mater. Trans. A 41 (2010) 823–837.
[27] F.A. Mohamed, On the creep transition from superplastic behavior to nanocrystal-
line behavior, Mater. Sci. Eng. A655 (2016) 396–398.
[1] H. Oikawa, T.G. Langdon, The creep characteristics of pure metals and metallic
[28] F.A. Mohamed, Correlation between the deformation of nanostructured materials
solid solution alloys, in: R. Wilshire, R.W. Evans (Eds.), Creep Behaviour of
and the model of dislocation accommodated boundary sliding, Metall. Mater. Trans.
Crystalline Solids, Pineridge Press, Swansea, 1985, pp. 33–82.
A 39 (2008) 470–472.
[2] J. Čadek, Creep in Metallic Materials, Elsevier, Oxford, UK, 1988, pp. 160–175.
[29] Y.T. Zhu, T.G. Langdon, Influence of grain size on deformation mechanisms: an
[3] M.E. Kassner, M.T. Pérez-Prado, Fundamentals of Creep in Metals and Alloys,
extension to nanocrystalline materials, Mater. Sci. Eng. A409 (2005) 234–242.
Elsevier, Oxford, UK, 2004, pp. 11–120.
[30] S. Ni, Y.B. Wang, X.Z. Liao, R.B. Figueiredo, H.Q. Li, S.P. Ringer, T.G. Langdon,
[4] W. Blum, J. Hausselt, G. König, Transient creep and recovery after stress reduction
Y.T. Zhu, The effect of dislocation density on the interactions between dislocations
during steady state creep of AlZn, Acta Metall. 24 (1976) 293–297.
and twin boundaries in nanocrystalline materials, Acta Mater. 60 (2012)
[5] I.S. Servi, N.J. Grant, Creep and stress rupture behaviour of aluminium as a function
3181–3189.
of purity, Trans. AIME 191 (1951) 909–916.
[31] T.G. Desai, P. Millett, D. Wolf, Is diffusion creep the cause for the inverse Hall–Petch
[6] H. Mecking, A. Styczynski, Y. Estrin, Steady state and transient plastic flow of
effect in nanocrystalline materials? Mater. Sci. Eng. A493 (2008) 41–47.
aluminum and aluminum alloys, in: P.O. Kettunen, T.K. Lepistö, M.E. Lehtonen
[32] R. Sandström, The role of cell structure during creep of cold worked copper, Mater.
(Eds.), Proc. Strength of Metals and Alloys-ICSMA8, Pergamon Press, Oxford, UK,
Sci. Eng. A674 (2016) 318–327.
1988, pp. 989–994.
[33] J.P. Hirth, J. Lothe, Theory of Dislocations, Krieger, Malabar, Florida, 1982.
[7] F.R.N. Nabarro, Do we have an acceptable model for power-law creep? Mater. Sci.
[34] U.F. Kocks, A.S. Argon, M.F. Ashby, Thermodinamics and kinetics of slip, Prog.
Eng. A 387–389 (2004) 659–664.
Mater. Sci. 15 (1979) 1.
[8] F.R.N. Nabarro, Creep in commercially pure metals, Acta Mater. 54 (2006)
[35] E. Nes, K. Marthinsen, Modeling the evolution in microstructure and properties
263–295.
during plastic deformation of f.c.c.-metals and alloys – an approach towards a
[9] A.A. Khamei, K. Dehghani, Effect of strain rate and temperature on the hot tensile
unified model, Mater. Sci. Eng. A322 (2002) 176–193.
deformation of severe plastic deformed 6061 aluminum alloy, Mater. Sci. Eng. A627
[36] H.D. Chandler, Effect of unloading time on interrupted creep in copper, Acta Metall.
(2015) 1–9.
Mater. 42 (1994) 2083–2087.
[10] Q. Zhang, W. Zhang, Y. Liu, Evaluation and mathematical modeling of asymmetric
[37] R. Sandström, Formation of a dislocation back stress during creep of copper at low
tensile and compressive creep in aluminium alloy ZL109, Mater. Sci. Eng. A628
temperatures, Mater. Sci. Eng. A700 (2017) 622–630.
(2016) 340–349.
[38] F. Roters, D. Raabe, G. Gottstein, Work hardening in heterogeneous alloys—a mi-
[11] P. Sundharshan Phani, W.C. Oliver, A direct comparison of high temperature na-
crostructural approach based on three internal state variables, Acta Mater. 48
noindentation creep and uniaxial creep measurements for commercial purity alu-
(2000) 4181–4189.
minium, Acta Mater. 111 (2016) 31–38.
[39] R. Wang, S. Wang, X. Wu, Edge dislocation core structures in FCC metals de-
[12] Y. Li, Z. Shi, J. Lin, Y.-L. Yang, Q. Rong, Extended application of a unified creep-
termined from ab initio calculations combined with the improved Peierls-Nabarro
aging constitutive model to multistep heat treatment of aluminium alloys, Mater.
equation, Phys. Scr. 83 (2011).
Des. 122 (2017) 422–432.
[40] M.E. Kassner, Taylor hardening in five-power-law creep of metals and class-M al-
[13] T. Subroto, A. Miroux, D.G. Eskin, L. Katgerman, Mater. Sci. Eng. A679 (2017)
loys, Acta Mater. 52 (2004) 1–9.
28–35.
[41] H. Mecking, Y. Estrin, The effect of vacancy generation on plastic deformation, Scr.
[14] Y. Xu, L. Zhan, L. Xu, M. Huang, Experimental research on creep aging behaviour of
Metall. 14 (1980) 815–819.
Al-Cu-Mg alloy with tensile and compressive stresses, Mater. Sci. Eng. A682 (2017)
[42] C.M. Young, S.L. Robinson, O.D. Sherby, Effect of subgrain size on the high tem-
54–62.
perature strength of polycrystalline aluminium as determined by constant strain
[15] L. Zuo, B. Ye, J. Feng, X. Kong, H. Jiang, W. Ding, Effect of Q-Al5Cu2Mg8Si6 phase
rate tests, Acta Metall. 23 (1975) 633–639.
mechanical properties of Al-Si-Cu-Mg alloy at elevated temperatures, Mater. Sci.
[43] F.R.N. Nabarro, The time constant of logarithmic creep and relaxation, Mater. Sci.
Eng. A693 (2017) 26–32.
Eng. A 309–310 (2001) 227–228.
[16] X. Jiang, Y. Zhang, D. Yi, H. Wang, X. Deng, B. Wang, Low-temperature creep be-
[44] F.A. Mohamed, T.G. Langdon, Deformation mechanism maps based on grain size,
haviour nd microstructural evolution of 8030 aluminum cables, Mater. Charact.
Metall. Trans. 5 (1974) 2339–2345.
130 (2017) 181–187.
[45] T.S. Lundy, J.F. Murdock, Diffusion of Al26 and Mn54 in aluminum, J. Appl. Phys. 33
[17] C. Li, M. Wan, X.-D. Wu, L. Huang, Constitutive equations in creep of 7B04 alu-
(1962) 1671–1673.
minum alloys, Mater. Sci. Eng. A527 (2010) 3623–3629.
[46] S.L. Robinson, O.D. Sherby, Activation energy for lattice self-diffusion in alumi-
[18] R. Sandström, J. Hallgren, The role of creep in stress strain curves for copper, J.
nium, Phys. Status Solidi A 1 (1970) K119–K122.
Nucl. Mater. 422 (2012) 51–57.
[47] L. Zhang, Y. Du, Q. Chen, I. Steinbach, B. Huang, Atomic mobilities and diffusivities
[19] R. Sandstrom, Basic model for primary and secondary creep in copper, Acta Mater.
in the fcc, L12 and B2 phases of the Ni-Al system, Int. J. Mater. Res. 101 (2010)
60 (2012) 314–322.
1461–1475.
[20] S. Vujic, R. Sandstrom, C. Sommitsch, Precipitation evolution and creep strength
[48] S. Straub, W. Blum, Does the “natural” third power law of steady state creep hold
modelling of 25Cr20NiNbN austenitic steel, Mater. High Temp. 32 (2015) 607–618.
for pure aluminun? Scr. Met. Mater. 24 (1990) 1837–1842.
[21] R. Sandström, Fundamental models for creep properties of steels and copper, Trans.
[49] J. Gubicza, N.Q. Chinh, T.G. Langdon, T. Ungár, Microstructure and strength of
Indian Inst. Met. 69 (2016) 197–202.
metals processed by severe plastic deformation, in: T.G. Langdon, Z. Horita,
[22] R. Sandström, Influence of phosphorus on the tensile stress strain curves in copper,
M.J. Zehetbauer, S.L. Semiatin, T.C. Lowe (Eds.), Proceedings Ultrafine Grained
J. Nucl. Mater. 470 (2016) 290–296.
Materials IV, TMS, Warrendale, PA, 2006, pp. 231–236.
[23] R. Sandström, H.C.M. Andersson, Creep in phosphorus alloyed copper during
[50] M.E. Kassner, M.E. McMahon, The dislocation microstructure in aluminum de-
power-law breakdown, J. Nucl. Mater. 372 (2008) 76–88.
formed to very-large steady state creep strains, Metall. Trans. 18A (1987) 835–846.
[24] T.G. Langdon, A unified approach to grain boundary sliding in creep and

349

You might also like