You are on page 1of 12

J Mech Behav Mater 2015; 24(3-4): 67–78

S. Tamimi*, A. Andrade-Campos and J. Pinho-da-Cruz

Modelling the Portevin-Le Chatelier effects in


aluminium alloys: a review
DOI 10.1515/jmbm-2015-0008 between solute atoms and mobile dislocations during
plastic deformation (e.g. [1, 2]). As opposed to DSA, in
static strain aging (SSA) (e.g. [3, 4]), the solute-dislocation
Abstract: Plastic deformation processes are among the interactions occur only after plastic deformation. DSA and
most demanding processes in manufacturing because SSA phenomena are known to bring loss in ductility in the
they lead to different microstructure features in the materi- metallic materials. In addition, in some cases, these may
als produced. Various dislocation patterns can be induced result in the heterogeneous plastic deformation and the
by plastic strain under different conditions. A serrated localisation of the plastic deformation in to slip-bands. As
yielding/jerky flow in some dilute alloys, such as alumin- a consequence of interactions between solute atoms and
ium-magnesium alloys during plastic deformation, is a dislocations, there exist two kinds of propagative patterns
well-known phenomenon under certain regimes of strain during plastic deformation: the Lüders bands and the
rate and temperature, as reported in a significant number ­Portevin-Le Chatelier (PLC) bands, which are caused by
of works. The serrated features in these materials reflect SSA and DSA, respectively. Both features are undesirable
the so-called Portevin-Le Chatelier effects. These unde- and should be avoided in manufacturing processes.
sirable effects are due to the interaction between solute Piobert and Lüders first reported the Lüders phenom-
atoms and mobile dislocation during the plastic deforma- enon in low carbon steels [5, 6]. They have shown that a
tion, which is known as dynamic strain ageing. There are Lüders band travels along the specimen during the tensile
a significant number of theoretical and numerical inves- test at a constant cross-head speed. The observed angle
tigations that have focused on describing the serrated of front line is approximately 50° from the sample axis,
behaviours of these materials during plastic deformation. which is the most common slip direction of  < 111 >  in BCC
Hence, the fundamental objective of this paper is to pro- materials. Figure 1 schematically illustrates the forma-
vide a general review of different constitutive modelling in tion of Lüders bands. As can be seen, after a drop in stress
regards this feature. The typical material models and new from the upper yield point to the lower yield point due to
constitutive models describing this feature are presented. the elastic relaxation of the rest of the sample, the plastic
In addition, applications of the models are provided along front of the Lüders bands moves at a constant lower yield
with their respective advantages and disadvantages. stress level.
The nucleation and movement of the Lüders band in
Keywords: aluminium-magnesium alloys; constitutive these materials at room temperature are associated with
modelling; dynamic strain ageing; Portevin-Le Chatelier the collective unpinning of the dislocations that are ini-
effects. tially locked at the upper yield point. The time required
for alloying atoms to diffuse into the core dislocations and
lock them is longer than the total time of the load apply-

1 Introduction ing; hence, no further repinning can occur. The stress


required to push the dislocations is lower than the upper
yield point. Consequently, the lower yield stress corre-
Dynamic strain aging (DSA) is a well-known phenomenon
sponds to the propagation of the Lüders band (Figure 1).
occurring in many alloys as a result of the interaction
However, SSA (via increasing the temperature for a suffi-
cient time) can lead to the diffusion of the alloying atoms
*Corresponding author: S. Tamimi, Center for Mechanical into the dislocations’ core, which causes the appearance
Technology and Automation, TEMA, Departmento de Engenharia of Lüders bands in the next plastic deformation.
Mecânica, Universidade de Aveiro, Portugal,
Meanwhile, the serration feature of PLC was first
e-mail: saeed.tamimi@ua.pt; saeed.tamimi@gmail.com
A. Andrade-Campos and J. Pinho-da-Cruz: Center for Mechanical
reported by Savart [8]. However, the details of this obser-
Technology and Automation, TEMA, Departmento de Engenharia vation were documented by Le Chatelier [9] for steel
Mecânica, Universidade de Aveiro, Portugal between 80°C and 250°C, and by Portevin and Le Chatelier
68      S. Tamimi et al.: Modelling the Portevin-LeChatelier effects

Upper yield
point

Yield elongation

Lower yield
point
Load

Figure 2: Diffusion of alloying atoms towards mobile dislocations


Lüders (DSA).
band

Unyielded metal

Elongation

Figure 1: A schematic illustration of the propagation of Lüders


bands in a stress-strain curve [7].

[10] for aluminium (Al) alloys at room temperature. There-


fore, the serration phenomenon was named Portevin-Le
Chatelier effect (PLC effect). Once the sample containing
the alloying elements is taken under a plastic load, a dis-
continuous behaviour occurs consisting of stress drops Figure 3: Range of strain rate-temperature for the occurrence of
and reloading sequences. It should be noted that each serrated follow of the PLC in the AA-5083 alloy [16].
load drop is related to the nucleation of a localised defor-
mation band that occurs within a range of strain rates and
temperatures. Later, many studies on the PLC effect, as
one type of defect in various metallic materials, have been
carried out (e.g. [11–13]). In a microscopic point of view,
dislocation distorts the surrounding lattice, thus causing
an elastic stress field that acts on the neighbouring solute
atoms. During plastic deformation, when mobile disloca-
tions are being blocked by dislocations tangles, a cloud of
solute is formed around it in order to reduce the distortion
in the crystal. This occurs with solute segregations by pre-
ferred pipe diffusion and the mobile dislocation is effec-
tively pinned (Figure 2). With the aid of effective stress,
obstacles can be overcome by thermally activated disloca-
tion motion; this unpinning process of dislocation, com-
Figure 4: Flow stress dependence of the strain rate for an Al-Mg
bined with the long range dislocation interactions, causes alloy at room temperature [17].
a stress drop on macroscopic stress curve [14].
The conditions of the occurrence of the PLC effect in
terms of strain, strain rate and temperature have been Furthermore, Figure 4 illustrates a nonlinear feature
taken into consideration by a considerable number of of stress vs. strain rate, namely, the existence of a range
authors [e.g. 3]. The PLC phenomena occur in a specific of ε , where the flow stress exhibits a negative sensitivity
range of temperature and strain rate in Al alloys [15, to strain rate. This results in an N-shaped dependence of
16]. Figure 3 illustrates a typical range of strain rate-­ stress on ε . Such behaviour is stress rate- and tempera-
temperature for a jerky flow of the PLC phenomena in an ture-dependent. Here, the arrows indicate the cyclic orbit
­Al-magnesium (Mg) alloy, which has been obtained using corresponding to unstable plastic flow [18, 19].
the experimental measurements and using an Arrhenius In polycrystals deformed with a constant imposed
type equation [16]. strain rate falling in the negative strain rate sensitivity
S. Tamimi et al.: Modelling the Portevin-LeChatelier effects      69

range, three types of serrated curves are commonly distin- σ = hε + F ( ε ),  (1)


guished [20]. These correspond to different dynamics of
where h is the strain hardening coefficient that is assumed
strain localisation evolving from the so-called type C to type
constant for simplicity, and F ( ε ) is a function exhibiting
B and to type A when the strain rate is increased or when the
a range of negative strain rate sensitivity. This expres-
temperature is reduced: type A serrations are associated with
sion results in the N-shape behaviour shown in Figure 4.
the repetitive continuous propagation of deformation bands
Numerical modelling studies of the PLC effect have been
nucleated at a certain point, often at a specimen end, and
reported in previous works. This approach, previously
moving along the specimen; type B serrations correspond to
­suggested by McCormick [21, 22], Estrin and ­McCormick
a hopping propagation of localised bands in the axial direc-
[23] and M ­ cCormick et  al. [24], was later revised by
tion of the sample; and type C serrations occur when local-
­McCormick and Ling [25] to simulate the serration behav-
ised deformation bands appear in a spatially non-correlated
iour of the materials. The significant aspects of the PLC
manner. As in the case of type B, each stress drop reflects the
effect (e.g. PLC instabilities and their critical plastic strain)
formation of a single localised band. As noted previously,
can be successfully predicted using this micromechanical
the DSA phenomenon occurs for a certain range of tempera-
constitutive model.
ture and strain rate; in some cases, a critical strain level has
The features of the PLC effect can be simulated by
to be reached for serrated yielding to take place.
discretising a tensile sample in a number of small sec-
In general, DSA and the PLC effects have been the
tions. The constitutive equation in each single section is
subject of numerous experimental and theoretical studies.
simultaneously solved while taking into account the con-
In the past decades, numerical modelling using the finite
straint of constant crosshead velocity. Here, according to
element (FE) method has also been used to investigate
the demonstrated works [21, 22, 24, 26], the DSA influence
the PLC effect. These studies aim to represent the spa-
on the following stress is taken into account by adding a
tio-temporal evolution of the local mechanical variables
term to Eq. (1). This additive term depends on the solute
using appropriate constitutive laws. There are various
concentration at the blocked mobile dislocations. Hence,
types of material models that attempt to describe the DSA
the applied stress σ is given as:
and the PLC effects. In the present work, a review of the
material models describing the spatio-temporal aspects  ε 
σ = hε + Si ln  + βC , (2)
of the PLC phenomena is presented. In models based on  ε 0  
the macroscopic description of deformation bands, the
negative strain rate sensitivity is explicitly defined, and where Si is the instantaneous strain rate sensitivity of the
serrations are obtained from strain rate jumps. The other flow stress in the absence of solute effects; the parameter β
type of materials models rely on a microscopic description represents the strength of dislocation pinning by solutes;
of the DSA based on the internal variable of ageing time. and C indicates the average solute concentration around
Here, the negative strain rate sensitivity and serrations are dislocations. The term of βC depends on the average
implicit consequences of constitutive equations. Addition- waiting time tw of dislocations at localised obstacles (or
ally, some more sophisticated models have been reported dislocation tangles), which is inversely proportional to
in previous works based on thermodynamic features. In the plastic strain rate, ε . The average waiting time can be
this work, the application of these material models and written as:
their capability to explain the serration behaviours of the

materials though plastic deformation are discussed. tw = , (3)
ε 

where Ω is a strain increment produced when all arrested


dislocations overcome localised obstacles and advance to
2 Micromechanical constitutive the next pinned configuration. The variation of C with the

formulation waiting time can be described [27–29] by:

  C  t  p 
C = Cm  1-exp  - 0  w    . (4)
2.1 The fundamental approach
 
 Cm  τ    

A simple relation between the flow stress (σ) the plastic In this equation, Cm is a saturation value of the solute
strain (ε) and the plastic strain rate ( ε ), first suggested by concentration at dislocations, and C0 is the nominal solute
Penning [18], can be written as: concentration in the crystal bulk, in the following form:
70      S. Tamimi et al.: Modelling the Portevin-LeChatelier effects

W  in Figure 4). In other words, in this range of strain rate, the


Cm = C0 exp  . (5)
 kBT   strain rate sensitivity is negative, resulting in the N-shaped
curve of Figure 4. The influence of temperature on the flow
The parameter W represents the binding energy
stress-strain rate curve plotted in this figure can also be
between a solute atom and a dislocation, and kB is the
perused. Regarding Eq. (5), with the rising temperature,
Boltzmann constant. It is noticeable that the binding
βCm decreases between two ascending branches of the
energy W influences the parameter β in Eq. (2), whose
w curve shown in Figure 4. Additionally, an increase of T
estimated order of magnitude is 3 . The parameter b is results in a decrease of the slope of the curve, which is
b
the Burgers vector of the dislocation. Here, Eq. (4) involves due to Si being inversely proportional to temperature in
the characteristic time τ, which is related to solute diffu- these ranges of strain rates. Based on Eq.  (8), the posi-
sion to dislocations. This characteristic time depends on tions of ε 1 and ε 2 shift to higher strain rates as ε ∗ grows
the temperature calculated as follows: with ­temperature; simultaneously, their differences also
increase (Figure 4). This means that the range of nega-
 Q 
τ = τ0 exp  , (6) tive strain rate sensitivity becomes smoother in the curve
 kBT   of σ ( ε ) vs. ε . It can also occur when the nominal solute
concentration C0 is reduced.
where Q is the activation energy for solute migration, and
In general, the mentioned constitutive parameters
the pre-exponential factor τ0 is constant.
should be experimentally determined by measurements.
In Eq. (4), regarding the classical Cottrell-Bilby theory
A comparison of the results obtained by the numerical
[30], the exponent p is 2/3. However, a smaller value of 1/3
study of the PLC effect based on the transient constitu-
has also been reported in [31]. The solute concentration
tive model for DSA and experimental observation in an Al
given by Eq. (4) obeys the Cottrell-Bilby type power-law
alloy indicates that the numerical simulations are in good
for small waiting times and saturates at C, for large tw. Sub-
agreement with experimental measurements [32, 33].
stituting Eqs. (3–6) in Eq. (2), σ ( ε ) can be written as:
Moreover, Jiang et al. [34] produced the elastic shrinkage
   ε ∗  p   outside an avalanche-like PLC deformation band using
 ε 
σ ( ε ) = Si ln  + βCm  1-exp  -     , (7) this model with modified boundary conditions. In that
 ε 0    
   ε     work, numerical simulations were carried out for three
types of PLC effects at different constant applied strain
where rates in a conventional tensile test, which were found to
be in good agreement with the experimental observations
 Ω  ( Q +W / p ) 
ε ∗ =   exp  - . (8) (Figure 5). They also investigated the dynamic interac-
 0
τ  K BT   tions between dislocation and diffusing solute atoms by
the constitutive equations.
The non-monotonic strain rate-dependent behaviour
Furthermore, Lebyodkin et al. [35, 36] investigated the
seen in Figure 4 can be explained by Eq. (7). In the large
statistics of the stress drops associated with the PLC effect
strain rate limit (i.e. ε 2 < ε in Figure 4), the last term is neg-
in an AI-Mg alloy (AA-5xxx) both experimentally and theo-
ligible and the logarithmic dependence, which is charac-
retically for single crystals and polycrystals. They used the
terised by a solute-free material, is recovered. However, for
constitutive models for dynamically strain ageing materi-
small strain rates (i.e. ε < ε 1 in Figure 4) when the waiting
als discussed above [21–24, 26] to describe the temporal
times are long enough for the solute atoms around dis-
behaviour associated with the PLC effect. Additionally,
locations reach saturation concentration, Eq. (7) can be
the localised deformation band patterns characteristic of
written as:
the PLC effect was simulated using this model. A principal
 ε  feature of the model used in their work is a dynamisation
σ = Si ln  + βCm . (9)
 ε 0  of the N-shaped curve proposed by Kubin, where it had

been shown that the elastic loading occurring during the
It is obvious that the low strain rate range of the flow waiting time leads to a deformation. The authors of that
 ε  work also indicated that the proposed model provides an
stress vs. ln  curve is simply vertically displaced by
 ε 0  adequate description of both the statistics of stress discon-
βCm, with respect to the high strain rate ranges. Hence, tinuities as well as the spatial features of the PLC effect.
there is a transition region between the two strain rate The model is able to reproduce the important features of
limits and the stress decreases by strain rate (i.e. ε 1 < ε < ε 2 the material behaviour in the PLC regime (e.g. different
S. Tamimi et al.: Modelling the Portevin-LeChatelier effects      71

A 207 B C 207
210
Simulated
Simulated Simulated
198
Stress (MPa)

Stress (MPa)

Stress (MPa)
198
195

189
180 189
Experiment
180 Experiment
Experiment
165 180

1600 1700 1800 1900 10 12 14 16 18 195 210 225 240


Time (s) Time (s) Time (s)

Figure 5: Experimental and simulated observation of three PLC types in tensile test at different applied strain rates.
(A) Type A, 5 E-3 s-1; (B) type B, 5E-4 s-1; and (C) type C, 5E-5 s-1 [34].

A B 30

60
Number of stress drops
Number of stress drops

20
40

10
20

0 0
0 1 2 3 0 2 4 6
Stress drop magnitude Stress drop magnitude

Figure 6: Simulated statistical distribution of the stress drops vs. different stress drop magnitudes.
(A) 1E-4 s-1 strain rate and (B) 4E-4 s-1 strain rate [35].

types of serrations, the variation of the character of the ageing time and a hardening term for three-dimensional
stress drop statistics with changing deformation condi- finite element analyses. The shape of PLC bands in
tions, and spatial deformation patterns). Figure 6 indi- AA-6xxx alloys was then investigated through FE analy-
cates a statistical observation of stress drops for various sis using this model. In order to conduct 3D simulations
stress drop magnitudes in an Al-Mg alloy for two different of the serration features and PLC effect, the constitutive
strain rates. These studies also indicated that a combi- model suggested by McCormick was reformulated for
nation of a local constitutive law capturing the temporal multiaxial loading [38]. Here, the increment of the total
aspects of the PLC effect with a certain type of spatial cou- strain rate tensor (i.e. ∆εtij ) is defined by the sum of the
pling can result in a simple and efficient model describing elastic (i.e. ∆εije ) and plastic (i.e. ∆εijp ) strain tensor incre-
the spatio-temporal behaviour of a dynamically strained ments given by:
ageing material exhibiting the PLC effect.
∆εtij = ∆εije + ∆εijp . (10)

According to Hooke’s law, the elastic part of the equa-


2.2 Developed approach and its application tion can be written in terms of the stress tensor increment
Δσij:
Zhang et al. [37] revised the model presented in the pre-
vious section. An elasto-viscoplastic constitutive law 1+ υ  υ 
∆εije = Cijkl ∆σ kl , where Cijkl = δ δ - δδ . (11)
accounts for DSA through a new internal variable, the E  ik jl 1 + υ ij kl  
72      S. Tamimi et al.: Modelling the Portevin-LeChatelier effects

Here, υ and E are the Poisson’s ratio and the Young’s the rate of saturation of solute atoms around dislocations.
modulus, respectively. Meanwhile, the plastic strain incre- The parameters α and n are constants pertaining to the
ment tensor, ∆εijp , can be written as: measurements of Ling and McCormick [33].
An important assumption regarding the work of
∆εijp = ε ijp ∆t , (12)
McCormick et  al. [24] concerns the ageing and waiting
times. The effective ageing time (i.e. ta) is not equal to
where Δt is the time increment. Assuming a plastically iso-
the average waiting time when a dislocation is arrested
tropic material, the plastic strain rate tensor ε ijp is written
at localised obstacles (i.e. tw). Hence, according to
in terms of the stress tensor σij or the deviatoric stress
­McCormick’s work, the effective ageing time (ta) is consid-
tensor of sij. The corresponding equivalent von Mises
ered to relax towards tw with time t. The time derivative of
quantities can be defined as:
ta is obtained by the following relationship:
2 p p 2
∆εijp = ε ε and σ v = ss . (13) dta ta
3 ij ij 3 ij ij  = 1- . (19)
dt tw 
Regarding the Prandtl-Reuss equation, the plastic
Regarding the previously mentioned works, the
strain rate tensor is given by:
waiting time tw is related to the equivalent plastic strain

3 ε p rate via Eq. (3) (i.e. tw = ). Here, the strain dependence
ε ijp = s. (14) ε p
2 σ υ ij 
of Ω can be obtained using the dislocation model [21, 22,
24]. This developed model gives a realistic description
Estrin and McCormick [23], McCormick et al. [24] and
of the spatio-temporal dynamic behaviours of the three
Estrin [38] suggested the equivalent plastic strain rate of
types of PLC effects by including the spatial coupling.
ε p as follows:
Kok et al. [39, 40] employed a crystal plasticity model
 σ -σ  embedded in an FE framework to study the PLC effect
ε p = ε 0 exp  v d -P1Cs′ . (15)
 S   in constant cross-head velocity-controlled tensile tests.

They used a constitutive law based on the Zhang model
This equation is valid only for the non-negativity of and introduced the model into a crystal plasticity law.
the exponent. The stress σd, representing the strain hard- The constitutive model of Zhang was later modified by
ening effect, is related to the dislocation density evolu- Graff et al. [41, 42] through the introduction of a thermally
tion, which is given by: activated elasto-viscoplastic law. This formulation, for
example, was used by Belotteau et al. [43] to predict both
  εp  
σ d = d1 + d2  1-exp  -   . (16) static (i.e. in Lüders bands) and dynamic strain ageing
  d3    (i.e. the PLC effects) of steel over a large range of tempera-
tures and strain rates. Maziere et  al. [44] used the same
As explained, the strain rate sensitivity (i.e. S) depends
constitutive model to simulate round smooth and notched
on stress and subsequently strain. This parameter has
specimens of a Ni based superalloy. Hopperstad et al. [45]
been obtained by experimental measurements for an Al
used an anisotropic elasto-viscoplastic model includ-
alloy in the work of Ling and McCormick [33] using the fol-
ing the McCormick model for DSA. In the works of Graff
lowing equation:
and Belotteau, the PLC effects in notched and cracked
1
specimens were numerically investigated with 2D models.
S = s1 + s2 ( ε p ) 2 . (17)
Accordingly, Wenman and Chard-Tuckey [46] applied a 3D
model of a notched specimen to study the strain localisa-
In Eqs. (15–17), ε 0 , P1, d1, d2, d3, s1, and s2 are con-
tion stemming from static ageing including residual stress
stants (the strain rate dependence of d1, d2 and d3 [38] is
and plastic strain obtained by preloading in compression.
neglected), and εp is the (von Mises) equivalent plastic
Benallal et  al. [47] used a phenomenological elasto-
strain. The non-dimensional solute concentration corre-
viscoplastic model based on the work of Penning [18]
sponds to:
to investigate the serration features in tensile tests with
Cs = ( 1-exp( - P2 ( ε p ) α ( tan )))Cm . (18) smooth axisymmetric samples (with and without U-notch)
at various strain rates. The simulated results of ­Benallal’s
The maximal overconcentration is Cm when the effec- work are in good agreement with the experimental obser-
tive ageing time ta tends to infinity, and P2 characterises vations reported in a previous work [48]. In order to
S. Tamimi et al.: Modelling the Portevin-LeChatelier effects      73

simulate the negative strain rate sensitivity being able to Regarding Eq. (24), the ageing time depends on ε P ,
predict strain localisation bands (i.e. PLC bands), differ- which has a maximum point, i.e. it increases until it
w
ent strain hardening laws have been used in this model for reaches the maximum value of the waiting time P and

different strain rates.
then decreases to zero. The PLC effects and strain localisa-
Manach et al. [49] used McCormick’s model to study
tion features appear after this point. The hardening caused
the PLC effects in an Al-Mg alloy. They used shear tests
by DSA depends on the Cs, which increases with ageing
at room temperature in order to compare the predicted
time from Cs = 0 (the unpinned situation) to Cs = 1 (the fully
load drops and kinetics of the bands with the experimen-
pinned situation) or inversely when a PLC band crosses.
tal data. They also considered anisotropy and kinematic
An extra stress P1 [shown in Eq. (20)] is required to switch
hardening by using an anisotropic yield criterion and a
between both situations while the kinetics of the pinning
nonlinear kinematic hardening given by:
process can be controlled by the parameter n in Eq. (18).
f ( σ , ε P , ta ) = σ v -R-P1Cs ( ε P , ta ), (20) Manach et  al. [49] showed that the McCormick model is
capable of predicting the PLC features. The amplitude is
where σv is the von Mises equivalent stress and R is the similar to those that have been measured experimentally.
isotropic work hardening defined as: Additionally, their study indicated the different types of
PLC (i.e. A and B PLC bands) as a function of the strain
R = σ y + Q( 1-exp( -b ε P )). (21)
rate with respect to the stress drop distribution. Strain,
time and strain rate jumps are also in good agreement
The term P1Cs(ε̅P, ta) in Eq. (20) is the extra-harden-
with experimental observations (Figure 7).
ing induced by strain ageing suggested by Zhang et  al.
Furthermore, it has been shown that the kinematics
[37]. Regarding Eq. (18), Cs is a function of both internal
of the bands using the McCormick model is in agreement
variables of the model, the equivalent plastic strain and
with experimental observations. In the shear stress direc-
the ageing time ta. The plastic strain follows a flow rule
tion and its perpendicular directions, the speed of growth
derived from a viscoplastic potential Ω given by:
of the PLC bands is faster. This leads to the production of
 f+ different bands instead of a well-defined band as observed
Ω( f ) = K ε 0 cosh  . (22) in the experimental results. Additionally, Manach et  al.
K 
[49] used the following model of Johnson-Cook [51, 52] to
In this equation, ε 0 represents the strain rate sensi- study the PLC effect in Al-Mg alloy due to its simplicity:
tivity coefficient, and K is a weighting coefficient of the
  ε  
viscous part of the stress, with f + being the positive part n
σ = [ A + B ε P ]  1 + D ln p   [ 1-T m ]. (25)
of the yield criterion. The equivalent plastic strain rate   ε 0   
tensor is defined by the plastic conservation principle in
the form of:
∂Ω ∂f (23)
ε P = = Ω′(f) .
∂σ ∂σ 

Additionally the ageing time increment, Δta, can be


calculated using the increment of the plastic strain, w.
This can be produced when all locked dislocations escape
from their obstacles. The ageing time increment, Δta, can
be obtained using the following equation:

ta  P
∆ta = 1- ε , (24)
w 

where the increment of the plastic strain, w, leads the


incremental strain resulting from the jump of unpinned
Figure 7: A comparison between the predicted and experimentally
dislocations, and has a significant influence on the tran-
measured shear stress-shear strain curves.
sition process between pinned and unpinned states. These values are for an Al alloy using McCormick model for three
According to Wang et al. [50], decreasing this value may different strain rates v1, v10 and V100 representing γ = 1.2 e-3,
promote strain localisation. γ = 1.2 e-2 and γ = 1.25 e-1 s-1 , respectively [49].
74      S. Tamimi et al.: Modelling the Portevin-LeChatelier effects

The parameters A, B, D, n, m and ε 0 are material constants,  -mQV  -m


Ω( ε , T ) = Ω0 + C exp   ε , (30)
and T is the non-dimensional temperature. The term  RT  
n
[ A + Bε P ] represents the plastic strain hardening (i.e.
  ε P  
Ludwik’s equation),  1 + D ln   represents the strain u( ρ ) =U 0 ρ . (31)
  ε 0  
rate hardening and the third term of [1-T m] indicates the Palumbo et  al. [55] also used the Bergstrom consti-
thermal softening. In this relation, the PLC effect occurs tutive model to simulate warm deep drawing of AA-5xxx
when D is negative. For the simulation of the behaviour using coupled thermomechanical FE analysis. Deep
of these Al alloys, the Voce-type law was employed in Eq. drawing experiments were performed with a cooled
(25). Hence, if the thermal effect is ignored, the following punch and heated dies. Axisymmetric simulations were
relation is used: performed with various coefficients of friction and punch
speeds. The calculated punch force had reasonable agree-
  ε P   ment with the experiments.
σ = [ σ y + Q( 1-exp( -b ε P ))]  1 + D ln    . (26)
  ε 0    Kurukuri et  al. developed an improved physically-
based constitutive model known as the NES model [56].
Generally, the work of Manach et  al. [49] indicates The NES model improves on the Bergstrom model by
that the McCormick model is capable of reproducing DSA incorporating a multiparameter description of micro-
in shear test at room temperature, and that the kinemat- structure. The dislocations are assumed to be stored in
ics and morphology of the localised plastic bands are in a finite cells, and both the dislocation density and cell size
satisfying agreement with experimental results. are tracked. The cells are bounded by finite walls, and at
The Bergstrom and Hallen model is a physically-moti- large strains, the walls collapse due to dynamic recovery.
vated material model based on dislocation densities [53]. The NES model has improved strain rate dependence over
The model has been used in thermomechanical coupled the Bergstrom model and can more accurately predict
processes of warm forming of Al by Van den Boogaard and localisation. Kuruki recommended investigating friction
Huetink [54] and Palumbo et  al. [55]. The model decom- in detail to improve warm forming simulations. The NES
poses the flow stress in to three components: a strain and model requires 30 independent parameters to be fully
strain rate independent stress σ0, a dynamic stress that defined.
depends on strain rate and temperature and a work hard-
ening component. The flow stress was decomposed into a
strain and strain rate independent stress σ0 and σw, which
incorporate work hardening as shown in the following
3 Thermodynamic constitutive
relation: formulation
σ f = σ 0 ( T ) + σ w ( ρ, T ). (27)
In the works of Rizzi and Hahner [57, 58] a model based
The work hardening-dependent stress, σw, is a func- on a thermodynamical approach was proposed. The
tion of the dislocation density ρ given by: model equations were developed in a one-dimensional
context and processes were taken as isothermal. At the
σ w = αG( T ) b ρ . (28)
starting point, Rizzi’s model assumes that plastic flow is
brought by thermally-activated dislocation glide, similar
Here, G and b are the shear modulus and the Burgers
to the other PLC models (e.g. McCormick’s model). Hence,
vector, respectively, and α is a scaling parameter. The
according to the constitutive flow rule of the Arrhenius
evolution of dislocation density is expressed as the com-
type relation, the strain rate is given as:
petition between dislocation storage and recovery by
remobilisation and annihilation expressed as:  G + ∆G σ ext -σ int 
ε t = ϑΩ exp  - 0 + , (32)
 kT S0  

=U ( ρ )-Ω( ε , T ) ρ, (29)
dε 
where ϑ and Ω are physical parameters representing the
where U represents the storage of mobile dislocations, and attempt frequency of thermal activation and the elemen-
Ω is the dynamic recovery by remobilisation and annihila- tary plastic strain corresponding to the activation of all
tion. The amounts of Ω and U determine the shape of the mobile dislocation segments, respectively. In addition, k
hardening curve, which can be written as: represents Boltzmann constant and T is temperature.
S. Tamimi et al.: Modelling the Portevin-LeChatelier effects      75

The salient feature of the present PLC model is that According to Eq. (32), it can be pointed out that a
the Gibbs free activation enthalpy G defining this energy constitutive equation of that type should be applied to
barrier is considered a dynamic internal variable, which is each single active dislocation slip system. Accordingly, σ
subjected to the dynamic strain ageing. It can be written can be considered as the shear stresses resolved on the
as: respective slip systems and the same for strains, i.e. t as
the corresponding plastic shear strain rates γ t . In order
G = G0 + ∆G , (33)
to simplify this process, the resolved stresses and strains
where G0 is the basic activation enthalpy in the absence are rather replaced by the equivalent quantities. Another
of DSA, and ΔG is an additional contribution related to relevant point is that yielding, as represented here, is
DSA, which is in proportion to the solute concentration smooth without any discontinuous elastic to plastic tran-
accumulated at the glide dislocation segments. In Eq. sition ruled by a yield function. However, in reality, yield-
(32), the equivalent stress (σeff ) does not coincide with the ing turns out to be quite sharp, with a well-defined yield
externally applied stress, σext, given that only the excess point due to the exponential term. Furthermore, aiming
of external stress over the internal stress σint resulting only at simulating tests with continuous plastic flow,
from dislocations as defects, is effective in driving dislo- material branching between plastic and unloading elastic
cation motion. Both σeff and σext are dynamic quantities responses is not taken into account. Notice that this does
that depend on DSA (due to ΔG), on plastic strain and on not impede local unloading associated to the plastic insta-
plastic strain rate. However, σint depends only on strain bilities in the PLC range.
hardening given by: Rizzi and Hahner took an additional activation
enthalpy G related to the kinetics of DSA as the main
σ eff ( ε, ε t , ∆G ) = σ ext ( ε, ε t , ∆G ) -σ int ( ε ). (34)
feature of his model. This is a (dynamic) internal variable
The material parameter S0 in Eq. (32) defines the of the model. The term ΔG is calculated as:
instantaneous strain rate sensitivity (SRS) of the flow
ε t
stress with respect to processes that can be obtained by ( ∆G )t = η( ∆G∞ -∆G )- ∆G . (38)
Ω 
inverting the Arrhenius law [Eq. (32)]:
Through plastic deformation, when ageing occurs
G0  ε  ∆G and dislocations are arrested at dislocation tangles, solute
σ ext ( ε, ε t , ∆G ) = σ int ( ε ) + S0 + S0 ln t  + S0 , (35)
kT  ϑΩ  kT  atoms diffuse towards the dislocations. Hence, the con-
centration of solute atoms on the dislocations increases,
in which thus leading to an increase of the activation enthalpy ΔG.
Afterwards, when dislocations are released by thermal
∂σ ext
S0 = . (36) activation, these escape from their solute clouds. Thus,
∂ ln ε t ε , ∆G  the solute concentration, and hence the additional activa-
tion enthalpy, is reset to zero at a rate controlled by the
The instantaneous SRS, which relates as S0 = kT/V to
plastic strain rate. DSA and then the PLC feature can be
the activation volume V of the thermally-activated dis-
explained by the interplay of these two mechanisms.
location glide, is positive, that is, the flow stress always
In Eq. (38), when ε t = 0 (during static ageing),
increases, almost instantaneously, with a sudden increase
ΔG = ΔG∞(1-exp(-ηt)) where the parameter η presents the
in strain rate. This does not imply, however, that the SRS
ageing rate. For t >  > η-1 the additional activation enthalpy
actually governing the stability of plastic flow does not
approaches the maximum value ΔG∞ and for t >  > η-1 the
become negative. In fact, a change in strain rate is fol-
solute concentration increases linearly resulting in
lowed by a stress transient towards a new steady state,
ΔG≈ΔG∞ηt2. The term of unpinning in Eq. (39) brings
which may be above or below the extrapolated stress-
about the decrease of ΔG due to thermal activation, which
strain curve before the strain-rate jump. This is accounted
occurs at a specific rate. The characteristic time scale of
for by introducing the so-called asymptotic SRS S∞, which Ω
assumes negative values in the PLC range. Formally, S∞ is this process is the waiting time t w = spent by disloca-
ε t
defined as the SRS of the flow stress as observed when DSA tions at the obstacles. This part of Eq. (38) is non-linear,
processes have relaxed to a new steady state expressed as which can be attributed to the dependence of ε t on ΔG.

The ratio of two time scales (i.e. η-1 and ) governs the
∂σ ext ∂σ d∆G  1 d∆G  ε t
S∞ = = S0 + ext = S0  1 + . (37) competition of the two processes and the resulting kinet-
∂ ln ε t ε ∂∆G ε dlnε t  kT dlnε t  
ics of DSA.
76      S. Tamimi et al.: Modelling the Portevin-LeChatelier effects

In a steady-state condition, where ΔG = 0 in Eq. (38), it and


can be written as follows:
 ∆G 
( ∆G )t = η( ∆G∞ -∆G )-ηf exp  - ∆G . (46)
 ∆G∞   kT  
∆Gs =  . (39)
 
 1 + εt  As seen in previous relations, the non-dimensional
 Ωη 
 additional activation enthalpy variable as defined by the
d∆G ∆G ∆G
Regarding Eq. (37) and with ΔG = ΔGs, the above energy ratio g = is employed. Considering g0 = 0 ,
dlnε t kT kT
relation becomes:
∆Gs ∆G∞
gs = and g ∞ = , the constitutive equations forming
 ε t / Ωη  kT kT
S∞ = S0  1-g ∞  , where g ∞ = kT . (40)
 1 + ε t / ( Ωη ) 2  the base of the present model of the PLC effect are now

expressed as:
Hence, SRS gets negative when:
f = σ f -θ exp( - g ) f 2 , (47)
g ∞ -2- g ∞ ( g ∞ -4 ) ε t g ∞ -2 + g ∞ ( g ∞ -4 )
≡ gl < < gu ≡ . (41)
2 2 g = g ∞ -g -f exp( - g ) g . (48)
Ωη 

Here, SRS becomes zero at g∞ = 4 when the two char- Here, the dimensionless stress rate σ and hardening
acteristic time scales match each other (i.e. ε t = Ωη ). coefficient θ are scaled parameters. These parameters are
However, for the values of g∞ > 4, Eq. (10) defines the range linked to σext, t and h using the relations below:
of plastic strain rates that give rise to a negative SRS. In
σ ext , t
this relation, g1 and gu are the lower and upper bounds, σ = and (49)
ηS0 
respectively, which are ruled by the saturation DSA
enthalpy ΔG∞. It is noticeable that the steady-state defor- hΩ
mation condition is σ ext , t = hε t , so that Eq. (41) also pro- θ= . (50)
S0 
σ
vides the stress rates g l < ext , t < g u entering the regime of When the non-dimensional hardening coefficient of θ
Ωηh
a negative SRS. is smaller than 1 (i.e. weak hardening), f is the slow vari-
With the goal of simplifying the equations, the non- able of the model, while g represents its fast variable in
dimensional driving force of f can be written as a function Eqs. (47) and (48).
of σeff: In constitutive Eqs. (47) and (48), the incorporation
of two available intrinsic time-scale parameters [one
σ 
f ≡ f0 exp  eff  , (42) associated with ageing – unit time scale characteristic
 S0   of g shown in Eq. (47) and the other one associated with
strain hardening characteristic time θ governing f, Eq.
 ϑ  G  (48)] constitutes a major difference between Rizzi’s model
where the factor of f0  f0 = exp  0   embeds the fre-
 η  kT   and Penning’s type models. The ratio between η-1 and
quency of ϑ. Hence we have: Ω
tw = governs the appearance of the PLC effect, while θ
ε t
 ∆G  determines the shape of the limit cycle in the weak hard-
ε t = ηΩ exp  - f. (43)
 kT   ening regime. Setting g = 0 in the Rizzi model may lead
to the recovery of the other previous model: one recov-
Taking the evolution equation of f by time differentia-
ers the flow stress rule Eq. (35), with ΔG replaced by its
tion, we have:
steady-state value ΔGs, Eq. (39). Given that the strain-rate
σ hε  dependence of this flow stress is “N-shaped” (provided
ft =  ext , t - t  f . (44) that ΔG∞ > 4kT), this is formally equivalent to Penning’s
 S0 S0  
model (Figure 4).
Replacing relation (44) in Eqs. (45) and (39), the dif- This model was used in the work of Hahner et al. [59]
ferential equations in f and ΔG become: to study the PLC band properties in a Cu-Al alloy. The
simulated results were compared with those obtained
σ ext , t hηΩ  ∆G  2
ft = f- exp  - f , (45) by experiments in Figure 8. Here, in both simulated
S0 S0  kT   and experimentally measured curves, different types of
S. Tamimi et al.: Modelling the Portevin-LeChatelier effects      77

A 100 55 are in good agreement with the experimental observa-


53
51
tions. Concerning 3D simulations, however, the model of
80 49 Zhang, which had originally been reformulated from the
47
McCormick model, indicates accurate descriptions of PLC
Block number

60 45
effect in engineering samples. Furthermore, the models

Stress (So)
43

40
41 of Rizzi and Hahner are also used to study the serration
39
37 behaviour of the materials considered in this review. This
20
35 model, which is based on thermodynamic relations, has
33
31 been evaluated in this work. The significant aspect of
29 this approach is the consideration of the Gibbs free acti-
0 200 400 600 800 1000 1200 140016001800 2000
vation enthalpy as the energy barrier subject to dynamic
Time (η-1)
strain ageing. Investigations on the PLC effect using this
B 40 200
idea indicate that the Rizzi and Hahner models are able to
35
predict the PLC band propagation in the material.
180
Space location (mm)

30
25 160 Stress (MPa) Acknowledgments: This work was co-financed by the
20 Portuguese Foundation for Science and Technology via
15 140 project PTDC/EMS-TEC/1805/2012 and by FEDER via the
10 ‘Programa Operacional Factores de Competitividade’ of
120
5 QREN.
0 100
0 5000 10,000 15,000 20,000
Time (s)

Figure 8: The PLC band properties in a Cu-Al alloy. (A) Space-time


References
plot of band propagation for ε 2 = 2E -6 s-1 obtained from the model.
[1] Winstone MR, Rawlings RD, West DRF. J. Less Comm. Metals.
(B) Experimental space-time plot of band propagation [59].
1973, 31, 143–150.
[2] Mulford RA, Kocks UF. ACTA Metall. 1979, 27, 1125–1134.
[3] Van den Beukel A, Kocks UF. ACTA Metall. 1982, 30, 1027–1034.
[4] Schwarz RB. Scripta Metall. 1982, 16, 385–390.
propagating modes (e.g. multiple band propagation with [5] Piobert A. Mem. Artillerie. 1842, 5, 525.
band collisions, serration propagation with reflections at [6] Lüders W. Dinglers Polytech. J. 1860, 155, 18.
one end and parallel propagation) can be indicated. This [7] Subramanian KH, Duncan AJ. Tensile Properties for Applica-
study shows that the model is able to monitor the promi- tion to Type I and Type II Waste Tank Flaw Stability Analysis
nent features of PLC bands in this material. The details of (U), WSRC-TR-2000-00232, Westinghouse Savannah River Com-
pany, Aiken: SC, 2000.
the numerical results are in good agreement with the ana-
[8] Savart F. Recherches sur les vibrations longitudinales Annales
lytic expressions for the band velocity, strain and width. de Chimie et de Phisique, 2nd series, 1837, 65.
[9] Le Chatelier F. Rev. Métall. 1909, 6, 914–917.
[10] Portevin A, Le Chatelier F. Comptes Rendus de l’Académie des
Sciences Paris. 1923, 176, 507–510.
4 Conclusion [11] Nabarro FR, Theory of crystal dislocations. Oxford: Clarendon
Press, 1967.
In this research, various constitutive approaches to [12] Brindley BJ, Worthington PJ. Metall. Rev. 1970, 15, 101–114.
explain the serration behaviour in metallic alloys induced [13] Robinson JM. Int. Mater. Rev. 1994, 39, 217–227.
[14] Gremaud G. Mater. Sci. Eng. A. 2004, 370, 191–198.
by the PLC effect have been studied. The theoretical
[15] Picu RC. ACTA Mater. 2004, 52, 3447–3458.
interpretation and modelling of this plastic instability in [16] Serajzadeh S, Sheikh H. Mater. Sci. Eng. A. 2008, 486,
materials was first suggested by the Penning and then 138–145.
developed later on by McCormick, Ling, Estrin, and Kubin [17] Chinh QN, Horvath G, Kovacs Z, Juhasz A, Berces G, Lendvai J,
during last decades. The previous works indicated that Mater. Sci. & Eng. A 2005, 409, 100–107.
[18] Penning P. ACTA Metall. 1972, 20, 1169–1175.
the model of McCormick, as a micromechanical constitu-
[19] Kubin LP, Estrin Y. ACTA Metall. 1985, 33, 397–407.
tive formulation, has potential to describe the PLC effects [20] Chihab K, Estrin Y, Kubin LP, Vergnol J. Scripta Metall. 1987, 21,
in different materials in terms of different types of PLC 203–208.
effects and statistical observation. The simulated results [21] McCormick PG. ACTA Metall. 1988, 36, 3061–3067.
78      S. Tamimi et al.: Modelling the Portevin-LeChatelier effects

[22] McCormick PG. Proc. ICSMA. 1988, 8, 409–414. [41] Graff S, Forest S, Strudel JL, Prioul C, Pilvin P, Bechade JL.
[23] Estrin Y, McCormick PG. ACTA Metall. Mater. 1991, 39, Mater. Sci. Eng. A. 2004, 387, 181–185.
2977–2983. [42] Graff S, Forest S, Strudel JL, Prioul C, Pilvin P, Bechade JL.
[24] McCormick PG, Estrin Y, Lowe TC, Rollett AD, Follansbee PS, Scripta Mater. 2005, 52, 1181–1186.
Daehn GS. Modelling the Deformation of Crystalline Solids, [43] Belotteau J, Berdin C, Forest S, Parrot A, Prioul C. Mater. Sci.
TMS: Warrendale, 1991, p 293. Eng. A 2009, 526, 156–165.
[25] McCormick PG, Ling CP. ACTA Metall. et Mater. 1995, 43, [44] Maziere M, Besson J, Forest S, Tanguy B, Chalons H, Vogel F.
1969–1977. Euro. J. Mech. A/Solids 2009, 28, 36–44.
[26] Kubin LP, Chihab K, Estrin Y. ACTA Metall. 1988, 36, [45] Hopperstad OS, Borvik T, Berstad T, Lademo OG, Benallal A.
2707–2718. Mod. Sim. Mater. Sci. Eng. 2007, 15, 747–772.
[27] Kubin LP, Estrin Y. ACTA Metall. et Mater. 1990, 38, 697–708. [46] Wenman MR, Chard-Tuckey PR. Int. J. Plasticity 2010, 26,
[28] Louat N. Scripta Metall. 1981, 15, 1167–1170. 1013–1028.
[29] Estrin Y, Kubin LP. J. Mech. Behavior Mater. 1989, 2, 255–292. [47] Benallal A, Berstad T, Borvik T, Clausen AH, Hopperstad OS.
[30] Cottrell AH, Bilby BA. Proc. Phys. Soc. A 1949, 62, 49. Eur. J. Mech. A-Solid 2006, 25, 397–424.
[31] Kalk A, Schwink CH. Phys Status Solidi B 1992, 172, 133–144. [48] Clausen AH, Borvik T, Hopperstad OS, Benallal A. Mater. Sci.
[32] Ling CP, McCormick PG. ACTA Metall. et Mater. 1990, 38, Eng. A 2004, 364, 260–272.
2631–2635. [49] Manach PY, Thuillier S, Yoon JW, Coer J, Laurenta H. J. Plasticity
[33] Ling CP, McCormick PG. ACTA Metall. et Mater. 1993, 41, 2014, 58, 66–83.
3127–3131. [50] Wang HD, Berdin C, Maziere M, Forest S, Prioul C, Parrot A,
[34] Jiang H, Zhang Q, Chen X, Chen Z, Jiang Z, Wu X, Fan XJ. ACTA ­Le-Delliou P. Mater. Sci. Eng. A 2012, 547, 19–31.
Mater. 2007, 55, 2219–2228. [51] Johnson GR, Cook WH. Pro. 7th Int. Symp. Ballistics 1983, 21,
[35] Lebyodkin M, Brechet Y, Estrin Y, Kubin L. ACTA Mater. 1996, 541–547.
44, 4531–4541. [52] Johnson GR, Cook WH. Eng. Fract. Mech. 1985, 21, 31–48.
[36] Lebyodkin M, Dunin-Barkowskii L, Brechet Y, Estrin Y, Kubin LP. [53] Bergstrom Y, Hallen H. Mater. Sci. Eng. 1982, 55, 49–61.
ACTA Mater. 2000, 48, 2529–2541. [54] Van den Boogaard AH, Huetink J. Comput. Methods Appl. Mech.
[37] Zhang S, McCormick PG, Estrin Y. ACTA Mater. 2001, 49, Eng. 2006, 195, 6691–6709.
1087–1094. [55] Palumbo G, Sorgente D, Tricarico L, Zhang SH, Zheng WT.
[38] Estrin Y. Dislocation-density-related constitutive modelling. In J. Mater. Pro. Tech. 2007, 191, 342–346.
Unified Constitutive Laws of Plastic Deformation, Krausz, AS, [56] Kurukuri S, Van den Boogaard AH, Miroux A, Holmedal B.
Krausz, K, Eds., Academic Press: New York, 1996, p. 69. J. Mater. Pro. Tech. 2009, 209, 5636–5645.
[39] Kok S, Beaudoin AJ, Tortorelli DA, Lebyodkin M. Mod. Sim. [57] Hahner P, Rizzi E. ACTA Mater. 2003, 51, 3385–3397.
Mater. Sci. Eng. 2002, 10, 745. [58] Rizzi E, Hahner P. J. Plasticity 2004, 20, 121–165.
[40] Kok S, Beaudoin AJ, Tortorelli DA, Lebyodkin M, Kubin L, [59] Hahner P, Ziegenbein A, Rizzi E, Neuhausser H. Phy. Rev. B
­Fressengeas C. J. Phys. IV 2003, 105, 191–197. 2002, 65, 134109.

You might also like