You are on page 1of 35

Accepted Manuscript

Determination of the RVE size for polycrystal metals to predict monotonic and cyclic
elastoplastic behavior: Statistical and numerical approach with new criteria

Moussa Bouchedjra, Toufik Kanit, Cherif Boulemia, Abdelwaheb Amrouche,


Mohamed El Amine Belouchrani

PII: S0997-7538(17)30779-9
DOI: 10.1016/j.euromechsol.2018.04.011
Reference: EJMSOL 3592

To appear in: European Journal of Mechanics / A Solids

Received Date: 18 October 2017


Revised Date: 11 April 2018
Accepted Date: 14 April 2018

Please cite this article as: Bouchedjra, M., Kanit, T., Boulemia, C., Amrouche, A., El Amine Belouchrani,
M., Determination of the RVE size for polycrystal metals to predict monotonic and cyclic elastoplastic
behavior: Statistical and numerical approach with new criteria, European Journal of Mechanics / A Solids
(2018), doi: 10.1016/j.euromechsol.2018.04.011.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT
Determination of the RVE size for polycrystal metals to predict monotonic and cyclic
elastoplastic behavior: Statistical and numerical approach with new criteria

Moussa Bouchedjra a,b,c, Toufik Kanit d, Cherif Boulemiab, Abdelwaheb Amrouche b,c*, Mohamed El Amine
Belouchrania, e
a
Laboratoire Génie des Matériaux, Ecole Militaire Polytechnique, BP17C, Bordj El Bahri, 16046
Alger, Algeria.

PT
b
Université d’Artois, Technoparc Futura, 62400 Béthune, France
c
Laboratoire Génie Civil et géo-Environnement (LGCgE, EA 4515), Technoparc Futura, 62400

RI
Béthune, France
d
Université de Lille, Unité de Mécanique de Lille, (UML EA 7512) 59650 Villeneuved’Ascq,

SC
France
e
Ecole Nationale Supérieure de Technologie, Dergana, Alger, Algeria

U
Corresponding authors: Abdelwaheb Amrouche; abdelwaheb.amrouche@univ-artois.fr
AN
Abstract
M

The determination of the size of the Representative Volume Element “RVE” is extremely important
to predict mechanical and physical effective properties of heterogeneous materials. In the present
D

work, the optimal size of RVE for polycrystal materials, which is made–up of grains having a
polyhedron Voronoi shape, is investigated for the prediction of an elastoplastic behavior using the
TE

crystal plasticity model. Ten realizations with a different number of grains ranging from 10 to 250
have been used in the numerical simulation. The homogenization technique combined with the
EP

Finite Element Method (FEM) is used to obtain the macroscopic strain and stress. Several new
criteria are introduced as indicators for estimating the optimal RVE's size. Their formulation is
C

based on the standard deviation with respect to the average value of different mechanical quantities.
These criteria ensure that, for an optimal RVE size, the response of the crystalline aggregate is not
AC

sensitive to the orientations evolution and grains’ dimensions. It turns out that, for the cyclic
behavior, the influence of the loading mode is very important, in particular the asymmetrical stress
mode. The influence of the value of strain or stress used to perform different simulation tests on the
RVE size has been analyzed. The simulation results are in good agreement with the experimental
results from the literature.

Keywords
Representative volume element; Homogenization; Finite element analysis; Polycrystals; Plasticity
1
ACCEPTED MANUSCRIPT

Nomenclature

Symbols
X average value of variable X
Σ effective (macroscopic) stress
Σ vM vom Mise effective stress
Σ ii axial effective stress (i=1, 2 and 3 according x, y and z direction, respectively)

PT
E effective (macroscopic) strain
E rat macroscopic ratcheting strain
∆E p macroscopic plastic range deformation

RI
Σ yshear macroscopic apparent elastic shear yield
σ local (microscopic) stress

SC
σ vM local (microscopic) von Mise stress
ε local (microscopic) strain
C fourth–rank tensor of elastic modulus

U
Li number of finite elements used to mesh one Voronoi cell (grain)
AN
ε 11 axial strain according x-direction
U1 displacement according x-direction
U2 displacement according y-direction
M

U3 displacement according z-direction


α har hardening rate
D

C criterion
Φ radius of equivalent sphere containing grain
Φ
TE

average of radius of equivalent sphere containing grain


Subscripts
ref reference
EP

used used
min minimum
max maximum
SD standard deviation
C

cy cyclic
S static
AC

K kinematic
SH shear
rat ratcheting
P plastic
e elastic
har hardening

Superscripts

s slip system
2
ACCEPTED MANUSCRIPT

1. Introduction
Homogenization techniques are robust methods for numerical determination of the effective
properties of heterogeneous materials. These have been applied in several works: for nonlinear
composites with almost periodic microstructures as well as random materials such as polycrystals in
Sanchez–Palencia and Zaoui [1], Nemat–Nasser and Hori [2] for linear properties, elasticity and
thermal conductivity, and the nonlinear properties, plasticity, in Salahouelhadj and Haddadi [3].

PT
The prediction of the macroscopic yield strength of polycrystalline metals subjected to plastic
forming in Watanable and Terada [4] and the computation of mechanical proprieties of porous

RI
material in Masmoudi et al. [5]. By evaluating effective elastic properties of heterogeneous
materials, Patil et al. [6] have shown that extended finite element method (XFEM) combined with

SC
the multi-scale approach leads to a significant reduction in the CPU time compared to the standard
XFEM method.

U
The resolution of homogenization problems is based on numerical simulation techniques applied to
samples representing the material. The latter is the Representative Volume Element “RVE” which
AN
plays an important role to ensure a good solution.
Over the last decades, several investigations have been conducted to determine the RVE size for all
M

categories of materials, composites, polymers and polycrystalline metals. Drugan and Willis [7]
have shown that for composite materials consisting of an isotropic matrix reinforced by random
D

distribution of isotropic spherical inclusions, the RVE size is at most twice the inclusion diameter
for any reinforcement concentration level. For an infinite medium, they considered several sets of
TE

matrix and inclusion module without considering statistical fluctuations of the effective properties
over finite domains.
EP

Gusev [8], in the statistical study of realizations of periodic elastic composite with disordered
distribution of spheres generated independently by Monte–Carlo calculation technique, showed that
C

the RVE containing a dozen of spheres were necessary to obtain a small scatter for calculation of
AC

the average of elastic properties. By conducting a statistical study of random composite and two–
phase 3D Voronoi mosaic microstructures, Kanit et al. [9] have shown that the determination of
RVE size depends on investigated properties such as thermal conductivity, elastic modulus, shear
modulus and volume fraction. Ostoja–Starzewski et al. [10] have found, in comparative study of
random composite materials that the discrepancy increases as follows: (i) linear elasticity, (ii)
plasticity, (iii) linear thermo–elasticity, (iv) nonlinear elasticity and (v) permeability. Recently,
Mirkhalaf et al. [11] have proposed a methodology based on statistical analysis and numerical
simulation to determine the size of the RVE for heterogeneous amorphous polymer subjected to

3
ACCEPTED MANUSCRIPT
finite deformation. They have used two criteria; one is the first invariant of the Piola–Kirchhof
stress tensor and the second checks if by increasing the size of the micro-structural sample, the
deformation behavior changes. If no considerable change is observed, while increasing the size,
these criteria are considered satisfying. Very recently, Koohbor et al. [12] have used an optical–
based experimental approach to estimate the length scale of the RVE of plain woven composite.
They have used a simple algorithm based on the strain averaging method to identify the length
scales.

PT
Ren and Zheng [13] have evaluated the effective shear and Young’s modulus using polycristal
made-up of square grains under a planar stress. They found that the RVE size has a roughly linear

RI
dependence on the anisotropy degree of the single crystal. Using RVE size of less than 256 grains
and with maximum relative error of 5%, these authors [14] also found an almost linear dependence

SC
for a 2D polycrystal of cubic structure with realistic shape formed by crystallization processes.
Nygards [15] studied cubic-symmetric materials using polycrystals consisting of periodic Voronoi

U
grain structures in 2D and 3D. He showed that the number of grains required in the RVE is
proportional to the anistropy degree. These same results were found and reported by Ren and Zheng
AN
[13, 14].
Ranganathan and Ostoja–Starzewski [16] have proposed a methodology, based on a scaling
M

function that relates the single crystal anisotropy to the scale of observation, for the determination
of the RVE size in linear elastic randomly structured polycrystals made–up of cubic single crystals.
D

The RVE size is identified by setting up stochastic Dirichlet and Neumann boundary value
problems consistent with the Hill–Mandel macro–homogeneity condition. The authors estimate the
TE

RVE size of several metallic materials. Ranganathan and Ostoja–Starzewski [17] applied the latter
approach to assess the RVE size of viscoplastic polycrystals. They succeeded to set a RVE of about
EP

8 grains for both simple and two-phase polycrystals.. Chih [18] have used variational bounds to
estimate elastic moduli of random orthorhombic polycrystals.
C

Salahouelhadj and Haddadi [3] have evaluated the number of grains needed in a RVE for random
AC

textured polycrystals of copper. They have used finite element simulations of tensile test on a
sufficient number of polycrystalline aggregates extracted at various positions, which are taken large
enough to be representative of mechanical behavior of the studied polycrystalline. They found that
400 grains for the 2D case and 1000 grains for the 3D case give a maximum relative error of 1.5%.
Up to now, research works focused on determination of the RVE size of polycrystal metals are only
performed with the monotonic tensile test, see Salahouelhadj and Haddadi [3], which is not
sufficient. On one hand, it cannot ensure that the RVE used is optimal for the other types of tests,
monotonic shear and cyclic loading. On the other hand, it does not take into account all the

4
ACCEPTED MANUSCRIPT
heterogeneities of polycrystal metals, the shape and the spatial orientation of grain, or accomplished
with cubic shape of grain which is not realistic.
The main objective of this paper is a qualitative study of the influence of several loadings under
monotonic and cyclic cases. These two configurations are conducted a with a static loading called
"Static Boundary Conditions SBC" and kinematic loading named "Kinematic Boundary Conditions
KBS" for determining RVE size of random polycrystals materials in elastoplastic behavior. For
each volume size and number of grains, 10 realizations with different independent orientations and

PT
grain sizes are generated. Then, different finite element analyses tests are performed on each
volume size. For each test, new criteria are introduced as indicator of accuracy for the determination

RI
of the effective elastoplastic properties of polycrystal materials. These are (i) C (Σ11 max ) is the

maximum effective axial stress Σ11max , (ii) C ( E11max ) is maximum effective axial deformation E11max ,

SC
(iii) C (α harK ) , C (α harS ) and C (α harSH ) are the rate of hardening in kinematic and static tensile test

and the rate of hardening shear test, respectively, (iv) C (Σ yshear ) is the apparent elastic shear yield,

U
(v) C ( E rat ) and C ( ∆E p ) are macroscopic ratcheting strain and plastic range with asymmetrical stress
AN
mode and (vi) C (Σ11cy max ) is maximum stress reached per cycle with symmetrical strain mode.

The paper is organized as follows:


M

• Section 2 outlines the methodology of generation of polycrystals.


• Section 3 is concerned with the study of the effect of mesh density on elastoplastic behavior of
D

polycrystals, where optimal mesh density (number of elements to mesh one grain) is investigated
TE

which ensures accuracy less than 1%.


• Section 4 deals with the constitutive elastoplastic model with boundary conditions used for
EP

performing finite element analysis tests.


• A definition of criteria is introduced for each test in Section 5.
• Section 6 provides a forum of discussion of then results, where the influence of types test on RVE
C

is shown.
AC

• Evolution of proposed new criteria with size of RVE of polycrystal materials is given in Section7.
• Comparison of simulation predictions with experimental results for 5083 Aluminum alloy is
performed in Section 8. Finally, conclusions are drawn based these simulation results.
2. Polycrystalline generation and meshing
Several methods have been proposed to construct random polycrystals morphologies. Zhang et al.
[19], Döbrich et al. [20] and Rowenhorst et al. [21] performed experimental characterizations of
real 3D polycrystals. They provide important information about the distribution of grain size and
grain boundaries, because this method is expensive, only few small samples have been analyzed.
5
ACCEPTED MANUSCRIPT
Analytical and numerical techniques based on Voronoi tessellations are used to generate random
polycrystal morphologies, see Barbe et al. [22, 23] and Zhao et al. [24]. These techniques show
significant discrepancies in grain size and shape which are representative of real polycrystal
morphologies, but it is very difficult to mesh them with high quality elements.
In this work, the polycristal morphology generation is based on 3D Voronoi tessellation which is a
collection of n-entities that fills the space with no overlaps and no gaps. These entities have a
polyhedral shape such as their center is defined as influence zones of a particular set of points

PT
which are considered to be randomly distributed. From a physical point of view, the generation of
Voronoi tessellations corresponds to a process of solidification or recrystallization where all grains

RI
nucleate at the same time and grow isotropically Quey et al. [25].
The aggregate was generated without a repulsion distance function. This function ensures that the

SC
distance between any two random Voronoi seeds is above a limit, resulting in a more homogeneous
grain volume distribution. According to Gonzalez et al [26], using this technique, small grains are

U
removed which may influence the polycrystal behavior since features such as small grains and small
edges are present in real microstructures.
AN
Quey et al. [25] have used an algorithm called ‘‘regularisation’’ (implemented in a Neper software
open source) to remove these small entities and faces that can lead to numerical issues. This work
M

revealed a small effect of regularization on the grain size distribution. It is worth noting that with
using this technique allow meshing the polycrystalline aggregates with high quality.
D

For each size of crystalline aggregate, 10 different realizations are generated. For each realization,
its orientations and dimensions of the grains are totally different to other realizations. Figure 1 and 2
TE

show respectively, the distribution of grain sizes and their orientations for crystalline aggregate
containing 250 grains.
C EP
AC

6
ACCEPTED MANUSCRIPT

PT
RI
SC
Figure 1. Grains size distribution for ten different realizations, Φ and Φ are radii and average of
radii of equivalent sphere, respectively.

U
AN
M
D
TE
C EP
AC

7
ACCEPTED MANUSCRIPT

Realization 9 Realization 7 Realization 5 Realization 3 Realization 1


Figure 2. Pole figures of ten realizations: orientations for each realization.

PT
RI
U SC
AN
M
Realization 10 Realization 8 Realization 6 Realization 4 Realization 2

D
TE
C EP
AC
8
ACCEPTED MANUSCRIPT
3. Determination of mesh density
This section is devoted to the effect of mesh density on elasto-plastic behavior of polycrystals with
the aim to ensure a sufficient quality of the presented results and to eliminate errors coming from
the mesh sensitivity. The tensile test with crystal plasticity model described in section 4 is used to
investigate optimal mesh density (average number of finite elements used to mesh one Voronoi
cell). The tensile test is simulated with displacements fixed in faces x0, y0 and z0 in x-, y- and z-
directions, respectively. Forces and displacements are applied in the opposite faces x1, y1 and z1.

PT
According to Pilvin et al [27], Héripré et al [28] and Cailletaud et al [29], this boundary condition is
referred to as mixed.

RI
The number of cells and the geometry of the microstructure are unchanged. Different mesh sizes Li

have been varied from the L1 level to Lref oreference level for each size. Different values of Li are

SC
given in Table 1.
Figure 3 shows meshing of polycrystalline aggregates made up of 250 grains for 3 different levels

U
while Figure 4 shows the meshing of some grains for these 3 different levels.
AN
Table 1. Different mesh refinement levels
Levels L1 L2 L3 L4 Lused Lref
M

10 grains 153 683 1391 3220 4732 5224


50 grains 194 327 702 1293 3951 4651
100 grains 195 210 251 719 2162 3173
D

150 grains 273 339 540 819 1054 1591


TE

200 grains 220 221 335 497 1007 1275


250 grains 291 357 412 610 1003 1294
C EP
AC

L2 Lused Lref

Figure 3. Discretization at different refinement levels of the 250-grains RVE.

9
ACCEPTED MANUSCRIPT

PT
L2 Lused Lref

RI
Figure 4. Discretization at different refinement levels of some grains.

SC
The relative difference (∆X) of Σ11max and of ΣvM (von Mises equivalent stress) between two
successive refinement levels is given by following formula:

U
X ( Li ) − X ( Li −1 )
∆X =
AN
X ( Li ) (1)

For both quantities (Σ11max and ΣvM ), the first Li which corresponds to a relative difference of less
M

than 1%, is taken as a reference Lref. Once done, the relative error of each level is calculated using
the following formula:
D

X ( Li ) − X ( Lref )
Err(%) =
TE

(2)
X ( Lref )

Table 1 shows that the mesh density to obtain accuracy less than 1% depends on both the size of
EP

polycrystalline aggregates and the number of grains.


For each volume size (number of grains), the relative error Err (%) with respect to the number of
C

elements Li to mesh one grain is depicted in Figures 5a-b.


AC

10
ACCEPTED MANUSCRIPT

PT
RI
Figure 5. Relative error Err[%] versus Li (number of elements to mesh one grain) for different

SC
sizes (number of grains) of polycrystalline aggregates: (a) Σ11max and (b) Σ vomMises.

Figures 5 (a) and (b) show that the convergence of error of Σ vM is faster than error of Σ11max .

U
Hence, the values Li for which the error of Σ vM is less than 1% are less than values Li of which the
AN
error of Σ11max is less than 1%. Thus, it can be concluded that the use of Σ11max for the determination

of the optimal density size is more accurate than with using Σ vM .


M

For both quantities ( Σ11max and Σ vM ) used for the determination of the optimal density mesh size,
D

Figures 5 a and b show that the convergence of relative error depends on volume size. This
convergence is faster for larger sizes (150, 200 and 250 grains) than for smaller ones 10, 50 and 100
TE

grains). The intersection of the line of the limit value with the error curves of large size, more than
150 grains, is practically at the same point, but in smaller cases, this intersection occurs at different
EP

points. Thereby, it can be concluded that for a size larger than 150 grains, the use of mesh densities
Li around 1000 elements per grain gives error of less than 1% using Σ11max .
C

The fluctuation of the error using Σ22max and Σ33max compared to those of Σ11max is less than 0.5%,
AC

this ensures that the optimal mesh density obtained by Σ11max being more rigorous in all directions
(x, y and z-directions).
For the polycrystalline aggregate and some grains meshed with 3 different levels L2 , Lused and Lref ,

Figures 6 and 7 show distribution of the von Mises equivalent stress σ vM and the axial deformation

ε11 , respectively. The distributions of the two quantities ( σ


vM and ε 11 ) are closer for

11
ACCEPTED MANUSCRIPT
polycrystalline aggregate meshed with density Lused and those meshed with density Lref . However,

they are obviously different with aggregate meshed with density L2 .


On one hand, this study ensures the use of the optimal mesh density for each size (number of
grains) of polycrystalline aggregates without precision effect of the of the RVE size determination,
and on the other hand, it allows to save CPU time.

PT
RI
U SC
AN
M
D

L2 Lused Lref
TE

Figure 6. Distribution of von Mises equivalent stress (MPa) for polycristal aggregate surfaces
and some grains extracted randomly from polycrystalline aggregate made–up of 250 grains for
different mesh densities.
C EP
AC

12
ACCEPTED MANUSCRIPT

PT
RI
SC
L2 Lused Lref

U
Figure 7. Distribution of ε 11[%] for polycrystal aggregate surfaces and of some grains extracted
AN
randomly from polycrystalline aggregate made–up of 250 grains for different mesh densities.

4. Elastoplastic constitutive model and boundary conditions


M

Simulations of the elastoplastic behavior of polycrystals have been performed with finite element
analysis, where an implicit method was adopted for the integration of constitutive laws with a
D

Newton-type method for handling the nonlinear problem.


TE

The behavior of each integration point of the finite element mesh is described by a constitutive
model that was initially introduced by Méric et al. [30]. It has been extensively used in several
EP

works which have proposed an important groundwork, see Barbe et al. [22, 31, 32], Diard et al.
[33], Osipov et al.[34], Sai et al. [34], and more particularly in the direct comparison of
experimental and simulated intracrystalline strain fields, see Musienko et al. [36].
C
AC

4.1. Elastoplastic constitutive model


In the framework of small perturbation, the total strain ε is divided additively into an elastic part
ε e and plastic part ε P as:

ε = εe + ε P (3)

The relationship between the elastic strain ε e and the stress σ is given by the Hooke’s law:

13
ACCEPTED MANUSCRIPT
σ = C :εe (4)

Where C is the fourth–rank tensor of the elastic modulus.


The plastic strain ε P results from dislocation slip on the crystallographic slip system is given by the
following expression:

• •
ε P = ∑ ms γ s (5)

PT
s

with:

RI
• •
γ s = v s sign(τ s − x s ) (6)

SC
• τ s − X s − Rs
v =<
s
>n (7)
K

U
where:
AN
< x >= Max(x,0) (8)

and:
M

v s ( t = t0 ) = 0 (9)

K and n are the material parameters that control the viscosity of the material; X s and R s are
D

kinematic and isotropic hardening variables of slip system s , respectively.


TE

The kinematic and isotropic hardening variables follow the nonlinear evolution defined by the
classical phenomenological formulation and are e given by the following expressions:
EP

X s = cα s (10)

• • •
α s = γ s − dα s v s (11)
C

with:
AC

α s (t = t 0 ) = 0 (12)

R s = R0 + Q∑ his (1 − e −bv ) (13)


i

c and d are the material parameters of kinematic hardening ; R0 , Q and b are the material
parameters of isotropic hardening, and his are components of the interaction matrix.

The resolved shear stress τ s directly results from the state of stress σ and the Schmidt tensor m s
as:

14
ACCEPTED MANUSCRIPT
τ s = σ : ms (14)
with:

ms = ns ⊗ l s + l s ⊗ ns (15)

where n s is the unit vector normal to the slip plane and l s is the slip direction.
Simulation of different tests has been performed using the material parameters given in Table 2.
These parameters have been chosen in such a way that: (i) Elasticity and Norton parameters from

PT
Lu et al [37] for 5083 Al alloy, (ii) the components of the matrix interaction values ( hi ) proposed

by Devincre et al. [38].

RI
Table 2. Material parameters of elastoplastic model.

Elasticity parameters E(MPa) 70000

SC
v 0.3
Norton parameters K (MPa) 20
n 50

UR0 (MPa) 46.43


Isotropic hardening
AN
Q (MPa) 11.08
b 12.01
Kinematic hardening c (MPa) 1100
M

d 1.0126
h1 0.03
h2 0.02
D

Components of interaction matrix h3 0.0454


TE

h4 0.624
h5 0.137
h6 0.122
EP

4.2. Boundary and loading conditions


This section is devoted to describing boundary conditions, which are used to simulate different
C

tests. Two types of boundary conditions are considered for all test types used.
AC

The used boundary conditions to simulate kinematic tensile and shear tests consist of prescribing a
uniform displacement U on the RVE’s faces of, (see Figure 8) as detailed in follows:
• Tensile monotonic tests:
To simulate monotonic kinematic tensile test, uniform displacement U is applied on face x1 and

face x0 is fixed as:

U1 ( x0 ) = 0 with: U1 ( x1 ) = U (KBC), the faces y0 , y1 , z0 and z1 remain free

15
ACCEPTED MANUSCRIPT
To simulate monotonic static tensile test, uniform traction Σ is applied on face x1 and face x0 is
fixed as:
U1 ( x0 ) = 0 with: Σ11 ( x1 ) = Σ (SBC), the faces y0 , y1 , z0 and z1 remain free

• Shear monotonic test:

U1 ( y0 ) = 0

U 2 ( y0 ) = 0 with: U1 ( y1 ) = U , the faces x0 , x1 , z0 and z1 remain free.
U ( y ) = 0
 3 0

PT
RI
U SC
AN
M
D
TE

Figure 8. Faces of the polycrystalline aggregate.

5. Procedure and criteria for the RVE’s determination


EP

The procedure for RVE size determination is a statistical technique, see Kanit et al. [9], combined
with numerical simulations. In this proposed methodology, the first step is to generate different
realizations, sizes and orientations of grains of one volume (number of grains) of polycrystals
C

materials. In the second step, the realizations are submitted to different types of tests: (i) Tensile and
AC

shear tests with a monotonic loading under static and kinematic configurations, (ii) Cyclic loading
tests under the same configurations. For each test type, criteria are introduced as an indicator of the
optimal RVE size. The formulation of these criteria is based on the percentage of the ratio of the
standard deviation in terms of averaged value, which ensures that changing the realization
(orientations and grain size), for the same realization size (number of grains), does not change the
elastoplastic behavior, see Mirkhalaf et al. [11].

16
ACCEPTED MANUSCRIPT
Using 10 realizations for one size, standard deviation X SD and average X are given by the
following relationships:
1 n=10
X SD = ∑(X n − X )2
10 n=1
(16)

1 n=10
X = ∑ Xn
10 n=1
(17)

PT
5.1. Criteria for monotonic test

RI
For the monotonic tensile tests, two criteria are used. In the case of a kinematic boundary condition,
the first criterion is the maximum effective axial stress Σ11max , see Figure 9a, which is given by the

SC
following expression:

U
Σ11 max SD
C ( Σ11 max ) = (18)
Σ11 max
AN
where Σ11max SD and Σ11max are the standard deviation and the average macroscopic axial stress σ 11
respectively.
M

In the case of a static boundary condition, the maximum effective axial deformation E11max , see
D

Figure 9b, is given by the following expression:


TE

E11max SD
C ( E11max ) = (19)
E11max

where E11max SD and E11max are the standard deviation and the average macroscopic axial strain ε 11
EP

respectively.

The second criterion is the slope α har , see Figure 9c, which represents the hardening rate for
C

kinematic and static tensile tests noted C (α harK ) and C (α harS ) , respectively.
AC

17
ACCEPTED MANUSCRIPT

PT
Figure 9. Tensile stress-strain curve.

RI
For shear tests, the first criterion is the slope that represents the rate of hardening, see Figure 9c, as
defined for tensile test. This is noted as follows:

SC
α harsSHSD
C (α harSH ) = (20)
α harSH

U
The second criterion, noted C (Σ yshear ) , is the apparent elastic shear yield Σ yshear , which is given by
AN
the intersection of the asymptote of the plastic part with the straight line of the elastic part of the
shear stress-strain (macroscopic stress-strain) curve, see Figure 10. It is given by:
M

Σ yshearSD
C (Σ yshear ) = (21)
D

Σ yshear
TE
C EP
AC

Figure 10. Shear strain-stress curve.

5.2. Criteria for cyclic test


18
ACCEPTED MANUSCRIPT
For cyclic tests, with asymmetrical stress mode, the criteria are defined from macroscopic
ratcheting strain Erat and plastic range ∆E p , see Figure 11a, which are used for studying the

fatigue behavior, see Shi et al. [39], Sun et al. [40], Hamidinejad and Varvani-Farahani [41], Pun et
al. [42] and Yuan et al. [43], which are given by following expressions:
E ratSD
C ( E rat ) = (22)
E rat
∆E PSD
C ( ∆E p ) = (23)
∆E P

PT
Where the ratcheting strain Erat is calculated by following expression:

RI
E11max + E11min
Erat = (24)
2

SC
where E11max and E11min are respectively the maximum and the minimum strain in one cycle, see
Figure 11a.

U
For the cyclic test with a symmetrical strain mode, one criterion is used and is defined from the
maximum stress reached per cycle, see Figure 11b. It is given by:
AN
Σ11cy max
C (Σ11cy max ) = (25)
Σ11cy max
M
D
TE
C EP
AC

Figure 11. Stress-strain hysteresis loops.

The macroscopic stress Σ and strain E are calculated by the following expressions :
1
Σ= ∫ σ dv
V RVE
(26)

1
E = ∫ ε dv
V RVE
(27)

where σ and ε are respectively the local stress and strain.

19
ACCEPTED MANUSCRIPT
6. Results
6.1. Influence of monotonic loading modes on the RVE size
As previously reported, the numerical simulations based on finite element method are carried out
for all tests described in prior sections. For each size of the polycrystalline aggregates, kinematic
and static tensile tests have been simulated with: (i) applying the uniaxial strain in face x1 up to 5%
and (ii) applying uniform stress of 280 MPa in face x1 according to the x direction respectively.

The average value and standard deviation of the macroscopic stress Σ11max corresponding to 5% of

PT
strain are plotted in Figure 12a. It can be noted that the standard deviation decreases with
increasing the size of polycrystalline aggregates (number of grains), with a slight fluctuation of the

RI
average value, this means that 10 realizations used for each polycrystalline aggregate size are
sufficient. For all realizations, the value of Σ11max has a tendency to converge to a common value

SC
which is the mechanical response of the polycrystalline material. The results are given in Figure
12b, in regards to the hardening rate α harK , which show the similar tendency than stress Σ11max .

U
The results plotted in Figure 12c and d show that similar conclusions can be drawn to that given in
AN
previous case of kinematic tensile test for strain Σ11max at 280 MPa which decreases with increasing

the number of grains (size of RVE), and slight fluctuation for the average value.
M

The monotonic shear tests are performed by applying a uniform deformation E12 of 5% in the x-
direction. The standard deviation and average value of apparent shear yield and hardening rate are
D

presented in Figures 12e and f respectively.


TE

We notice that the standard deviation decreases with increasing the size of polycrystalline aggregate
for the apparent shear yield and the hardening rate. One note that the average value of apparent
elastic shear yield Σ yshear fluctuates slightly, which varies between 58 to 60 MPa for a percentage of
EP

about 3%. For the hardening rate, we note a slight fluctuation for sizes larger than 100 grains.
However, those of 10 and 50 grains have a fluctuation of about 5% compared to 250 grains.
C

It should be noted that these results show that polycrystalline aggregate of less than 100 grains
AC

required more than 10 realizations for statistical analysis. As pointed out by Kanit et al. [9], the
evaluation of the effective elastic constants of the polycrystalline can be established either by
performing a large number of simulations on small aggregates or by using a small number of
simulations on large aggregates.

20
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Figure 12. Average value and standard deviation vs. size of polycrystalline aggregates: (a) Σ11max

, (b) α harK , (c) E11max , (d) α harS , (e) Σ yshear and (f) α harSH .

21
ACCEPTED MANUSCRIPT

For the two realizations, which were having maximum discrepancy, Figures 13a and b show a
comparison of monotonic stress-strain curves of two sizes (100 and 250 grains) of polycrystalline
aggregates,. For static and kinematic boundary conditions, the gap is evident on the size of 100
grains, see Figures 13a, but in a size of 250 grains, see Figures 13b, the curves are closer with a
very slight gap for the tensile test with two types of boundary conditions.
These results confirm that the responses are independent of types of boundary conditions for the

PT
RVE of 250 grains. It can be noted that the estimation of the RVE of polycrystals with the
polyhedron shape of grains for predicting the effective elastoplastic behavior is about 250 grains.

RI
U SC
AN
M
D

Figure 13. Monotonic strain–stress curves with static and kinematic boundary conditions: (a) RVE
contains 100 grains and (b) RVE contains 250 grains.
TE

6.2. Influence of cyclic loading modes on the RVE size.


EP

In this Section, the influence of cyclic loading on the RVE size of polycrystal materials is studied.
Cyclic test simulations were performed under different cyclic loading modes, namely the
C

symmetrical strain and asymmetrical stress modes, which are used for studying the materials fatigue
behavior.
AC

For cyclic test under asymmetrical stress mode, Figure 14a and b show that average value and
standard deviation of ratcheting strain ( Erat ) and plastic range deformation ( ∆E p ) decrease with

increasing polycrystalline aggregates size. However, for cyclic test under symmetrical strain mode,
the average value of Σ11cy max increases with increasing size of polycrystalline aggregates and this

standard deviation decreases with increasing polycrystalline aggregates size, see Figure 14c.

22
ACCEPTED MANUSCRIPT

PT
Figure 14. Average value and standard deviation versus size of polycrystalline aggregates (number
of grains): (a) Erat , (b) ∆E p and (c) Σ11cy max .

RI
For two realizations which were having maximum discrepancy and for symmetrical strain mode, as

SC
shown in Figure 15, the macroscopic stress-strain hysteresis loops are seen with little gap for small
size (50 and 100 grains) and become very closer for 250 grains. But, for the asymmetrical stress
mode, the gap is evident for small size and decreases with increasing size of RVE. With a size of

U
250 grains, the stress-strain hysteresis loops are not superposed; see Figure 16, as found in
AN
symmetrical strain mode.
M
D
TE
EP

50 grains 100 grains 250 grains


Figure 15. Stress-strain hysteresis loops in the first cycle under symmetrical strain loading mode
for different sizes of RVE.
C
AC

50 grains 100 grains 250 grains


Figure 16. Stress-strain hysteresis loops in the first cycle under asymmetrical stress loading mode
for different sizes of RVE.
23
ACCEPTED MANUSCRIPT

Finally, we can consider the sensitivity of the RVE size with the cyclic loading mode. It can be
stated that, for a cyclic loading under asymmetrical stress mode, the RVE size has to be more than
250 grains in order to obtain an accuracy of 1%.
In order to account of the influence of the cyclic amplitude loading, cyclic tests are added and
simulated with two others amplitudes of 0.6% and 1%, on two realizations contain 250 grains,
which exhibited maximum discrepancy. The results given in Table 3 show that the variation of the

PT
percentage of the maximum deviation from the average value of stress Σ11cy max is less than 1% with

a cyclic amplitude increasing from 0.3% to 1%.

RI
Table 3. The percentage of the maximum deviation to the average value of stress Σ11cy max as

SC
function of cyclic load amplitude.
Σ11max − Σ11min
Amplitude cyclic load
Σ

U
0.3 % 1.41%
0.6 % 1.84%
AN
1% 2.07%

For 2 different realizations with 250 grains, which were presenting maximum discrepancy. Figure
M

17 shows that there is no gap on stress-strain hysteresis loops. Therefore, one can state that the RVE
size depends slightly on the amplitude of the cyclic load.
D
TE
C EP
AC

Figure 17. Stress-strain hysteresis loops during the first cycle under symmetrical strain loading
mode for different amplitudes.

24
ACCEPTED MANUSCRIPT
7. Approach with new criteria.
The evolution of the new criteria, described in the previous section 5, according to the size of the
RVE of the polycrystalline materials, as well as the influence of the strain and the stress ratio used
for monotonic and cyclic simulation tests, are developed here.
7.1 Influence of strain and stress rate on RVE size
7. 1.1. Monotonic loading case
The aim seeked here is to study the influence of strain and stress value applied on RVE to simulate

PT
the kinematic and static tensile tests, respectively. For kinematic tensile, Figure 18a shows a slight
increasing of C (Σ11max ) when the applied strain increases. For different polycrystal aggregate sizes

RI
(50, 150 and 250 grains), fluctuations do not exceed 1%.
Nevertheless, for static tensile, Figure 18b shows that C ( E11max ) decreases with the increase of the

SC
applied stress with fluctuation around 2%. For two types of boundary conditions used, the
fluctuations of the two criteria depend slightly on value of strain or stress used for simulation.

U
Therefore, the RVE size has a very low dependence on the value of strain or stress used to perform
AN
the tensile test. For kinematic tensile test, Salahouelhadj and Haddadi[3] found similar results when
considering a cubic shape of grains.
M
D
TE
C EP
AC

Figure 18.Influence of value of strain or stress used for performing tensile test: (a) kinematic tensile
test and (b) static tensile test.

7. 1.2. Cyclic loading case


The asymmetrical stress mode is performed by applying load amplitude of 160 MPa and a mean
stress of 20 MPa in face x1 in the x-direction. The results depicted in Figure 19a and b, for three

sizes of polycrystalline aggregates, show very slight fluctuation of C ( ∆E p ) and C ( E rat ) with

increasing number of cycles. For a size corresponding to 250 grains, the fluctuations are less than
25
ACCEPTED MANUSCRIPT
1%. For the first cycle, C ( ∆E p ) and C ( E rat ) are 5.28% and 4.02% respectively. Then for tenth

cycle, they become 4.96% and 4.18%, respectively.


The symmetrical strain mode is performed by applying a uniform deformation E11 of 0.3% in the x-

direction. Figure 19c shows that C (Σ11cy max ) is slightly dependent on number of cycles applied.

Note that similar results were found with applying asymmetrical stress mode.
It should be noted that the criteria having been considered for the RVE size study for cyclical

PT
behavior are not sensitive to the number of cyclic loads. As a result, cyclic test simulations were
performed by applying one cycle load thereby saving CPU time.

RI
U SC
AN
Figure 19.Evolution of criteria of cyclic tests versus number of cyclic loading: (a) C ( ∆E p ) , (b)
M

C ( E rat ) and (c) C (Σ11cy max ) .


D

7. 2. Evolution of the proposed criteria with RVE size


TE

The convergences of the maximum relative deviation of all monotonic tests criteria, described in
Section 5, are depicted in Figure 20a. It turns out that the convergence of C (Σ11max ) and C (Σ yshear )

is faster than that of the other criteria C ( E11max ) , C(α harK ) , C (α harS ) and C (α harSH ) .
EP

The results show that, for polycrystalline aggregates with the smallest size (10 grains), the
maximum discrepancy is 11.58% for C ( E11max ) and 9.17% for C (α harS ) . For other criteria, this
C

discrepancy is between 4% and 6%. Then, for size of 50 grains, the relative dispersion is reduced to
AC

half for all the criteria (2.36% for C (Σ11max ) , 3.32% for C(α harK ) , 6. 27% for C ( E11max ) , 5.56% for

C (α harS ) and 2,39% C (Σ yshear ) , except C (α harSH ) which remain about 6%.

For polycrystalline aggregates with large size (250 grains), the accuracy of estimation the effective
stress Σ11max is 0.79% < 1%, and it is greater than 1% for the other criteria: 1.51% for estimation
αharK, 2.25% for E11, 1.52% for αharS, 1.25% for Σyshear and 1.83% of αharSH. For crystalline
aggregates made–up of grains with cubic shape, Salahouelhadj and Haddadi [3] have shown that, to
obtain an accuracy of 2% in the estimation of the effective stress, the RVE size for FCC copper has
26
ACCEPTED MANUSCRIPT
to be of 250 and 512 grains using 2D and 3D models, respectively. For accuracy of 1.5%, the
selected RVE size is doubled. In present work, the accuracy is less than 1% to estimate the effective
stress considering a RVE having 250 grains Voronoi-polyhedron shape, in a 3D tensile test.
The convergence of criteria used for studying the influence of cyclic loading is depicted on
Figure 20 b. For smallest size of RVE (10 grains), the discrepancy is more than 20% for criteria
C(Erat) and C(∆Ep) considered for the asymmetrical stress mode, but the discrepancy is about 2%
for criterion C(Σ11cymax) considered for the symmetrical strain mode. For size 50 grains, the

PT
relative dispersion is 11.97 %, 17.84 %, 1.38 % for C(Erat), C(∆Ep) and C(Σ11cymax),
respectively. Then, for polycrystalline aggregates with the large size (250 grains), the accuracy of

RI
estimation the effective stress Σ11cymax is less than 1%, 4.02% of ratcheting strain and 5.28% for
plastic range deformation.

U SC
AN
M
D
TE

Figure 20. Evolution of criteria as function of RVE size (numbers of grains): (a) monotonic tests
and (b) cyclic tests.
EP

8. Comparison of simulation predictions with experimental results


C

The validation of the proposed approach has been carried out by comparison of simulation results
with corresponding experimental results on 5083 Aluminum alloy from Lu et al. [37]. The
AC

parameters of the crystal plasticity model are identified using the optimization technique.
Concerning the components of the interaction matrix ( hi ), we have considered the values proposed

by Devincre et al. [38] who used discrete dislocation dynamics simulations to quantify these
parameters. At first, the optimal RVE (250 grains) is verified by the simulation of the monotonic
kinematic tensile test, and the resulting findings are shown in Figure 21a where the simulated
stress-strain curve is in good agreement with experimental one. Then, for cyclic test with
symmetrical strain load, Figure 21b shows a good agreement between numerical and experimental

27
ACCEPTED MANUSCRIPT
results for 1st, 10th and 30th cycles. Finally, the ability to predict cyclic behavior under asymmetrical
stress load is shown in Figure 21c where stress-strain hysteresis loops of simulation and
experimental are observed with low gap for the 1st, 10thand 50thcycles.
The predictions of mechanical quantities, which are used to define some new criteria to determine
the optimal RVE size, are compared with corresponding experimental values. The relative error of
estimation of the effective stress Σ11 max is 0.59 % <1%. For Σ11cy max , relative errors are less than 3%

for 1st, 10th and 30th cycle. However, the relative estimation errors of Erat and of ∆E p are relatively

PT
higher: less than 6.5% for Erat for the 1st, 10th and 50th cycles with regards to Erat . For ∆E p , the

RI
error is of 24% , 20 % and 10% for 1st, 10th and 50th cycles, respectively.

U SC
AN
M
D
TE
C EP
AC

Figure 21.Comparison of simulated and experimental results of different configurations.


9. Conclusion
The present study deals with statistical and numerical computation of the RVE size of polycrystal
metals via new criteria. On the basis of a statistical approach combined with a F.E. analysis, the
RVE size of polycrystal materials consisting of grains having a realistic shape (Voronoi polyhedral
28
ACCEPTED MANUSCRIPT
form) is assessed to determine the effective elastoplastic behavior. For both types of behavior,
monotonic and cyclic, new criteria were introduced as indicators of an efficient estimate of the
optimal RVE size. In addition, the accuracy of the calculation of the effective properties of the
elastoplastic behavior is related to the size of the polycrystalline aggregate. The following major
conclusions can be drawn:
• For all criteria used, results show that the scatter decreases with the increase of the RVE size.
However, the accuracy achieved for each size, depends on the proprieties investigated, which

PT
confirm that RVE concept is not unique.
• For smaller sizes (10 and 50 grains), the discrepancy of the ratcheting strain prediction and

RI
the plastic range is more than 20%, while it is about 5%. for other effective properties. For the
largest size (more than 200 grains), the prediction accuracy of all quantities used herein is less

SC
than 3%, except ratcheting strain and plastic range, which still relatively higher than 6%.
• For cyclic loading, the results show that RVE size is highly dependent on the cyclic loading

U
mode. The criteria introduced for asymmetrical stress mode provided a high relative error
than all the other criteria used in this study.
AN
• In both loading cases considered (cyclic and monotonic), it has been shown that RVE size
does not depend on the value of strain or stress used for simulation in the different tests.
M

• For the tensile test and strain-controlled cyclic test, simulation results performed on RVE of
250 grains corroborate experimental results. However, such a size does not seem to be
D

sufficient to predict the elastoplastic behavior during a stress-controlled cyclic test. For this
TE

type loading, the proposed procedure allows to assess the optimal RVE considering a size
greater than 250 grains.
EP

References
C

[1] E. Sanchez–Palencia and A. Zaoui, Homogenization techniques for composite media, Lecture
AC

Notes in Physics, 1985, n272, p193–278, Springer Verlag, New York.


[2] S. Nemat–Nasser and M. Hori, Micromechanics: Overall Properties of Heterogeneous Materials,
1993, v37, 1st Edition, North Holland.
[3] A.Salahouelhadj and H. Haddadi, Estimation of the size of the RVE for isotropic copper
polycrystals by using elastic–plastic finite element homogenization, Computational Materials
Science, 2010, v48 (3), p447–455.
[4] I. Watanable and K. Terada, A method of predicting macroscopic yield strength of
polycrystalline metals subjected to plastic forming by micro-macro de-coupling scheme, 2010,
International Journal of Mechanical Sciences 52, p343-355.

29
ACCEPTED MANUSCRIPT
[5] M. Masmoudi, W. Kaddouri, T. Kanit, S. Madani, S. Ramtani and A. Imad, Modeling of the
effect of the void shape on effective ultimate tensile strength of porous materials: Numerical
homogenization versus experimental results, International Journal of Mechanical Sciences, 2017,
130, p497-507.
[6] R. U. Patil, B. K. Mishra and I. V. Singh, A new multiscale XFEM for the elastic properties
evaluation of heterogeneous materials, International Journal of Mechanical Sciences, 2017,122,
p277-287.
[7] W.J. Drugan and J.R. Willis, A micromechanics–based nonlocal constitutive equation and
estimates of representative volume element size for elastic composites, J. Mech. Phys. Solids,1996,

PT
v44 (4), p497–524.
[8] A.A. Gusev, Representative volume element size for elastic composites: A numerical study, J.

RI
Mech. Phys. Solids, 1997, v45 (9), p1449–1459.
[9] T. Kanit, S. Forest, I. Galliet, V. Mounoury and D. Jeulin, Determination of the size of the
representative volume element for random composites : statistical and numerical approach, Int. J.

SC
Solids Struct., 2003,v40 (13–14), p3647–3679.
[10] M. Ostoja–Starzewski, X. Du, Z.F. Khisaeva and W. Li, Comparisons of the size of

U
representative volume element in elastic, plastic, thermoelastic, and Permeable random
microstructures, Int. J. Multiscale Comp. Eng., 2007, v5 (2), p73–82.
AN
[11] S.M. Mirkhalaf, F.M. Andrade–Pires and R. Simoes, Determination of the size of the
Representative Volume Element (RVE) for the simulation of heterogeneous polymers at finite
strains, Finite Elements in Analysis and Design, 2016, v119, p30–44.
M

[12] B. Koohbor, S. Ravindran and A. Kidane, Experimental determination of Representative


Volume Element (RVE) size in woven composites, Optics and Lasers in Engineering, 2017, v90,
D

p59–71.
[13] Z.Y. Ren and Q.S. Zheng, A Quantitative study of minimum sizes of representative volume
TE

elements of cubic polycrystals – Numerical experiments, J. Mech. Phys. Solids, 2002, v50(4),
p881–893

[14] Z.Y. Ren and Q.S. Zheng, Effects of grain sizes, shapes, and distribution on minimum sizes of
EP

representative volume elements of cubic polycrystals, Mech. Mater., 2004, v36(12), p1217–1229.
[15] M. Nygards, Number of grains necessary to homogenize elastic materials with cubic
C

symmetry, Mech. Mater., 2003, v35 (11), p1049–1057.


AC

[16] S.I. Ranganathan and M. Ostoja–Starzewski, Scaling function, anisotropy and the size of RVE
in elastic random polycrystals, J. Mech. Phys. Solids, 2008a, v56 (9), p2773–2791.
[17] S.I. Ranganathan and M. Ostoja–Starzewski, Scale–dependent homogenization of inelastic
random polycrystals, Journal of Applied Mechanics, Transactions ASME, , 2008b, v75 (5),
p0510081–0510089.
[18] P. D. Chinh, Estimates for the elastic moduli of random orthorhombic polycrystals,
International Journal of Mechanical Sciences, 2004, 46, pp1005-1019.
[19] C. Zhang, M. Enomoto, A. Suzuki and T. Ishimaru, Characterization of three–dimensional
grain structure in polycrystalline iron by serial sectioning, Metal. Mater. Trans. A, 2004, v35 (7),
p1927–1933.
30
ACCEPTED MANUSCRIPT
[20] K.–M. Döbrich, C. Rau and C.–E. Krill, Quantitative characterization of the three–
dimensional microstructure of polycrystalline Al–Sn using X–ray microtomography, Metal. Mater.
Trans. A: Phys. Metall. Mater. Science, 2004, v35 A(7), p1953–1961.
[21] D.J. Rowenhorst, A.C. Lewis and G. Spanos, Three–dimensional analysis of grain topology
and interface curvature in a β –titanium alloy, Acta Mater., 2010, v58 (16), p5511–5519.

[22] F. Barbe, L. Decker, D. Jeulin and G. Cailletaud, Intergranular and intragranular behavior of
polycrystalline aggregates. Part 1: F.E. Model, Int. J. Plast., 2001a, v17 (4), p513–536.
[23] F. Barbe, R. Quey, A. Musienko and G. Cailletaud, Three–dimensional characterization of

PT
strain localization bands in high–resolution elastoplastic polycrystals, Mech. Res. Commun., 2009,
v36 (7), p762–768.

RI
[24] Z. Zhao, S. Kuchnicki, R. Radovitzky and A. Cuitiño, Influence of in–grain mesh resolution on
the prediction of deformation textures in fcc polycrystals by crystal plasticity FEM, Acta Mater.,
2007, v55 (7), p2361–2373.

SC
[25] R. Quey, P.–R. Dawson and F. Barbe, Large–scale 3D random polycrystals for the finite
element method: Generation, meshing and remeshing, Comput. Methods Appl. Mech. Engrg., 2011,
v200 (17–20), p1729–1745.

U
[26] D. Gonzalez, I. Simonovski, P. J. Withers, J. Quinta da Fonseca, Modelling the effect of elastic
AN
and plastic anisotropies on stresses at grain boundaries, Int. J. Plasticity, 2014, 61, p49-63

[27] P. Pilvin, F. Onimus, R. Brenner, S. Pascal, X. Feaugas and K. Sai, 2017, Finite element
assessment of an affine selfe consistent model for Hexagonal polycrystals, European Journal of
M

Mechanics A/ Solids, 2017, 61, 345-356.

[28] E. Héripré, M. Dexet, J. Crépin, L. Gélébart, A. Roos, M. Bornert, D. Caldemaison, Coupling


D

between experimental measurements and polycrystal finite element calculations for


micromechanical study of metallic materials, Int. J. Plas. 2007, 23, 1512-1539.
TE

[29] G. Cailletaud, S. Forest, D. Jeulin, F. Feyel, I. Galliet, V. Mounoury, S. Quilici. Some elemnts
of microstructural mechanics, Comp. Mater. Sci., 2003, 27, 351-347.
EP

[30] L. Méric, P. Poubanne and G. Cailletaud, Single crystal modeling for structural calculations.
Part 1: Model presentation, J. Engng. Mat. Technol., 1991, v113, p162–170.
C

[31] F.Barbe, S. Forest and G. Cailletaud, Intergranular and intragranular behavior of


polycrystalline aggregates. Part 2 : Results, Int. J. Plasticity, 2001b, v17 (4), p537–563.
AC

[32] F.Barbe, S. Forest and G. Cailletaud, Polycrystalline plasticity under small strains. Toward
finer descriptions of microstructures. In : Bouchaud, E. et al. (Eds.), NATO Proceedings, Physical
Aspects of Fracture, Kluwer Academic Publishers, 2001c, p191–206.
[33] O.Diard, S. Leclercq, G. Rousselier and G. Cailletaud, Evaluation of finite element based
analysis of 3D multicrystalline aggregates plasticity : Application to crystal plasticity model
identification and the study of stress and strain fields near grain boundaries, Int. J. Plasticity, 2005,
v21 (4), p691–722.
[34] N. Osipov, A.F. Gourgues–Lorenzon, B. Marini, V. Mounoury, F. Nguyen and G. Cailletaud,
FE modeling of bainitic steels using crystal plasticity, Philos. Mag., Taylor & Francis, 2009, v88
(30–32), p3757–3777.
31
ACCEPTED MANUSCRIPT
[35] K. Sai, G. Cailletaud and S. Forest, Micro–mechanical modeling of the inelastic behavior of
directionally solidified materials, Mech. Mater., 2006, v38 (3), p203–217.
[36] A. Musienko, A. Tatschl, K. Schmidegg, O. Kolednik, R. Pippan and G. Cailletaud, Three–
dimensional finite element simulation of a polycrystalline copper specimen, Acta Mater., 2007, v55
(12), p4121–4136.
[37] F. C. Lu, G. Z. Kang, Y. J. Liu, K. K. Shi, Experimental study on uniaxial cyclic deformation
of rolled 5083Al alloy plate. 7th ICLCF, 2013, Aachen, Germany.

[38] B. Devincre, L. P. Kubin and T. Hoc, Physical analyses of crystal plasticity by DD simulations.

PT
Scripta Materialia, 2006, 54:741–746.

[39] Z.Shi, Q. Gao, G.Z. Kang and Y.J. Liu, Uniaxial time–dependent ratcheting behaviors of

RI
1Cr18Ni9 stainless steel at elevated temperature, Gongcheng Lixue / Eng. Mech., 2007, v24(9),
p159–165.
[40] Y. Sun, S.–L. Shen, X.–H. Xia and Z.–L. Xu, A numerical approach for predicting shakedown

SC
limit in ratcheting behavior of materials, Materials and Design, 2013, v47, p106–114.
[41] S. M. Hamidinejad and A. Varvani-Farahani, Ratcheting of 304 stainless steel under multiaxial

U
step-loading conditions, International Journal of Mechanical Sciences,2015, 100, p80-89.
[42] C. L. Pun, Q. Kan, P. J Mutton, G. Kang and W. Yan, An efficient computational approach to
AN
evaluate the ratcheting performance of rail steels under cyclic rolling contact in service,
International Journal of Mechanical Sciences,2015, 100-102, p214-226.
M

[43] X. Yuan, W. Yu, S. Fu, D. Yu and X. Chen, Effect of mean stress and ratcheting strain on the
low cycle fatigue, Materials Science & Engineering A, 2016, v677, p193–202.
D
TE
C EP
AC

32
ACCEPTED MANUSCRIPT
HIGHLIGHTS

The new proposed criteria allow to identify the optimal size of RVE of virtual
polycrystal material.

The response of crystalline aggregate is not sensitive to the evolution of the grains
orientations and dimensions of identified optimal RVE.

PT
For cyclic behavior, the influence of loading mode is very important, in particular the
asymmetrical stress mode.

RI
The simulation results present a good agreement with the experiment results.

U SC
AN
M
D
TE
C EP
AC

You might also like