You are on page 1of 19

JID: SAS

ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

International Journal of Solids and Structures 0 0 0 (2016) 1–19

Contents lists available at ScienceDirect

International Journal of Solids and Structures


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

A methodology for the estimation of the effective yield function of


isotropic composites
I. Papadioti a,b, K. Danas c, N. Aravas a,d,∗
a
Department of Mechanical Engineering, University of Thessaly, 38334 Volos, Greece
b
Institute for Research and Technology Thessaly (I.RE.TE.TH.), Center for Research and Technology - Hellas, 38333 Volos, Greece
c
Laboratoire de Mécanique des Solides, C.N.R.S., École Polytechnique, University of Paris-Saclay, Palaiseau, France
d
International Institute for Carbon Neutral Energy Research (WPI-I2CNER), Kyushu University, 744 Moto-oka, Nishi-ku, Fukuoka 819-0395, Japan

a r t i c l e i n f o a b s t r a c t

Article history: In this work we derive a general model for N−phase isotropic, incompressible, rate-independent elasto-
Received 4 November 2015 plastic materials at finite strains. The model is based on the nonlinear homogenization variational (or
Revised 8 February 2016
modified secant) method which makes use of a linear comparison composite (LCC) material to estimate
Available online xxx
the effective flow stress of the nonlinear composite material. The homogenization approach leads to an
Keywords: optimization problem which needs to be solved numerically for the general case of a N−phase composite.
Homogenization In the special case of a two-phase composite an analytical result is obtained for the effective flow stress
Elasto-plasticity of the elasto-plastic composite material. Next, the model is validated by periodic three-dimensional unit
Composite materials cell calculations comprising a large number of spherical inclusions (of various sizes and of two differ-
Finite strains ent types) distributed randomly in a matrix phase. We find that the use of the lower Hashin–Shtrikman
bound for the LCC gives the best predictions by comparison with the unit cell calculations for both the
macroscopic stress-strain response as well as for the average strains in each of the phases. The formu-
lation is subsequently extended to include hardening of the different phases. Interestingly, the model is
found to be in excellent agreement even in the case where each of the phases follows a rather different
hardening response.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction phases/particles tend to reinforce the yield strength of the com-


posite while they usually have different strength and hardening
The present work deals with the analytical and numerical es- behavior than the host matrix phase.
timation of the effective as well as the phase average response of In the literature of nonlinear homogenization there exists a
N−phase incompressible isotropic elasto-plastic metallic compos- large number of studies for two-phase composite materials. The
ites. Special attention is given to particulate microstructures, i.e., reader is referred to Ponte Castañeda and Suquet (1998), Ponte
composite materials which can be considered to comprise a dis- Castañeda (2002), Idiart et al. (2006), and Idiart (2008) for a re-
tinct matrix phase and an isotropic distribution of spherical parti- view of the nonlinear homogenization schemes such as the ones
cles (Willis et al., 1982) (or in a more general setting an isotropic used in the present work and relevant estimates. Nonetheless, very
distribution of phases (Willis, 1977)). In the present study, the par- few studies exist in the context of three- or N−phase rate indepen-
ticles are considered to be stiffer than the matrix phase, which dent elasto plastic composites.
is the case in most metallic materials of interest, such as TRIP In view of this, the present work uses the nonlinear variational
steels, dual phase steels, aluminum alloys and others. Such ma- homogenization method (Ponte Castañeda, 1991) or equivalently
terials, usually contain second-phase particles (e.g., intermetallics, the modified secant method (Suquet, 1995), which makes use of
carbon particles) or just second and third phase variants (e.g., re- a linear comparison composite (LCC) material, to estimate the ef-
tained austenite, bainite, martensitic phases). In addition, these fective response of a N−phase nonlinear composite material. Even
though, this method exists for several years most of the studies

in the context of composite materials have been focused on two-
Corresponding author at: Department of Mechanical Engineering, University of
Thessaly, 38334 Volos, Greece. Tel.: +30 2421074002; fax: +30 2421074009.
phase composites where the optimization process required by the
E-mail addresses: kdanas@lms.polytechnique.fr (K. Danas), aravas@uth.gr, method can be done analytically (see for instance deBotton and
aravasn@gmail.com (N. Aravas). Ponte Castañeda (1993)). Nevertheless, as the number of phases

http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
0020-7683/© 2016 Elsevier Ltd. All rights reserved.

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

2 I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19

increases to three or more the optimization can only be done nu- composite flow stress becomes
merically. Perhaps, that is the reason that in his original work, ⎧ (2 ) (1 ) √
Ponte Castañeda (1992) proposed general expressions (and bounds) ⎨ 5 c r+c 9+6(2c) (2) −6 c(2) r2 if 1 ≤ r ≤ 5 4 + 6 c (2 ) ,
σ˜ 0 3+2 c
for N−phase composites, but its numerical/analytical resolution re- =
mained untractable until today due to the complex optimization σ0(1) ⎩ 1 4 + 6 c (2 ) if r ≥ 5 4 + 6 c (2 ) ,
2
procedures required by the nonlinear homogenization method.
(2)
It should be pointed out at this point that these homogeniza-
(2 ) (1 )
tion theories treat separately the elastic (which in the present where r = σ0 /σ0 is the contrast ratio. The predictions of the
case is trivial) and the plastic homogenization problem. That of homogenization model agree well with the predictions of detailed
course has certain impact if cyclic loading is considered which is three-dimensional unit cell finite element calculations as shown in
beyond the scope of the present work and is not considered here. the following.
Nevertheless, recently, Lahellec and Suquet (2007) proposed an The homogenization technique provides also accurate estimates
incremental variational formulation for materials with a hereditary for the average strains in the constituent phases. These estimates
behavior described by two potentials: a free energy and a dissi- form the basis for the development of an approximate analytical
pation function. This method has been introduced mainly to deal model for the elastoplastic behavior of a composite with harden-
with the coupled elasto-plastic response of composites in an at- ing phases. A method for the numerical integration of the result-
tempt to resolve the cyclic response of these materials (see also re- ing elastic-plastic equations is developed and the model is imple-
cent work by Brassart et al. (2011)). Note that these more advanced mented into the ABAQUS general purpose finite element code. The
methods use the aforementioned or variants of the LCC estimates. predictions of the model agree well with the results of detailed
In this regard, the present study, albeit not using this coupled unit cell finite element calculations of a composite with hardening
scheme, reveals the nature of equations required to deal with phases.
a general N−phase composite material and could be potentially Standard notation is used throughout. Boldface symbols denote
useful in the future for such more complete incremental schemes, tensors the orders of which are indicated by the context. The usual
which are based upon those simpler LCC homogenization theories. summation convention is used for repeated Latin indices of ten-
sor components with respect to a fixed Cartesian coordinate sys-
tem with base vectors ei (i = 1, 2, 3 ). The prefice det indicates
1.1. Scope of the present work and major results
the determinant, a superscript T indicates the transpose, and the
subscripts s and a the symmetric and anti-symmetric parts of a
The scope of the present work is to provide a semi-analytical
second-order tensor. A superposed dot denotes the material time
model for N−phase isotropic, incompressible rate-independent
derivative. Let A, B be second-order tensors, and C, D fourth-order
elasto-plastic materials. Simple analytical expressions are given
tensors; the following products are used in the text: (A · B )i j =
for the effective yield stress of a two-phase composite (see also
Aik Bk j , A : B = Ai j Bi j , (A B )i jkl = Ai j Bkl , (C : A )i j = Ci jkl Akl , and (C :
(deBotton and Ponte Castañeda, 1993)), while a simple semi-
D )i jkl = Ci j pq D pqkl . The inverse C−1 of a fourth-order tensor C that
analytical expression (requiring the solution of a constrained op-
has the “minor” symmetries Ci jkl = C jikl = Ci jlk is defined so that
timization problem for N − 1 scalar quantities) is given for the
C : C−1 = C−1 : C = I , where I is the symmetric fourth-order iden-
N−phase composite. Additional analytical expressions are also pro-
tity tensor with Cartesian components Ii jkl = (δik δ jl + δil δ jk )/2, δ ij
vided for the phase concentration tensors and average strains in
being the Kronecker delta.
each phase in terms of the aforementioned optimized scalar quan-
tities. In the context of two- and three-phase materials the model
2. Power-law creep and perfect plasticity
is assessed by appropriate three-dimensional multi-particle two-
and three-phase periodic unit cell calculations considering both
We consider an incompressible creeping solid characterized by
hardening and non-hardening phases. The agreement is found to
a power-law stress potential U of the form
be good not only for the effective yield stress but also for the
phase average strains thus allowing for the extension of this model σ0 ε˙ 0 σe n+1
U ( σe ) = , (3)
to include arbitrary isotropic hardening of the phases. n + 1 σ0
Specifically, we use the methodology developed by Ponte Cas-
where σ 0 is a reference stress, ε˙ 0 
a reference strain rate, n the
tañeda and co-workers (Ponte Castañeda, 1991; Suquet, 1995) to
derive a model for the rate-independent elastoplastic behavior of creep exponent (1 ≤ n ≤ ∞), σe = 32 s : s the von Mises equiv-
a macroscopically isotropic composite comprising N phases. When alent stress, σ the stress tensor, p = σkk /3 the hydrostatic stress,
the constituent phases are perfectly plastic the corresponding flow and s = σ − p δ the stress deviator, δ being the second-order iden-
stress of the composite material σ˜ 0 is determined from the solu- tity tensor. The corresponding deformation rate D is defined as
tion of a constrained optimization problem:
∂U σ n ∂σe 3
    −1 D= = ε̄˙ N, ε̄˙ = ε˙ 0
e
, N= = s, (4)
 ∂σ σ0 ∂σ 2 σe
 N N
c ( p) N
c (s ) y (s )
σ˜ 0 = 
2
(r ) σ (r ) y (r ) .
 (i )
inf c
 y ≥0 r=1
0
p=1
3 y ( p ) + 2 y0 s=1
3 y ( s ) + 2 y0  magnitude (N : N =
where N is a second order tensor of constant
y (1 ) =1
i=2,...,N 2)
3
that defines the direction of D and ε̄˙ = 2
3 D : D is the equiva-
(1) lent plastic strain rate that defines the magnitude of D. Note that
Dkk = 0.
where N is the number of phases, (c (i ) , σ0(i ) ) are the volume frac- The special case in which the exponent takes the value of unity
tion and flow stress of phase i, and y(i) are positive optimization (n = 1) corresponds to a linearly viscous solid:
parameters. In turn, y0 is a reference scalar to be chosen accord-
σe2 ∂ UL s
ing to various linear homogenization schemes. For instance, best UL (σe ) = , D= = , (5)
results are obtained with the well known Hashin–Shtrikman lower 6μ ∂σ 2μ
bound choice, i.e., y0 = y(1 ) = 1. where μ = σ0 /(3 ε˙ 0 ) is the viscosity.
In the special case of a two-phase composite (N = 2 ), the op- The other limiting case n → ∞ corresponds to a perfectly plas-
timization problem is solved analytically and the estimate for the tic solid that obeys the von Mises yield condition with flow stress

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19 3

σ 0 . In this case the stress function (3) becomes1 estimate (10) of U˜ may have the character of a rigorous bound pro-
 vided that the corresponding estimate U˜L has also the same charac-
0 when σe ≤ σ0 ,
U∞ (σe ) = (6) ter of a bound as discussed in the following. Nonetheless the scope
∞ when σe > σ0 . of the present work is to insist mainly on a good estimate by com-
The threshold stress σ 0 in (6) is the flow stress of the material, parison with numerical unit cell calculations and not necessarily
and the flow rule is written in the form on rigorous bounds.
In this view, the quadratic potential U˜L of the LCC in (10 b) uses
3
D = ε̄˙ N, N= s, (with ε̄˙ = 0 if σe < σ0 ), (7) the effective viscosity μ
˜ of the LCC that depends on the individual
2 σe viscosities μ(r) and the corresponding volume fractions c(r) . One
where the equivalent plastic strain rate ε̄˙ is not defined locally by way to estimate μ ˜ is to use the well-known Hashin–Shtrikman
the constitutive equations and becomes one of the primary un- relationship for particulate composites (e.g., see Willis et al.
knowns in the rate boundary value problem. (1982))
  −1
3. The homogenization method  (r )
 
N
c (s ) μ (s ) 
N
c (r )
μ˜ μ = , (13)
3 μ0 + 2 μ ( s ) 3 μ0 + 2 μ ( r )
We consider a composite material made of N isotropic, incom- s=1 r=1
pressible viscoplastic phases. The phases are distributed randomly
and isotropically and are characterized by viscoplastic stress poten- where μ0 is a “reference viscosity” to be chosen appropriately. An
tials U(r) of the form (3) with constants (σ0(r ) , ε˙ 0 , n(r ) ) and μ(r) in upper bound for μ ˜ is produced by (13) when μ0 is chosen to be
the linear case, i.e., the maximum of all μ(r) and a lower bound is produced when μ0
 n(r) +1 is the minimum of all μ(r) (Willis et al., 1982).
 (r )
 σ0(r ) ε˙ 0 σe(r ) σe(r )
2
An important observation made by several authors is that the
U (r )
σe = , UL(r ) (σe(r ) ) = , (8)
n (r ) + 1 σ0(r ) 6 μ (r ) Hashin–Shtrikman bounds are accurate estimates for composites
with particulate microstructures, at least for two-phase systems at
where σe(r ) is the von Mises equivalent stress in phase r. The vol- moderate volume fraction (Bonnenfant et al., 1998); in particular,
 ( r ) = 1 ).
ume fraction of each phase is c(r) ( N r=1 c the upper bound is a good estimate when the stiffest material is
The constitutive equation of the isotropic nonlinear compos- the matrix phase and contains inclusions of the most compliant
ite is written in terms of the effective viscoplastic stress potential material, whereas the lower bound is a good estimate for the in-
U˜ (σ ), so that verse situation in which the most compliant material is the matrix
phase containing inclusions of the stiffest material.
∂ U˜
D= , (9) When no phase plays clearly the role of a matrix, the effec-
∂σ tive properties of the composite may be estimated by the self-
where σ and D are respectively the macroscopic stress and defor- consistent method of Hill (1965). In this case, the relevant mi-
mation rate in the composite. crostructure is granular in character, being composed of ellipsoidal
An estimate for U˜ is obtained by using the variational method- particles of the different phases with varying size so as to fill
ology of Ponte Castañeda and co-workers (Ponte Castañeda (1991), space. Eq. (13) provides Hill’s self-consistent estimate, if μ0 is
Ponte Castañeda and Suquet (1997), Ponte Castañeda and Suquet identified with the effective modulus μ ˜ ; in this case, (13) becomes
(1998)). This methodology has also been proposed independently a polynomial equation of order 2 N for μ ˜ Willis et al. (1982).
for power-law materials by Michel and Suquet (1992) and inter-
preted as a secant homogenization method by Suquet (1995). The
final form of the estimate reads (Ponte Castañeda, 1992)
  3.1. Strain-rate concentration in the phases
  
N  
U˜ (σe ) = sup U˜L σe , μ˜ (μ(r ) ) − c (r ) v (r )
μ (r )
, An approximation for the strain field in the non-linear compos-
μ(r ) ≥0 r=1 ite may be obtained from the strain field in the LCC evaluated at
σe2 the optimal comparison moduli μ ˆ (r ) defined by the optimization
U˜L = , (10) problem in (10). In particular, the average deformation rate field in

˜ ( μ (r ) )
the phases D(r) may be written in terms of the macroscopic defor-
where σ e is the macroscopic von Mises equivalent stress, mation rate D in the form (Kailasam and Ponte Castañeda, 1998;
      
v(r ) μ(r ) = sup UL(r ) σe(r ) , μ(r ) − U (r ) σe(r ) , (11) Ponte Castañeda, 2005; Ponte Castañeda and Suquet, 1997; Ponte
σe(r ) ≥0 Castañeda and Zaidman, 1994):
2  n(r) +1  
σe(r ) σ (r ) ε˙ 0 σe(r ) D (r ) = A (r ) μ
ˆ ( i ) ( σe ) : D ,
UL(r ) = , U (r ) = (0r ) . (12) r = 1, 2, . . . , N, (14)
6μ ( r ) n +1 σ0(r )
where A(r) are the fourth-order strain concentration tensors of
The effective stress potential U˜ (σ ) is defined in (10) in terms of
the LCC, evaluated at the optimal values2 , μˆ (r ) , of the compari-
the quadratic effective stress potential U˜L of a “linear comparison
son moduli, defined by the solution of the optimization problem
composite” (LCC) evaluated at the macroscopic stress σ e and the
in (10). It is emphasized that the optimal values μ ˆ (r ) depend in a
“corrector functions” v(r) , which are defined in (11) as the opti-
nonlinear manner upon the macroscopic von Mises equivalent σ e ,
mal difference between the quadratic potentials UL(r ) and the ac-
and consequently the strain concentration tensors A(r) are in gen-
tual potentials of the non-linear materials U(r) . The stress tensors
eral nonlinear functions of the macroscopic stress tensor σ .
σe(r ) in (11) are obtained by the “sup” operation in that equation
and hence v(r) are only functions of the individual viscosities of
the linearized phases, μ(r) . It is worth noting at this point that the
2
Henceforth the superscript (ˆ. ) serves to denote the optimal value of the rele-
0 vant quantity obtained by the corresponding optimization described in the previous
An+1 if A ≤ 1,
1
Here we take into account that lim = ∞ A > 1.
section.
n→∞ n+1 if

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

4 I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19

For isotropic composite materials with random microstructures 3.2. Perfectly plastic phases
and “ellipsoidal symmetry”, A(r) is of the form (Ponte Castañeda
and Suquet, 1997) We consider the case of perfectly plastic phases (n(r) → ∞). The
 −1 optimization in (10) and (11) as n(r) → ∞ is carried out in three
 steps. In the first step, we consider the optimization over σe(r ) in
N
A (r ) = E (r ) : c (s ) E (s ) ,
(11). All creep exponents are set equal in the second step, i.e., we
s=1
set n(1 ) = n(2 ) = · · · = n(N ) ≡ n. In the final third step we consider
 (r )
 −1
E(r ) = I + S0 : L−1
0 : L − L0 , (15) the limit n → ∞. Details of the calculations are given in the fol-
lowing.
where I is the symmetric fourth order identity tensor with Carte-
sian components Ii jkl = (δik δ jl + δil δ jk )/2, S0 is the well known Step 1: Calculation of σe(r ) in (11)
tensor of Eshelby (1957) for the linear “reference material” with The “inner” optimization in (11) is carried out by setting equal
elasticity tensor L0 introduced in (13), and to zero the derivatives

L0 = 2 μ0 K + 3 κ0 J , L(r ) = 2 μ
ˆ (r ) K + 3 κ (r ) J , ∂  (r ) 
(r )
UL − U (r ) = 0, (22)
1 ∂σe
J = δ δ, K = I − J . (16)
3
which defines the optimal values of σe(r ) as
The quantities (μ0 , κ 0 ) and (μˆ (r ) , κ (r ) ) in (16) are the shear and
bulk viscosities of the LCC; the bulk viscosities κ 0 and κ (r) are set ⎡
 (r) ⎤ n(r) −1
1

(r ) n
to ∞ after the final expression for D(r) in (14) is derived, in order σ
σe(r ) = ⎣ 0 (r ) ⎦ ≡ σˆ e(r ) . (23)
to take into account the incompressible nature of the phases and 3 μ ε˙ 0
the composite.
For composites consisting of an isotropic matrix and a uniform
distribution of spherical inclusions, the Eshelby tensor has the form When the optimal values σˆ e(r ) are substituted into (10), the expres-
sion for the estimate of the effective stress potential becomes
6(κ0 + 2 μ0 ) 3 κ0 ⎧ ⎡
S0 = K+ J. (17) ⎪  (r ) ⎤ n(r ) −1
2

5(3 κ0 + 4 μ0 ) 3 κ0 + 4 μ0 ⎨  σ (r ) n
σ 2
1
N
n (r ) − 1
⎣ 0 ⎦
Using (16) and (17) in (15 b) and taking into account that J : J = U˜ (σe )= sup e  −
μ(r ) ≥0⎪
⎩ 6 μ˜ μ(r ) 2 n (r ) + 1 ε˙ 0
J , K : K = K, and J : K = 0, we conclude that r=1
⎫ (24)
5 μ0 (3 κ0 + 4 μ0 ) 3 κ0 + 4 μ0 (r ) ⎬
E (r ) = K+ J. c
μ0 (9 κ0 + 8 μ0 ) + 6(κ0 + 2 μ0 )μˆ (r ) 3 κ ( r ) + 4 μ0   n((rr )) +1 ⎭,
(18) 3 μ (r ) n −1

Then, using (15), after some lengthy but otherwise straightforward where μ ˜ (μ(r ) ) is defined in (13). Substitution of the expression
calculations we reach the following expression for the strain con- (13) for μ
˜ into (24) leads to
centration tensors: $ %
 −1   σe2  (1 ) (r ) 
1 
N
c (s ) U˜ (σe ) = sup sup F y (r ) − I μ , y , (25)
A (r ) = J y (r ) ≥0 μ ( 1 ) >0 6 μ (1 )
3 κ (r ) + 4 μ0 s=1
3 κ (s ) + 4 μ0 y (1 ) =1

1 where
+   −1
μ0 (9 κ0 + 8 μ0 ) + 6(κ0 + 2 μ0 )μˆ (r )
 −1  (r )
 μ (1 ) N
c (r ) y (r ) N
c (s )
 F y = = , (26)
c (s )
N
μ˜ 3 y ( r ) + 2 y0 3 y ( s ) + 2 y0
K. (19) r=1 s=1
s=1
μ0 (9 κ0 + 8 μ0 ) + 6(κ0 + 2 μ0 )μˆ (s)
⎡ n(r) ⎤ n(r) −1 
2
Finally, using (14), taking into account the incompressibility condi-  n((rr)) +1
tion Dkk = 0 (or J : D = 0), and considering the limit κ 0 → ∞, we
  1  (r ) n (r ) − 1
N
σ0(r ) y (r ) n −1
I μ (1 ) , y (r )
= c ⎣ ⎦ ,
find 2 n (r ) + 1 ε˙ 0 3μ ( 1 )
  r=1
D(r ) = lim A ( r ) : D = α ( r ) D, (27)
κ0 →∞
 −1
1 
N
c (s ) μ (1 ) μ (1 )
α (r ) = . (20) y (r ) = (with y(1) = 1 ), and y0 = . (28)
3 μ0 + 2 μ
ˆ (r ) 3 μ0 + 2 μ ˆ (s ) μ (r ) μ0
s=1

The optimal values of y(r) in (25) depend on the values of the vol-
We emphasize again that the strain concentration factors α (r) depend
ume fractions c(r) , the material properties (σ0(r ) , n(r ) , ε˙ 0 ), and the
in general on the macroscopic stress σ (or macroscopic deformation
macroscopic von Mises equivalent stress σ e .
rate D) through the optimal moduli μˆ (i ) . Eq. (20) implies that
The strain concentration values α (r) defined in (20) can be writ-
  ten in the form
2 (r ) 2 dε̄ (r )
ε̄˙ (r ) = D : D (r ) = α (r ) D : D = α (r ) ε̄˙ or = α (r ) ,  −1
3 3 dε̄ 
yˆ(r ) c(s ) yˆ(s )
N
(r )
(21) α = , r = 2, . . . , N, (29)
3 yˆ(r ) + 2 y0 s=1
3 yˆ(s ) + 2 y0
where ε̄˙ (r ) and ε̄˙ are the average equivalent strain rates in the
phases and the average macroscopic equivalent strain rate respec- where yˆ(r ) are the optimal values of y(r) resulting from the opti-
tively. mization in (25).

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19 5

 
Step 2: Equal creep exponents n(1 ) = n(2 ) = · · · = n(N ) ≡ n with
When all “creep exponents” are set equal, i.e., n(1 ) = n(2 ) =   

· · · = n(N ) ≡ n, Eq. (25) becomes  H∞ y(r )
 σ˜ 0 =  inf   , r = 2, . . . , N
   (r ) (r )
≥0 F y
(38)
σe2  (r ) 
y
n−1 H y (r ) y (1 ) =1
U˜ = sup sup F y −  n+1 ,
2 (n + 1 ) 
(30)
6 μ (1 )
y (r ) ≥0
y (1 ) =1
μ ( 1 ) >0 3 μ(1) n−1 where H∞ (y(r) ) and F(y(r) ) are defined in (36) and (26) respectively,
i.e.,
where F(y(r) ) is defined in (26) and


    −1
 N N ( p) N (s ) y (s )
(r ) (r ) 
σ˜ 0 (c , σ0 ) =  (inf ( ) 2
c ( r ) σ0 y ( r )
r c c
. (39)
y i ) ≥0
r=1 p=1
3 y ( p ) + 2 y0 s=1
3 y ( s ) + 2 y0
(1 ) y =1
i=2,...,N

 n  n−1
2
Comparing the above Eq. (37) with (6), we conclude that, when all
  
N
σ0(r )   nn+1
H y (r )
= c (r ) y (r ) −1
. (31) phases are perfectly plastic (n = ∞ ), the form of the estimated ef-
r=1
ε˙ 0 fective stress potential U˜ (σe ) corresponds to a perfectly plastic ma-
terial that obeys the von Mises yield condition with a flow stress
The optimal value of μ(1) in (30) is determined by calculating the σ˜ 0 defined in (39). This effective flow stress, in turn, is a func-
partial derivative of the quantity in brackets with respect to μ(1) tion of the phase volume fractions c(r) as well as of the phase flow
and setting it equal to zero. The resulting value for μ(1) is stresses σ0(r ) .
    n−1 Calculation of the estimated effective yield stress σ˜ 0 requires
H y (r ) 1  
2
1 the solution of the constrained optimization problem in (39) for
μ (1 ) =   ˆ (1 ) y (r ) > 0
≡μ (32)
3 F y(r ) σe2 the values of y(r) , which define in turn the appropriate values of
the viscosities μ(r) (see (28)). In the special case of a two-phase
and (30) becomes composite the solution of the optimization problem in (39) can
be found analytically as described in Section 3.2.1. The solution of
 ⎡  (r )  ⎤ 2
n+1

n+1 
more general cases presented in the following are obtained by us-
σ  [ F ( y (r ) )]n+1 σ n+1 F y
U˜ (σe ) = e  sup = e ⎣ sup   n−1 ⎦ .
ing the methodology of Kaufman et al. (1995) and the CONMAX
n + 1 y(r ) ≥0 [H (y(r ) )]n−1 n + 1 y(r ) ≥0 (r ) n+1 software (http://www.netlib.org/opt/conmax.f) for the solution of
y (1 ) =1 y (1 ) =1
H y
the optimization problem in (39).
(33) The optimal values y(r) in (39) depend on the values of the
volume fractions c(r) and the flow stresses σ0(r ) of the phases but
It is interesting to note that the expression for the effective stress are independent of the macroscopic stress state. Also, depending on
potential given in (33) is of the power-law type defined in (3), i.e., the parameters of the problem, the optimal values yˆ(r ) = μ ˆ ( 1 ) /μ
ˆ (r )
when all phases have the same creep exponent n, the effective be- may be one of the extreme values 0 or ∞. The value yˆ(r ) = 0
havior of the composite is also of the power-law type with creep corresponds to a rigid comparison material for phase r, whereas
exponent n implying that U˜ is a homogeneous function of degree yˆ(r ) = ∞ corresponds to an incompressible comparison material
n + 1 in σ . Also, the optimal values of y(r) in (33) are now indepen- with zero stiffness (i.e., to an “incompressible void” comparison
dent of the macroscopic von Mises equivalent stress σ e . material). It should be noted that it is possible to have yˆ(r ) =
Step 3: Perfectly plastic phases (n → ∞) μˆ (1) /μˆ (r ) = 0 (rigid comparison material) even for finite σ0(r ) (e.g.,
Using (33) and taking into account that see deBotton and Ponte Castañeda (1993) and Section 3.2.1 below).
n+1
 The strain concentration values α (r) defined in (20) can be writ-
[a ( n ) ] 0 when a(∞ ) ≤ 1, ten in the form
lim = (34)
n→∞ n+1 ∞ when a(∞ ) > 1 ,
 −1
yˆ(r ) 
N
c(s ) yˆ(s )
we find α (r ) = (r ) ,
⎧    3 yˆ + 2 y0 3 yˆ(s ) + 2 y0
(40)
⎪  F y (r )
s=1

⎪ 

⎪0 when σe  sup   ≤ 1,

⎪ (r ) H∞ y where yˆ(r ) are the optimal values of y(r) resulting from the opti-


y (r ) ≥0
y (1 ) =1
mization in (39).
lim U˜ (σe ) =  (35)
n→∞ ⎪
⎪   

⎪  F y (r ) 3.2.1. The two-phase perfectly plastic composite—An analytic estimate

⎪∞ when σe  sup   > 1,

⎪ (r ) for the effective flow stress and the strain concentration factors
⎩ y (r ) ≥0
y (1 ) =1
H∞ y
We consider an isotropic two-phase composite (N = 2, c1 + c2 =
1 ). Each phase is perfectly plastic with flow stress σ0(1 ) and σ0(2 ) .
where F(y(r) ) is defined in (26) and We treat phase 1 as the “matrix” and phase 2 with σ0(2 ) > σ0(1 ) as
    
N
2
the reinforcing particles. In that case it is possible to obtain ana-
H∞ y(r ) = lim H y(r ) = c(r ) σ0(r ) y(r ) . (36) lytical expressions for the effective flow stress σ˜ 0 .
n→∞
r=1 The estimate for σ˜ 0 depends on the chosen value of the ref-
Eq. (35) can be written also as erence viscosity μ0 in (13). Results for various choices of μ0 are
 reported in Papadioti and will be discussed briefly later in this
0 when σe ≤ σ˜ 0 , section. Here we present in some detail the formulation based on
lim U˜ (σe ) = (37)
n→∞ ∞ when σe > σ˜ 0 , a Hashin–Shtrikman lower bound with μ0 = μ(1 ) (y0 = 1 ); as it
will be discussed in the following Section 4, this particular choice

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

6 I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19

of μ0 shows the best agreement with detailed unit cell finite ele- where
ment calculations. For μ0 = μ(1 ) , the ratio H∞ /F in (38) takes the
value c (1 ) c(2) yˆ(2)
D= + (47)
  2 y0 + 3 2 y0 + 3 yˆ(2)
H∞ y(2) 2  2 + 3 c (2 ) + 3 c (1 ) y (2 )
  = σ0(1) c(1) + c(2) r2 y(2)   ,
F y (2 ) (1 ) 2c (2 ) (2 )
+ 3 + 2c y and yˆ(2 ) is defined in (43).

σ0(2) 4. Unit cell finite element calculations and assessment of the


r= > 1. (41)
σ0(1) models

The optimum value of y(2) to be used in (38) is calculated by using In this section we present the results of unit cell finite element
the condition calculations for a composite material made up of a statistically
∂ H∞  isotropic random distribution of isotropic, linearly-elastic perfectly-
=0 (42)
∂ y (2 ) F plastic spherical inclusions embedded in a continuous, isotropic,
linearly-elastic perfectly-plastic matrix. The elastic Young modulus
together with the constraint y(2) ≥ 0. After some lengthy, but used in the finite element calculations for all phases is three orders
straightforward, calculations we find the resulting optimal value of magnitude higher than the highest yield stress involved; this
yˆ(2 ) to be
⎧   

⎨ 1 5  1  
−2 c(1) + √ 3 + 2 c (2 ) − 2 c (2 ) if 1 ≤ r ≤ r cr c(2) ≤ c(cr
2)
,
yˆ(2) = 3 + 2 c (2 ) 3 r2 (43)

⎩  
(2 )
0 if r ≥ rcr c (2 ) ≥ c cr ,

where minimizes the effects of elasticity and the results are very close
$ 2 % to those of rigid-perfectly-plastic materials.
5 1 5
r cr = and c(cr
2)
= −4 . (44) We study numerically two- and three-phase composites. The
4 + 6 c (2 ) 6 r matrix is labelled as phase 1 and the reinforcing particles are
spherical and have higher flow stresses (σ0(i ) > σ0(1 ) , i > 1 ).
According to (43), for a given particle concentration c(2) , when the The periodic unit cell is a cube with edge size L and is con-
contrast ratio r = σ0(2 ) /σ0(1 ) is larger than a value r cr , the compar- structed using the method presented by Segurado and Llorca
ison material for phase 2 (particles) is rigid (yˆ(2 ) = 0 ). (2002) (see also Fritzen et al., 2012) and extended to polydisperse
The corresponding estimate for the effective flow stress result- inclusion distributions by Lopez-Pamies et al. (2013). The virtual
ing from (38) is microstructure contains a dispersion of a sufficiently large number

1
  
⎨ 5 c (2 ) r + c (1 ) 9 + 6 c (2 ) ( 1 − r 2 ) if 1 ≤ r ≤ rcr c(2) ≤ c(cr
2)
,
σ˜ 0 3 + 2 c (2 )
= (45)
σ0(1) ⎩ 1 4 + 6 c(2)  
if r cr ≤ r c(cr
2)
≤ c (2 ) .
2

The result stated in (45) was first presented by deBotton and Ponte
Castañeda (1993), who used a “dissipation function” formulation of non-overlapping spheres of uniform (monodisperse) or differ-
(as opposed to the “stress potential” approach used here). For all ent (polydisperse) size. The inclusions are randomly located within
volume fractions c(2) , there is a value r cr of the contrast ratio the cell and are generated using the Random Sequential Adsorp-
r = σ0(2 ) /σ0(1 ) beyond which the predicted effective flow stress σ˜ 0 tion Algorithm (RSA) (Rintoul and Torquato, 1997). In addition,
does not vary with r. For values of r larger than r cr , the optimal the unit cell is periodic, i.e., it can be repeated in all three di-
value of y(2 ) = μ(1 ) /μ(2 ) vanishes or μ(2 ) = ∞, i.e., for r ≥ r cr the rections to represent a 3-D periodic structure. For the two-phase
comparison material 2 (particles) does not deform; therefore, fur- composite and for c(2) ≤ 0.20 monodisperse spheres are used; for
ther increase of σ0(2 ) does not change the effective flow stress σ˜ 0 . higher volume fractions polydisperse (variable size) distributions
The estimate for the effective flow stress σ˜ 0 depends on the are used. In the present study, the two-phase polydisperse ap-
choice of the reference viscosity μ0 . Fig. 1 shows the predicted proach of Lopez-Pamies et al. (2013) is readily extended to obtain
σ˜ 0 for various choices of μ0 for a volume fraction c(2) = 0.30. virtual microstructures with three-phases or more. For instance,
The curves marked H-S− and H-S+ correspond to μ0 = μ(1 ) and denoting the matrix phase with 1 and the two inclusion phases
μ0 = μ(2) respectively, and “self consistent” corresponds to μ0 = with 2 and 3, the extension is straightforward and requires the
μ˜ . We emphasize that the Hashin–Shtrikman lower bound H-S− continuous alternation of spheres of phase 2 and spheres of phase
(μ0 = μ(1) ) shows the best agreement with detailed unit cell fi- 3 during the RSA process. Of course this simple extension can be
nite element calculations presented in the following section. repeated as often as necessary to obtain an N−phase virtual mi-
The strain concentration values α (r) given in (29) can be written crostructure provided that the concentration of each of the phases
in the form is known. Moreover, as discussed briefly in the following, a conver-
gence study with respect to the number of spheres is done for all
dε̄ (1) 1 virtual microstructures used in this study to ensure isotropy and
α (1 ) = = ,
dε̄ (2 y0 + 3 )D ergodicity of the virtual unit cell.
Finite element discretizations of the cubic unit cell were cre-
dε̄ (2) yˆ(2) ated from the particle center distributions using the mesh gen-
α (2 ) = =   , (46) erator code NETGEN (Schöberl, 1997), which has the capability
dε̄ 2 y0 + 3 yˆ(2) D
to create periodic meshes as required. All calculations were car-
ried out using the ABAQUS general purpose finite element code

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19 7

Fig. 1. Variation of effective normalized flow stress σ˜ 0 /σ0(1) with contrast ratio r = σ0(2) /σ0(1) as predicted by various models for a volume fraction c (2) = 0.30.

(Hibbitt, 1977). Three dimensional 10-node quadratic tetrahe- 5 of the unit cell in the form
dral elements with a constant pressure interpolation were used
(C3D10H in ABAQUS); all analyses were carried out incrementally ui(2) = (Fi1 − δi1 ) L, ui(4) = (Fi2 − δi2 ) L, ui(5) = (Fi3 − δi3 ) L,
and accounted for geometry changes due to deformation (finite (48)
strain solutions). where Fij are the components of the macroscopic deformation gra-
Fig. 2 shows the finite element meshes used for a two-phase dient F. The periodicity of the problem requires also that the dis-
composite with volume fractions c (2 ) =0.10, 0.20, 0.30, and 0.40. placements of material points at the same position on opposite
The distributions are monodisperse for c (2 ) =0.10 and 0.20, and faces of the cell should satisfy the conditions
polydisperse for c (2 ) =0.30 and 0.40. Fig. 3 shows a typical finite
element mesh of a unit cell for a three-phase polydisperse com- uRIGHT − uLEFT = u(2) , uTOP − uBOTTOM = u(4) ,
posite for a matrix with volume fraction c (1 ) = 0.60 and two dif- FRONT BACK (5 )
u −u =u , (49)
ferent inclusion types with c (2 ) = 0.25 and c (3 ) = 0.15.
where the superscripts (LEFT, RIGHT), (BOTTOM, TOP), and (BACK,
FRONT) denote collectively all material points located respectively
on the faces of the cell at (X1 = 0, X1 = L ), (X2 = 0, X2 = L ), and
4.1. The effective yield stress
(X3 = 0, X3 = L ). Eq. (49) show that the periodic constraints be-
tween all corresponding opposite boundary points can be written
We determine numerically the effective yield stress by solving
in terms of the displacements of the three vertex points (u(2) , u(4) ,
the problem of a unit cell loaded in uniaxial tension. Periodicity
u(5) ), which are defined, in turn, in (48) by the macroscopic de-
conditions are imposed on the boundary of the unit cell. A detailed
formation gradient F. In ABAQUS, for given F, we impose boundary
discussion of the periodic boundary conditions on a unit cell can
conditions on (u(2) , u(4) , u(5) ) according to (48), and the periodicity
be found in (Suquet et al., 1987) or (Michel et al., 1999). Here, the
constraints (49) are enforced through a “user MPC” subroutine (or
periodic boundary conditions on the unit cell are imposed as fol-
the “EQUATION” option).
lows (see Mbiakop et al. (2015b) and Papadioti for more details).
For the problem of uniaxial tension in direction 1, the deforma-
Referring to Fig. 4, if we fix vertex 1 in order to eliminate rigid
tion gradient is of the form
body translations, then, in view of the periodicity of the displace-
ment field, we can write the displacements u at vertices 2, 4, and F = λ e1 e1 + λt (e2 e2 + e3 e3 ), (50)

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

8 I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19

Fig. 2. Finite element discretization of cubic unit cells for two-phase composites containing a random distribution of 30 spherical particles for volume fractions of 10, 20,
30 and 40%. The finite element meshes have (200,869; 112,281; 165,371; 159,303) nodes and (83,270; 45,679; 67,790; 65,543) elements respectively. The corresponding total
numbers of degrees of freedom, including pressures, are (436,067; 245,485; 360,533; 346,823).

Fig. 3. Finite element discretization of a cubic unit cell for a three-phase composite Fig. 4. Periodic unit cell.
containing a random distribution of 30 polydisperse spherical particles with vol-
ume fractions c (2) = 0.25 (yellow) and c (3) = 0.15 (blue). The finite element mesh
has 303,953 nodes, 124,225 elements, and the total number of degrees of freedom,
including pressures, is 663,409. (For interpretation of the references to color in this <σ ij > in the finite element solution:3
figure legend, the reader is referred to the web version of this article.) &
1
<σi j > = σi j (x ) dV, (53)
V cell
V cell
where (λ, λt ) are the axial and transverse stretch ratios and ei the
base vectors along the coordinate axes shown in Fig. 4; the bound- where V cell is the total volume of the deformed finite element
ary conditions (48) become mesh.
The conditions u2(4 ) = u3(5 ) and <σ22 > = <σ33 > = <σ12 > =
u1(2) = (λ − 1 ) L, u2(4) = u3(5) = (λt − 1 ) L, (51) <σ13 > = <σ23 > = 0 are used to verify the correctness of the finite
element solution.
The nodal displacement u1(2 ) was increased gradually, the solu-
u2(2) = u3(2) = u1(4) = u3(4) = u1(5) = u2(5) = 0. (52) tion was developed incrementally, and the average stress <σ 11 >
was determined by (53) at the end of every increment. As u1(2 )
In ABAQUS, we prescribe u1(2 ) (i.e., λ) and set R2(4 ) = R3(5 ) = 0,
where Ri(N ) denotes the ith component of the force at node N. The
3
The alternative calculation <σ11 > = R1(2) /A cell appears to be less convenient as
quantities R1(2 ) and (u2(4 ) , u3(5 ) ), i.e., λt , are determined by the finite it requires evaluation of the current cross sectional area of the deformed cell A cell
element solution. The corresponding macroscopic stresses σ ij are which in general does not remain flat due to the complex microstructure of the
determined from the numerical calculation of the average stresses unit cell.

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19 9

Fig. 5. Variation of normalized effective flow stress σ˜ 0 /σ0(1) with contrast ratio r = σ0(2) /σ0(1) for different values of the volume fraction c(2) . The full triangles are the results
of the unit cell finite element calculations and the solid lines are the predictions (39) of the model based on the H-S− estimate (μ0 = μ(1) ). The maximum difference
between the numerical results and the analytical estimates is 3%.

increases, the calculated average stress <σ 11 > reaches a constant In order to check the isotropy of the unit cell, we carried out
value, which defines the effective flow stress of the composite σ˜ 0 . calculations for uniaxial tension in directions 2 and 3. In all cases,
Fig. 5 shows the variation of the calculated effective flow stress the results were identical to those shown in Figs. 5 and 6.
from the unit cell finite element calculations with the contrast
ratio r = σ0(2 ) /σ0(1 ) for various volume fractions, together with 4.2. The strain concentration tensors
the predictions (39) of the homogenization model, based on the
Hashin-Shtrikman lower bound H-S− (μ0 = μ(1 ) ).4 For that data The unit cell finite element calculations discussed above were
shown in Fig. 5, the maximum difference between the predictions used also to determine the strain concentration factors defined in
(39) and the results of the unit cell finite element calculations is (20) as follows. At the end of every increment in the finite element
3% (note that the vertical axis in Fig. 1 starts at the value of 1). It is solution the average value of the Eulerian logarithmic strain tensor
also interesting to mention that an increase of the flow stress σ0(2 ) ε(r) was determined in every phase of the composite, where the
in the inclusions beyond (approximately) two times the flow stress superscript (r) denotes “phase r”. The macroscopic axial logarith-
of the matrix (2 σ0(1 ) ) does not change the effective flow stress of mic strain was also determined as ε̄ = ln λ, where λ is the axial
the composite for all volume fractions considered here. The finite stretch ratio used in (51) to drive the finite element calculations.
element calculations confirm the fact that, for σ0(2 )  2 σ0(1 ) , the Interestingly, the components of <ε(r) > are found to be propor-
inclusions do not deform plastically in the deforming unit cell and tional to ε̄ in the context of the present study; in particular, it is
are in agreement with earlier numerical results of Suquet (1997) found that
for c (2 ) = 30% and by Ponte Castañeda et al. (2001) and Idiart et al.
(2006) for c (2 ) = 15%. As we will see in the following, this result is <εi(jr ) > = Ci j ε̄ , (54)
due to the fact that the particles behave as being rigid beyond fur-
which leads to the following estimate for the strain concentration
ther increase of σ0(2 ) .
(1 ) α (r) :
Fig. 6 shows the variation of σ˜ 0 /σ0 of a three-phase com- 
posite for different values of the volume fraction c(3) as deter- (r ) dε̄ (r ) 2
α = = Ci j Ci j . (55)
mined from the unit cell finite element calculations and the pre- dε̄ 3
dictions (39) of the homogenization model. The material data are
typical for a TRIP5 steel with a ferritic matrix (phase 1) contain- Fig. 7 shows the variation of the strain concentration factors α (r)
ing retained austenite (phase 2), which transforms gradually to in a two-phase composite with the contrast ratio r = σ0(2 ) /σ0(1 ) for
martensite (phase 3) as the TRIP steel deforms plastically (e.g., see various volume fractions as determined from the unit cell finite el-
Papatriantafillou et al. (2006)). ement calculations (Eq. (54)) and the homogenization theory (Eqs.
(46)–(47)).
An important observation in the context of this figure is that at
a contrast ratio of r ࣃ 2, a sharp transition is observed where the
particles start behaving as being rigid, i.e., the average strain in
the particle is almost zero. This is validated by both the model and
4
Of all possible choices for μ0 shown in Fig. 1, the Hashin–Shtrikman lower
bound H-S− (μ0 = μ(1) ) gives the best estimate by comparison to the predictions the numerical results. In terms of the homogenization procedure,
of the unit cell results. this implies that the case of infinite contrast, i.e., rigid particles,
5
TRIP is the acronym for TRansformation Induced Plasticity. and finite contrast is very similar beyond a value of r ࣃ 2. A weak

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

10 I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19

Fig. 6. Variation of effective normalized flow stress σ˜ 0 /σ0(1) of a three-phase composite with a matrix volume fraction c (1) = 0.60 for different values of the volume fraction
c(3) . The homogenization estimates are based on H-S− and the contrast ratios are σ0(2) /σ0(1) = 1.875 and σ0(3) /σ0(1) = 5.

dependence of this sharp transition upon the volume fraction c is Angle θ takes values in the range −30◦ ≤ θ ≤ 30◦ , where, to within
observed in these figures. a given hydrostatic stress, θ = −30◦ corresponds to uniaxial ten-
Similar plots for a three-phase composite are shown in Fig. 8. sion, θ = 0 to pure shear, and θ = 30◦ to uniaxial compression.
The predictions of the homogenization theory agree well with the It is stressed at this point that the composite materials consid-
results of the unit cell finite element calculations. ered in this work are plastically incompressible and thus the ap-
Fig. 8 shows, in turn, the strain concentration factors in a three- plied stress triaxiality affects only the elastic part which is of no
phase material. The comparison between the model and the finite interest here. Thus the only relevant invariant studied in this sec-
element simulations is qualitatively good, whereas the model tends tion, apart from the J2 invariant, is the third deviatoric invariant J3
to underestimate the straining of the middle phase, i.e., the one defined above. The study of the effect of J3 , in turn, allows for a
with yield stress σ0(2 ) /σ0(1 ) = 1.875. Again, in the case of the third complete analysis of general triaxial loading states.
phase, when σ0(3 ) /σ0(1 ) = 5, the particle behaves as rigid which is As a consequence of the applied periodic boundary conditions
consistent with the observations of the previous figure. and the symmetry of the problem, the macroscopic (average) de-
formation of the unit cell is entirely described by the displace-
ments of the “reference vertices” (2,4,5), as shown in Fig. 4, which
can be written in the form
4.3. On the possible dependence of the effective flow stress on J3
u(2) = U1 e1 , u(4) = U2 e2 , u(5) = U3 e3 . (58)
Ponte Castañeda and Suquet (1995); Suquet and Ponte Cas-
tañeda (1993) studied the effective mechanical behavior of weakly In ABAQUS, the displacements (U1 , U2 , U3 ) are tied, through “user
inhomogeneous composites and showed that, for the case of in- multipoint constraints”, to the degrees of freedom of a fictitious
compressible “power-law” phases, the effective potential of the node, which is properly displaced so that the desired triaxiality
composite may depend, to second order, on the third invariant of X and Lode angle θ are achieved. Details of the numerical for-
the applied strain. mulation can be found in (Mbiakop et al., 2015a; 2015b) (see also
We carry out detailed unit cell finite element calculations in or- Barsoum and Faleskog (2007) and Papadioti).
der to check for a possible dependence of the effective yield stress We carry out finite element calculations in which the unit cell
σ˜ 0 on the third invariant J3 of the stress deviator s (J3 = dets, is loaded with X = 1/3 and Lode angles in the range −30◦ ≤ θ ≤
where ‘ det’ denotes the determinant). We identify the coordinate 30◦ . The finite element analysis is carried out incrementally; at
axes shown in Fig. 4 with the principal directions of the stress ten- the end of each increment the average stress  < σ > and the cor-
sor and write the principal stresses in the form responding von Mises equivalent stress σ̄e = 3
2 < s >:< s > are
⎛ ⎧ π  ⎫⎞ calculated. As the applied displacement of the fictitious node in-
' ( ' ( ⎪ cos θ + ⎪ creases, σ̄e takes a constant value, which defines the effective flow
σ1 1 ⎨ 6 ⎬⎟
⎜ 2
stress σ˜ 0 of the periodic composite.
σ2 = σe ⎝X 1 + θ
sin  ⎠, (56)
3⎪⎩− cos θ − π ⎪ ⎭
σ3 1 In order to verify that the desired values have been indeed
6 achieved, the triaxiality and Lode angle corresponding to the av-
erage stress < σ > are determined at the end of every increment.
where X = p/σe is the stress triaxiality and θ is the “Lode angle”, Also, since the coordinate axes in the finite element solution are
so that assumed to coincide with the principal stress directions, the condi-
tions < σ12 >=< σ13 >=< σ23 >= 0 are checked at the end of ev-
2
J3 = dets = − σe3 sin 3θ . (57) ery increment to verify the accuracy of the finite element solution.
27
Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19 11

Fig. 7. Strain concentration factors α (i ) = dε̄ (i ) /dε̄ as determined from unit cell finite element calculations and homogenization theory (Eqs. (46) and (47)) for a two-phase
composite.

Fig. 9 shows the variation of the effective flow stress σ˜ 0 , For simplicity, we consider infinitesimal displacement gradi-
as determined from unit cell finite element calculations, with ents (small strains and rotations); the method is easily extended
Lode angle θ for particle volume fractions c (2 ) =0.10, 0.20, and to cover the case of finite geometry changes as discussed in
0.40. Fig. 9 shows that the effective flow stress of the com- Appendix. The elastic and plastic response of the homogenized
posite is essentially independent of the third stress invariant J3 , composite are treated independently, and combined later to obtain
which is in agreement with earlier results by Suquet (1997), Ponte the full elastic-plastic response. The infinitesimal strain tensor ε at
Castañeda et al. (2001) and Idiart (2008) in the case of rigid every point in the homogenized material is written as
particles. ε = εe + ε p , (59)
where εeand εp
are the elastic and plastic parts. Linear isotropic
5. Hardening phases elastic behavior is assumed:
εe = Me : σ or σ = Le : εe , (60)
In this section we present an approximate method for the pre- e
where M is the isotropic elastic compliance tensor, which is the
diction of the incremental elastoplastic behavior of macroscopically
inverse of the isotropic elasticity tensor Le :
isotropic composites made of N isotropic, rate-independent, elastic-
 −1 1 1
plastic hardening phases. Let the flow stresses σy(i ) of each phase Le = 2 μ K + 3 κ J , M e = Le = K+ J, (61)
be known functions of the corresponding equivalent plastic strains 2μ 3κ
ε̄ (i) (i = 1, 2, . . . , N ). At every point of the homogenized composite and μ and κ denote the effective elastic shear and bulk moduli of
the “internal variables” that characterize the local state of the ho- the composite.
mogenized continuum are the local values ofthe equivalent plastic Let t be a loading (time-like) parameter and consider an in-
strains in the phases q = ε̄ (1 ) , ε̄ (2 ) , . . . , ε̄ (N ) . finitesimal change from tn to tn+1 = tn + t, where t is “small”.

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

12 I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19

Fig. 8. Strain concentration factors α (i ) = dε̄ (i ) /dε̄ as determined from unit cell finite element calculations and homogenization theory (Eq. (40)) for a three-phase composite.

Fig. 9. Variation of effective normalized flow stress σ˜ 0 /σ0(1) with Lode angle θ for particle volume fractions of 10, 20, and 40%. The results show almost no dependence on
J3 .

We use the notation An and An+1 to denote the values of A at the updated at every increment. The value of σ˜ 0 (qn+1 ) is calculated by
start tn and the end tn+1 of the increment and set A = An+1 − An . the solution of the corresponding optimization problem (39) using
We assume that the effective flow stress is, to a first approxima- the σ0(i ) values defined in (62). The solution of the optimization
tion, constant over the period (tn , tn+1 ) and can be determined by problem (39) defines also the optimal values yˆ(r ) (qn+1 ), which de-
the optimization problem in (39), in which the flow stresses of the termine the corresponding strain concentration factors α (i) in (40)
phases take values for the increment. The actual calculation is implicit in general, ex-
/ /
σ0(i) = (1 − β )σ0(i) /n + β σ0(i) /n+1 , 0 ≤ β ≤ 1. (62) cept when β = 0 is used in (62).
Over any time increment (tn , tn+1 ) the effective yield condition
where
/   of the composite is written in the form
σ0(i) /n = σy(i) ε̄n(i) and
/  )    (σ , qn+1 ) = σe − σ˜ 0 (qn+1 ) = 0, (64)
σ0(i) /n+1 = σy(i) ε̄n(i+1 = σy(i ) ε̄n(i ) + ε̄ (i ) . (63)
Put in other words, the composite is assumed to behave as “incre- where σ˜ 0 (qn+1 ) is determined from the solution of the optimiza-
mentally perfectly plastic” with a flow stress σ˜ 0 (qn+1 ), which is tion problem (39) with σ0(i ) defined in (62), and the associated

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19 13

flow rule is Remark 1. In the special case where the value β = 0 is used in
3 (62), the effective flow stress of the composite σ˜ 0 and the strain
ε˙ p = ε̄˙ N, N = s, (65) concentration factors α (i) are determined using the values of the
2 σe
 flow stresses of the phases σ0(i ) |n at the start of the increment, and
where ε̄˙ = 2
3 ε˙ p : ε˙ p is the macroscopic effective equivalent plas- Eqs. (70) and (71) can be solved analytically:
tic strain rate. The evolution of the equivalent plastic strains in the σee − σ˜ 0 (qn )
ε̄ = and qi = ε̄ α (i ) (qn ). (72)
phases are written in terms of the strain concentration factors α (i) 3μ
defined in (40) in terms of the optimal values yˆ(r ) (qn+1 ), i.e.,
The integration scheme described above is implemented into
q˙ i = ε̄˙ α (i ) (qn+1 ), i = 1, 2, . . . , N. (66) the ABAQUS general purpose finite element program (Hibbitt,
1977). This code provides a general interface so that a particular
5.1. Numerical integration of constitutive equations constitutive model can be introduced as a user subroutine (UMAT).
The finite element formulation is based on the weak form of
In the following, we present a method for the numerical inte- the momentum balance, the solution is carried out incrementally,
gration of the resulting constitutive equations in the context of a and the discretized nonlinear equations are solved using Newton’s
displacement driven finite element formulation. In a finite element method. In the calculations, the Jacobian of the global Newton
environment, the solution is developed incrementally and the con- scheme is approximated by the tangent stiffness matrix. Such an
stitutive equations are integrated at the element Gauss points. At approximation of the Jacobian is first-order accurate as the size of
a Gauss point in the finite element mesh, the solution (εn , σ n , qn ) the increment t → 0; it should be emphasized, however, that the
at time tn as well as the infinitesimal strain εn+1 at time tn+1 are aforementioned approximation influences only the rate of conver-
known, and the problem is to determine (σ n+1 , qn+1 ). gence of the Newton loop and not the accuracy of the results.
A backward Euler integration scheme is used for the numerical
integration of the flow rule (65): 5.2. Unit cell calculations and assessment of the model with
ε p = ε̄ Nn+1 . (67) hardening phases

The elasticity Eq. (60) is written in the form In this section we present the results of unit cell finite el-
σ n+1 = σ n + L : ( ε − ε ) = σ − 2 μ ε̄ Nn+1 ,
e p e
(68) ement calculations for a composite material made up of a sta-
tistically isotropic random distribution of isotropic, linearly-elastic
where σ e = σ n + Le : ε is the (known) “elastic predictor”. Con- hardening-plastic spherical inclusions embedded in a continu-
sidering the deviatoric part of last equation and using the defini- ous, isotropic, linearly-elastic hardening-plastic matrix. All analyses
tion (65) of Nn+1 we conclude that the stress deviator sn+1 is co- were carried out incrementally and accounted for geometry
linear with the deviatoric part of the elastic predictor se . Therefore, changes due to deformation (finite strain solutions).
we can determine the direction Nn+1 of the plastic strain rate at In all cases analyzed, the matrix material is identified as “phase
tn+1 by using the known elastic predictor as 1” and the flow stress σy(i ) of “phase i” is a function of the corre-
3 3 e sponding equivalent plastic strain ε̄ p :
Nn+1 = sn+1 = s = known, (69)
2 σe 2 σee   1
   ε̄ (i) η(i) σ (i )
σy(i) ε̄ (i) = σ0(i) 1 + (i) , ε0(i ) = 0 , (73)
where σee = 3
2 se : se is the von Mises equivalent stress of the ε0 E
elastic predictor. Projecting (68) in the direction of the plastic
strain rate Nn+1 and taking into account that σ : N = σe and N : where σ0(i ) = σy(i ) (0 ) is the yield stress of phase i, E is the elas-
N = 32 , we find σe |n+1 = σee − 3 μ ε̄ . Therefore, the yield condi- tic Young’s modulus, and the hardening exponents η(i) take values
tion (64) can be written at the end of the increment in the form in the region 1 ≤ η(i) ≤ ∞, with the limiting case η (i ) = ∞ corre-
sponding to perfect plasticity. Note that this hardening exponents
σee − 3 μ ε̄ − σ˜ 0 (qn + q ) = 0. (70) are completely uncorrelated to the creep exponent n(i) used in the
definitions of the stress potentials in the previous sections.
The evolution of the equivalent plastic strains in the phases (66) The values E = 917 σ0(1 ) and ν = 0.3 for Young’s modulus E and
are written also as Poisson ratio ν are used in the calculations. In addition, one-
qi = ε̄ α (i) (qn + q ), i = 1, 2, . . . , N. (71) element finite element calculations were carried out, in which the
element is subjected to the same deformation gradient as the unit
The problem of integrating the elastoplastic equations for the cell and the corresponding uniform stress state in the element is
homogenized composite reduces to the solution of the set calculated by using the algorithm described in Section 5.1 for the
of N + 1 non-linear Eqs. (70) and (71) for ε̄ and q = homogenized material.
( ε̄ (1) , ε̄ (2) , . . . , ε̄ (N ) ). These equations are solved by using
Newton’s method. In every Newton iteration the values of q are
5.2.1. Two-phase composites
used to calculate the corresponding σ0(i ) from (62) and then the
We analyze first a two-phase composite with
optimization problem (39) is solved by using the CONMAX soft-
ware (Kaufman et al., 1995) to determine the optimal values yˆ(i ) ; σ0(2)
the values of the effective flow stress σ˜ 0 and the strain concentra- = 1.5, η ( 1 ) = 5, η ( 2 ) = 3. (74)
σ0(1)
tion factors α (i) are then determined and the iterations are contin-
ued until the set tolerances are met. Details on the calculation of The corresponding stress-strain curves of the phases in uniaxial
the Jacobian of the Newton loop are given in Papadioti. tension are shown in Fig. 10.
It is emphasized that the calculations are much simpler for a Fig. 11 shows the results of the unit cell finite element calcula-
two-phase composite; in that case, one does not need to invoke tions together with the predictions of the homogenization model
CONMAX, since σ˜ 0 is defined analytically by (45). for the case of uniaxial tension in direction 1 and for inclusion
Once ε̄ and q are calculated, Eqs. (68) and (71) are used to volume fractions c2 =0.10, 0.20, 0.30, and 0.40. The quantity σ˜ in
determine the stress σ n+1 and the state variables qn+1 . Fig. 11 is the average stress < σ 11 > in the unit cell calculations

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

14 I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19

Fig. 10. Uniaxial stress–strain curves of phases.

Fig. 11. Stress–strain curves of the two-phase composite in uniaxial tension for different values of the volume fraction c(2) . The solid lines are the results of the unit cell
finite element calculations and the dash lines are the predictions (39) of the model based on the H-S− estimate (μ0 = μ(1) ).

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19 15

Fig. 12. Deformed configurations of unit cells in simple shear (γ = 0.15 ) for various values of the volume fraction c(2) .

Fig. 13. Shear stress-shear strain curves of the two-phase composite in simple shear for different values of the volume fraction c(2) . The solid lines are the results of the unit
cell finite element calculations and the dash lines are the predictions (39) of the model based on the H-S− estimate (μ0 = μ(1) ).

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

16 I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19

and the uniform σ 11 stress component in the corresponding one- Fig. 13 shows the results of the unit cell finite element calcula-
element homogenization calculation. The predictions of the ho- tions together with the predictions of the homogenization model
mogenization model agree well with the numerical results. It is for the case of finite shear on the 1–2 plane and for inclusion
also evident from these figures that as we increase the volume volume fractions c2 =0.10, 0.20, 0.30, and 0.40. The quantity τ˜ in
fraction of the stiffer particle phase which also has a higher hard- Fig. 13 is
ening exponent, this leads to a reinforcement of the composite 
both at the level of the yield strength as well as in its hard- 1
τ˜ = si j si j , (76)
ening response. It is also interesting to note that even though 2
we have added the hardening behavior of the phases heuristi-
where sij is identified with the average deviatoric stresses < sij > in
cally to the homogenization model for perfectly plastic phases (see
the unit cell calculations and with the uniform deviatoric stresses
Eq. (45)), the corresponding analytical estimates are in excellent
sij in the one-element homogenization calculations. Again, the pre-
agreement with those obtained by the finite element calculations
dictions of the homogenization model agree well with the numer-
(see also relevant discussion in (Suquet, 1997)). This, in turn, sug-
ical results. Similar observations to those made in the context of
gests that such a simplified approach is sufficient for this class of
Fig. 11 could also be made in Fig. 13 regarding the effect of vol-
materials.
ume fraction and the hardening exponent of the phases upon the
Calculations are also carried out for finite shear deformation. In
effective response under shear loadings.
this case, the deformation gradient used in (48) is of the form

F = δ + γ e1 e2 , (75) 5.2.2. Three-phase composites


where γ is the amount of shearing on the 1–2 plane. Fig. 12 shows We consider next a three-phase composite with
the deformed unit cell at γ = 0.15 for various inclusion volume σ0(2) σ0(3)
fractions c2 . (1 )
= 1.875, = 5, η ( 1 ) = 5, η ( 2 ) = 3, η ( 3 ) = 2.5. (77)
σ0 σ0(1)

Fig. 14. Deformed configurations of unit cells of the three-phase composite in uniaxial tension (λ = 1.20 ) and simple shear (γ = 0.20 ).

Fig. 15. Stress–strain curves of the three-phase composite in uniaxial tension and simple shear. The solid lines are the results of the unit cell finite element calculations and
the dash lines are the predictions (39) of the model based on the H-S− estimate (μ0 = μ(1) ).

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19 17

The problems of uniaxial tension and finite shear deformation are on the third invariant (J3 ) of the stress tensor, i.e., on the Lode pa-
solved. rameter or Lode angle. Nonetheless, this dependence is extremely
Fig. 14 shows the deformed unit cells for uniaxial tension at weak and thus the present model, which does not take into ac-
λ = 1.20 and finite shear γ = 0.20. count this dependence, is sufficiently accurate for the estimation
Fig. 15 shows the stress-strain curves in uniaxial tension and of the effective response as well as of the phase average strains
finite shear, for a three-phase composite with composition c (1 ) = (which depend apriori upon the normal to the yield surface). This
0.60, c (2 ) = 0.25, and c (3 ) = 0.15. The predictions of the homoge- observation of course is valid for phases described by a von Mises
nization model agree well with the results of the unit cell finite el- (J2 ) yield response and does not hold in the case of plastic solids
ement calculations. The model is capable of predicting sufficiently that depend directly on the third invariant J3 via Tresca, Hosh-
well both the initial yield strength of the three-phase composite as ford or Drucker-Prager plasticity (see for instance Barthélémy and
well the hardening evolution as a function of the applied strains Dormieux (2004) and Maghous et al. (2009)).
both in uniaxial tension and shear loadings. A third, and perhaps more important finding of this work at
least from a more practical point of view, is related to the exten-
6. Concluding remarks sion of the model to arbitrary isotropic hardening of the phases.
In the present work, we carry out first the nonlinear homogeniza-
The present work presents a simple semi-analytical model for tion for perfectly plastic phases and then the hardening is added
the estimation of the effective as well as the phase average re- heuristically at the definition of the yield stresses of each of the
sponse of N−phase incompressible isotropic elasto-plastic metallic phases. This of course is an approximation and does not take into
materials. The model is based on the original variational method account explicitly the coupling between the different hardening
of Ponte Castañeda (1991), which is based on a linear compari- exponents of the different phases. Nevertheless, the resulting esti-
son composite technique. The resulting expression for the effec- mates are in very good agreement with the full field finite element
tive yield strength of the composite requires the solution of a results (which include this coupling) suggesting that this strategy
constrained optimization problem for N − 1 scalar variables and is is sufficient for the materials and loadings considered in this study.
much simpler and tractable than the original expressions given in This good agreement can be directly associated with the fact that
(Ponte Castañeda, 1992). This is achieved by use of the method- the model predicts accurately enough the phase average strains.
ology of Kaufman et al. (1995). In the special case of a two-phase This, however, may not be true if one considers kinematic hard-
composite, we provide a fully explicit expression which is given via ening of the phases or more complex non-proportional loadings
a piecewise function defined in Eq. (45). Due to its simplicity and but again in this case a more advanced homogenization method
numerical efficiency, the proposed N−phase model is numerically needs to be used such as the one proposed by Lahellec and Suquet
implemented in a user-material subroutine which, in turn, allows (2007).
for the simulation of three dimensional geometries. Finally, in this study, we do not consider the extreme case of
In addition, the N−phase analytical model is compared with a three-phase composite comprising stiff particles and voids. The
full field three dimensional finite element simulations of two- and reason is that the presence of a soft compressible phase would
three-phase multi-particle periodic unit cells. The proposed model introduce a dependence on an additional invariant, i.e, pressure
is found to be in good agreement with the finite element results (I1 ) and the material would be plastically compressible (see for
in most of the cases considered here, even at large particle vol- instance Gărăjeu and Suquet (1997) and He et al. (2013)). A vast
ume fractions and different hardening exponents. The agreement is amount of studies has been carried out in the context of voided
good both for the effective average stress strain response, as well materials and is very well known that the present method by de-
as for the phase average strain fields. This last observation allowed fault would lead to overly stiff estimates unless corrected (see for
to extend the model in a heuristic manner to account for arbitrary instance recent work of Danas and Aravas (2012) and Cao et al.
isotropic hardening of the phases, both in a small and finite strain (2015)). Such work is now underway and will be reported else-
formalism. The present combined analytical and numerical study where.
reveals several nontrivial features which are summarized in the
following. Acknowledgments
One of the main non-intuitive observations in the present work,
which is in accord with former literature, is that in the context of This work was carried out while I.P. and N.A. were supported by
a two-phase composite when the strength of the particles is al- the Greek General Secretariat for Research and Technology (GSRT)
most twice that of the matrix the particles behave as being rigid through the KPHI Project. K.D. greatfully acknowledge partial
for all volume fractions considered here. In other words, we ob- support by the French National Research Agency (ANR) via the
tain a rather sharp transition when the yield stress of the parti- grant INDiANA (ANR-12-RMNP-0011). The authors would like also
cle is about two times that of the matrix beyond which the strain to acknowledge the help of late Prof. A. Stamatis of University
in the particle is negligible, thus leading to an almost rigid re- of Thessaly on the solution of the optimization problem. Use of
sponse of the particle in the sense of straining for a large range the FORTRAN code CONMAX (http://www.netlib.org/opt/conmax.f)
of volume fractions. This result was shown for the effective yield is gratefully acknowledged.
stress by Suquet (1997) and Idiart (2008) for given particle volume
fraction, whereby it is further shown here that this sharp transi- Appendix. Finite strain formulation
tion is weakly sensitive to the particle volume fraction (at least
for volume fractions up to 40%). This, in turn, may have signifi- The constitutive equations become
cant implications in the strengthening of the composite and pos-
sible debonding/failure of the particle/matrix interface (Bignonnet D = De + D p , (78)
et al., 2015), since beyond that contrast ratio the particle stops de-
forming. This of course leads to stress and strain concentrations in D e = Me : σ ∇ , (79)
the matrix/article interface.
3
A second observation, which has already been made in the con- D p = ε̄˙ N, N= s, (80)
2 σe
text of a plastic matrix with rigid particles by Suquet (1997) and / /
Idiart (2008) is that the numerical estimates exhibit a dependence σ0(i) = (1 − β ) σ0(i) /n + β σ0(i) /n+1 , (81)

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

18 I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19

References
(σ , q ) = σe − σ˜ 0 (q ) = 0, (82)
Barsoum, I., Faleskog, J., 2007. Rupture mechanisms in combined tension and shear—
q˙ i = ε̄˙ α (i ) (q ), i = 1, 2, . . . , N, (83) micromechanics. Int. J. Solids Struct. 44, 5481–5498.
Barthélémy, J.F., Dormieux, L., 2004. A micromechanical approach to the strength
where ∇ denotes the Jaumann or co-rotational derivative. criterion of Drucker–Prager materials reinforced by rigid inclusions. Int. J. Nu-
mer. Anal. Methods Geomech. 28, 565–582.
In a finite element environment, the solution is developed in- Bignonnet, F., Dormieux, L., Lemarchand, E., 2015. Strength of a matrix with elliptic
crementally and the constitutive equations are integrated at the el- criterion reinforced by rigid inclusions with imperfect interfaces. Eur. J. Mech. A
ement Gauss integration points. Let F denote the deformation gra- Solids 52, 95–106.
Bonnenfant, D., Mazerolle, F., Suquet, P., 1998. Compaction of powders containing
dient tensor. At a given Gauss point, the solution (Fn , σ n , qn ) at hard inclusions: experiments and micromechanical modeling. Mech. Mater. 29,
time tn as well as the deformation gradient Fn+1 at time tn+1 are 93–109.
known, and the problem is to determine (σ n+1 , qn+1 ). Brassart, L., Stainier, L., Doghri, I., Delannay, L., 2011. A variational formulation for
the incremental homogenization of elasto-plastic composites. J. Mech. Phys.
The time variation of the deformation gradient F during the Solids 59, 2455–2475.
time increment [tn , tn+1 ] can be written as Cao, T.-S., Maziere, M., Danas, K., Besson, J., 2015. A model for ductile damage pre-
diction at low stress triaxialities incorporating void shape change and void ro-
F(t ) = F(t ) · Fn = R(t ) · U(t ) · Fn , tn ≤ t ≤ tn+1 , (84) tation. Int. J. Solids Struct. 63, 240–263.
Danas, K., Aravas, N., 2012. Numerical modeling of elasto-plastic porous materials
where R(t) and U(t) are the rotation and right stretch tensors as- with void shape effects at finite deformations. Compos. Part B 43, 2544–2559.
deBotton, G., Ponte Castañeda, P., 1993. Elastoplastic constitutive relations for fiber-
sociated with F(t). The corresponding deformation rate D(t) and reinforced solids. Int. J. Solids Struct. 30, 1865–1890.
spin W(t) tensors are given by Eshelby, J.D., 1957. The determination of the elastic field of an ellipsoidal inclusion,
    and related problems. Proc. R. Soc. Lond. A 241, 376–396.
D(t ) ≡ F˙ (t ) · F−1 (t ) = F˙ (t ) · F−1 (t ) , (85) Fritzen, F., Forest, S., Böhlke, T., Kondo, D., Kanit, T., 2012. Computational homoge-
s s
nization of elasto-plastic porous metals. Int. J. Plast. 29, 102–119.
and Gărăjeu, M., Suquet, P., 1997. Effective properties of porous ideally plastic or vis-
    coplastic materials containing rigid particles. J. Mech. Phys. Solids 45, 873–902.
W(t ) ≡ F˙ (t ) · F−1 (t ) = F˙ (t ) · F−1 (t ) , (86) He, Z., Dormieux, L., Kondo, D., 2013. Strength properties of a Drucker–Prager
porous medium reinforced by rigid particles. Int. J. Plast. 51, 218–240.
a a
Hibbitt, H.D., 1977. ABAQUS/EPGEN—a general purpose finite element code with em-
where the subscripts s and a denote the symmetric and anti- phasis in nonlinear applications. Nucl. Eng. Des. 77, 271–297.
symmetric parts, respectively. Hill, R., 1965. A self-consistent mechanics of composite materials. J. Mech. Phys.
If it is assumed that the Lagrangian triad associated with F(t) Solids 13, 213–222.
Idiart, M., 2008. Modeling the macroscopic behavior of two-phase nonlinear com-
(i.e., the eigenvectors of U(t)) remains fixed over the time interval posites by infinite rank laminates. J. Mech. Phys. Solids 56, 2599–2617.
(tn , tn+1 ), it can be shown readily that Idiart, M.I., Moulinec, H., Ponte Castañeda, P., Suquet, P., 2006. Macroscopic behavior
and field fluctuations in viscoplastic composites: Second-order estimates versus
D(t ) = R(t ) · E˙ (t ) · RT (t ), W(t ) = R˙ (t ) · RT (t ), full-field-simulations. J. Mech. Phys. Solids 54, 1029–1063.
Kailasam, M., Ponte Castañeda, P., 1998. A general constitutive theory for linear and
σ ∇ (t ) = R(t ) · σˆ˙ (t ) · RT (t ), (87) nonlinear particulate media with microstructure evolution. J. Mech. Phys. Solids
46, 427–465.
where a superscript T indicates the transpose of a second-order Kaufman, E.H., Leeming, D.J., Taylor, G.D., 1995. An ODE-based approach to nonlin-
tensor, E(t ) = ln U(t ) is the logarithmic strain relative to the con- early constrained minimax problems. Numer. Algorithms 9, 25–37. (CONMAX
software available at www.netlib.org/cgi-bin/search.pl)
figuration at the start of the increment, and σˆ (t ) = RT (t ) · σ (t ) · Lahellec, N., Suquet, P., 2007. On the effective behavior of nonlinear inelastic com-
R(t ). posites: I. incremental variational principles. J. Mech. Phys. Solids 55, 1932–
It is noted that at the start of the increment (t = tn ) 1963.
Lopez-Pamies, O., Goudarzi, T., Danas, K., 2013. The nonlinear elastic response of
F n = R n = U n = δ , σˆ n = σ n , and En = 0, (88)
suspensions of rigid inclusions in rubber. II. A simple explicit approximation for
finite-concentration suspensions. J. Mech. Phys. Solids 61, 19–37.
Maghous, S., Dormieux, L., Barthélémy, J.F., 2009. Micromechanical approach to the
whereas at the end of the increment (t = tn+1 ) strength properties of frictional geomaterials. Eur. J. Mech. A Solids 28, 179–188.
Mbiakop, A., Constantinescu, A., Danas, K., 2015a. An analytical model for porous
Fn+1 = Fn+1 · F−1
n = Rn+1 · Un+1 = known, and single crystals with ellipsoidal voids. J. Mech. Phys. Solids 84, 436–467.
Mbiakop, A., Constantinescu, A., Danas, K., 2015b. On void shape effects of periodic
En+1 = ln Un+1 = known. (89) elasto-plastic materials subjected to cyclic loading. Eur. J. Mech. A Solids 49,
481–499.
Then, the constitutive equations of the model can be written as Michel, J.C., Moulinec, H., Suquet, P., 1999. Effective properties of composite mate-
rials with periodic microstructure: a computational approach. Comp. Methods
E˙ = E˙ e + E˙ p or E = Ee + E p , (90) Appl. Mech. Eng. 172, 109–143.
Michel, J.C., Suquet, P., 1992. The constitutive law of nonlinear viscous and porous
  materials. J. Mech. Phys. Solids 40, 783–812.
E˙ e = Me : σˆ˙ or Een+1 = Me : σˆ n+1 − σ n , (91) Papadioti, I., Non-linear Homogenization Theories with Applications to TRIP Steels.
Ph.d. thesis. University of Thessaly. Greece. In preparation, 2016.
3 Papatriantafillou, I., Agoras, M., Aravas, N., Haidemenopoulos, G., 2006. Constitu-
E˙ p = ε̄˙ N
ˆ, ˆ =
N sˆ , (92) tive modeling and finite element methods for TRIP steels. Comp. Methods Appl.
2 σe Mech. Eng. 195, 5094–5114.
/ /
σ0(i) = (1 − β ) σ0(i) /n + β σ0(i) /n+1 ,
Ponte Castañeda, P., 1991. The effective mechanical properties of nonlinear isotropic
(93) composites. J. Mech. Phys. Solids 39, 45–71.
Ponte Castañeda, P., 1992. New variational principles in plasticity and their applica-
tion to composite materials. J. Mech. Phys. Solids 40, 1757–1788.
(σˆ , q ) = σe − σ˜ 0 (q ) = 0, (94) Ponte Castañeda, P., 2002. Second-order homogenization estimates for nonlinear
composites incorporating field fluctuations: I–theory. J. Mech. Phys. Solids 50,
q˙ i = ε̄˙ α (i ) (q ), i = 1, 2, . . . , N, (95) 737–757.
Ponte Castañeda, P., 2005. Heterogeneous Materials, Lecture Notes, École Polytech-
nique, Department of Mechanics, ISBN 2-7302-1267-1.
where
 we took into account that Een = 0, σˆ n = σ n , and σe = Ponte Castañeda, P., Suquet, P., 1995. On the effective mechanical behavior of weakly
3 3 inhomogeneous nonlinear materials. Eur. J. Mech. A Solids 14, 205–236.
2 s:s=
ˆ ˆ
2 s : s. The constitutive equations listed above are Ponte Castañeda, P., 1997. Nonlinear composite materials: effective constitutive be-
identical to those of the infinitesimal strain formulation and can havior and microstructure evolution. In: Suquet, P. (Ed.), Continuum Microme-
chanics, CISM Courses and Lectures No. 377. Springer-Verlag, pp. 131–195.
be integrated as described in Section 5. The integration determines
Ponte Castañeda, P., Suquet, P., 1998. Nonlinear composites. Adv. Appl. Mech. 34,
σˆ n+1 and the true stresses are calculated as σ n+1 = Rn+1 · σˆ n+1 · 171–302.
RTn+1 .

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022
JID: SAS
ARTICLE IN PRESS [m5G;March 8, 2016;19:11]

I. Papadioti et al. / International Journal of Solids and Structures 000 (2016) 1–19 19

Ponte Castañeda, P., Suquet, P., 2001. Nonlinear Composites and Microstructure evo- Suquet, P., Ponte Castañeda, P., 1993. Small-contrast perturbation expansions for the
lution. In: Aref, H., Phillips, J. (Eds.), Mechanics for a new Millenium. Kluwer effective properties of nonlinear composites. C. R. Acad. Sci. Paris (Série II) 317,
Academic Publishers, pp. 253–273. 1515–1522.
Ponte Castañeda, P., Zaidman, M., 1994. Constitutive models for porous materials Suquet, P., 1987. Elements of homogenization for inelastic solids. In: Sanchez-
with evolving microstructure. J. Mech. Phys. Solids 42, 1459–1497. Palencias, E., Zaoui, A. (Eds.), Homogenization Techniques for Composite Media.
Rintoul, M.D., Torquato, S., 1997. Reconstruction of the structure of despersions. J. Springer-Verlag, pp. 194–278.
Collois Inter. Sci. 186, 467–476. Suquet, P., 1997. Effective properties of nonlinear composites. In: Suquet, P. (Ed.),
Schöberl, J., 1997. NETGEN an advancing front 2d/3d-mesh generator based on ab- Continuum Micromechanics, CISM Courses and Lectures No. 377. Springer-
stract rules. Comput. Vis. Sci. 1, 41–52. Verlag, pp. 197–264.
Segurado, J., Llorca, J., 2002. A numerical approximation to the elastic properties of Willis, J.R., 1977. Bounds and self-consistent estimates for the overall properties of
sphere-reinforced composites. J. Mech. Phys. Solids 50, 2107–2121. anisotropic composites. J. Mech. Phys. Solids 25, 185–202.
Suquet, P., 1995. Overall properties of nonlinear composites: a modified secant mod- Willis, J.R., 1982. Elasticity theory of composites. In: Hopkins, H.G., Sewell, M.J.
uli theory a nd its link with Ponte Castañedas nonlinear variational procedure’. (Eds.), Mechanics of Solids, The Rodney Hill 60th Anniversary Volume. Perga-
C. R. Acad. Sci. Paris (Série II) 320, 563–571. mon Press, pp. 653–686.

Please cite this article as: I. Papadioti et al., A methodology for the estimation of the effective yield function of isotropic composites,
International Journal of Solids and Structures (2016), http://dx.doi.org/10.1016/j.ijsolstr.2016.02.022

You might also like