Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/283875313

Anisotropic damage modeling of concrete materials

Article in International Journal of Damage Mechanics · September 2015


DOI: 10.1177/1056789515606509

CITATIONS READS

48 630

1 author:

Rodrigue Desmorat
École normale supérieure Paris-Saclay
186 PUBLICATIONS 4,204 CITATIONS

SEE PROFILE

All content following this page was uploaded by Rodrigue Desmorat on 03 August 2020.

The user has requested enhancement of the downloaded file.


Article
International Journal of Damage
Mechanics
2016, Vol. 25(6) 818–852
Anisotropic damage modeling ! The Author(s) 2015
Reprints and permissions:
of concrete materials sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/1056789515606509
ijd.sagepub.com

Rodrigue Desmorat

Abstract
An anisotropic damage model is proposed for concrete materials. As required by thermodynamics a single
damage variable, tensorial, is considered for any loading: as a state variable it represents the micro-
cracking pattern whatever the loading sign. Damage anisotropy is used to model the strong dissymmetry
tension/compression.The Ladevèze damage variable H ¼ (1  D)1/2 is introduced within a deviatoric/
hydrostatic split. An original shear-bulk coupling is derived, in accordance with numerical discrete element
computations. The sought property of gradual stress softening, with a tail in stress–strain diagram, is
obtained. Stress triaxiality is used to enhance the performance of Mazars criterion and therefore of the
full anisotropic damage model in bicompression.

Keywords
Anisotropic damage, concrete, shear-bulk coupling, triaxiality, dissipation, nonlocal

Introduction
Damage anisotropy is loading induced. From a micro-mechanics point of view it is due to an
oriented micro-cracking pattern. From a continuum damage mechanics point of view anisotropic
damage is represented by a tensorial damage variable, either a fourth-order tensor D of components
Dijkl (Chaboche, 1982; Chaboche, 1984; Chaboche and Maire, 2000; Ju, 1989; Leckie and Onat,
1981; Lemaitre and Chaboche, 1985; Maire and Chaboche, 1997) or a symmetric second-order
damage tensor D of components Dij (Cordebois and Sidoroff, 1982; Ladevèze, 1983; Murakami,
1988) such as Dijkl ¼ Dij ¼ 0 for a virgin material and such as rupture at a vanishing stress corres-
ponds to maximum principal damage equal to one.
Second-order anisotropic damage representation is restrictive compared to fourth-order tensorial
formulation, but since its interpretation is quite simple it has been widely and successfully used for
either metallic or quasi-brittle materials (Badel et al., 2007; Billardon and Pétry, 2005; Brünig, 2003;
Carol et al., 2001; Desmorat and Otin, 2008; Desmorat et al., 2007; Gatuingt, 2008; Halm and

Laboratoire de mécanique et Technologie, Université Paris Saclay, France


Corresponding author:
Rodrigue Desmorat, LMT-Cachan (ENS Cachan, CNRS, Université Paris-Saclay), 94235 Cachan, France.
Email: desmorat@lmt.ens-cachan.fr

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 819
Dragon, 1998; Lemaitre et al., 2000; Kattan and Voyiadjis, 1990; Menzel and Steinmann, 2001;
Menzel et al., 2002; Papa and Taliercio, 1996; Ramtani et al., 1992; Ragueneau et al., 2008;
Steinmann and Carol, 1998). The three principal values Di correspond to three orthogonal families
of microcracks. The damage evolution law then takes the tensorial rate form D _ ¼     0 with the
need to properly bound the principal damages to one for finite element computations. In the context
of quasi-brittle failure, the alternative theories are the microplane damage models, usually introdu-
cing several scalar damage variables dk, defined for fixed damage directions k all over a unit sphere
(Bažant and Gambarova, 1984; Bažant and Ozbolt, 1990; Bažant and Prat, 1988a,b; Fichant et al.,
1997; Kuhl and Ramm, 1998).
Other tensorial damage representations exist in literature such as the thermodynamics damage
1 1
variable of integrity tensor ( ¼ 1  D, its invert (1 or even their square root (2 ¼ ð1  DÞ2 ,
1 1

(2 ¼ ð1  DÞ 2 ¼ H (Badel et al., 2007; Chambart et al., 2014; Chen et al., 2011; Desmorat
et al., 2010a,b; Ladevèze, 1983, 1995; Ladevèze et al., 1994; Menzel and Steinmann, 2001;
Steinmann and Carol, 1998; Souid et al., 2009) as introduced in pioneering work of Cordebois
and Sidoroff (1982). The tensorial damage evolution laws take different forms according to the
authors, from ( _ ¼     0 to H
_ ¼     0. Note that, in a similar spirit a compliance version of
microplane damage modeling, Jirasek (1999) performed the unit sphere averaging process with the
damage evolution law applied to the inverse of the integrity tensor in order to ensure gradual
softening and numerical efficiency. As the principle of energy equivalence is used the tensor defined
is in fact tensor H. Carol et al. (2001) and Pröchtel and Häußler-Combe (2008) for instance used the
1 1
Lavedèze general framework but defined damage evolution as L_ ¼ 2 (2  ðd=dtÞ(1  (2 ¼
2H  D _  H ¼ 2ðH _  H þ H  HÞ
1 1 _ ¼     0 from the concept of pseudo-logarithmic damage
tensor rate. Lastly, some authors instead straightforwardly consider the effective (damaged) elasti-
city tensor as damage variable (Govindjee et al., 1995; Meschke et al., 1998).
One aims in the present work to model monotonic multiaxial concrete materials behavior and
loading induced anisotropic damage. It has been shown that a proper consideration of damage
anisotropy allows the number of material parameters, needed to represent a concrete response in
not too a confined state of stresses, to diminish (Desmorat, 2004; Desmorat et al., 2007). The strong
dissymmetry of tension/compression behaviors can be obtained by means of five material param-
eters, including elasticity parameters, i.e. less than in the damage model by Mazars (1984, 1986).
This property has been gained in the so-called ‘initial anisotropic damage model’:

. within the Ladevèze second-order damage thermodynamics framework;


. as required by thermodynamics (Lemaitre and Desmorat, 2005), by considering a single thermo-
dynamics variable, standard damage variable D: as a state variable it represents the micro-crack-
ing pattern whatever the loading sign;
. by means of a deviatoric/hydrostatic splitting within effective stress (Lemaitre et al., 1999, 2000;
Papa and Taliercio, 1996);
. from a full hydrostatic stiffness recovery in compressive stress states; and
. with damage intensity and anisotropy both governed by the extensions (Mazars, 1986).

The constitutive equations are recalled in the section on initial anisotropic damage model, with an
emphasis on the shear-bulk coupling. The biaxial responses of the initial model are derived next
from a closed form polar representation in principal stresses plane. The model advantages, but also
the drawbacks, are discussed so that a novel shear-bulk coupling is then proposed, in accordance
with the numerical discrete element results of Delaplace and Desmorat (2007). The main features of
the finally proposed anisotropic damage model are (i) the use of unbounded second-order tensor H

Downloaded from ijd.sagepub.com by guest on July 26, 2016


820 International Journal of Damage Mechanics 25(6)

as damage variable instead of D, (ii) that both damage level and damage anisotropy are assumed to
be governed by the extensions (Mazars, 1984) and (iii) that an improved multiaxial behavior in
confined stress states is gained from an original stress triaxiality enhancement of Mazars criterion
function for anisotropic damage. In this paper the stress triaxiality is classically defined as the ratio
hydrostatic stress sH/von Mises equivalent stress seq

1 tr r H
TX ¼ ¼ ð1Þ
3 eq eq

if p is the stress tensor.

Initial anisotropic damage model


For concrete, the microcracks due to tension are mainly orthogonal to the loading direction when
the microcracks due to compression are mainly parallel to the loading direction. The damage state
is then chosen to be represented by the tensorial variable D, bounded to a unit tensor 1. As
mentioned in the introduction the use of a second-order damage tensor is convenient for practical
applications but also for material parameters identification. The damage anisotropy induced by
either tension or compression is simply modeled by the consideration of damage evolution laws
ensuring a damage rate proportional to the positive part of the strain tensor, i.e. damage – and its
anisotropy – governed by the extensions hii ¼ max(0, i) if i stands for the eigenstrains (Mazars,
1984, 1986).

Thermodynamics framework
In the initial anisotropic damage model (Desmorat, 2004), Gibbs free enthalpy is assumed to be a
function of the stress tensor and of the second-order damage tensor, as follows
" #
1  1 1
 1 htr ri 2
 ? ðr, DÞ ¼ tr ð1  DÞ2  r0  ð1  DÞ2  r0 þ þ h tr ri2 ð2Þ
4G 18K 1  13  tr D

where E is Young’s modulus, n is Poisson’s ratio, G ¼ E/2(1 þ n) and K ¼ E/3(1  2n) are the shear
and bulk moduli, respectively, hxi ¼ max(x, 0) denotes the positive part of a scalar and
ð:Þ0 ¼ ð:Þ  13 trð:Þ 1 stands for the deviatoric part of a tensor.
A splitting between deviatoric and hydrostatic contributions has been made and the hydrostatic
term has itself been split into two parts, the part at positive hydrostatic stresses H ¼ 13 trp being
affected by damage (through its mean/hydrostatic damage value DH ¼ 13 tr D, not an additional
variable), the negative hydrostatic stresses remaining unaffected by damage.
The corresponding state laws are

@ ? @ ?
¼ Y¼ ð3Þ
@r @D

The elasticity law coupled with anisotropic damage can be recast as equations (4) and (5). See the
works of Lemaitre and Desmorat (2005) and Lemaitre and Chaboche (2009) for the derivation of the
thermodynamics force Y associated with anisotropic damage D.

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 821
The model is completed by a criterion function f  0 defining both the elasticity domain and the
mean damage growth and by a tensorial non-standard damage evolution law D _ ¼ . . .  0 ensuring
positivity of the damage rate tensor and of the intrinsic dissipation.

Constitutive equations
The full set of constitutive equations for the initial anisotropic damage model are as follows
(Chambart, 2009; Desmorat et al., 2007).

(1) Initially isotropic elasticity coupled with damage


8
>
> ~0
r
1þ  < 0 ¼
 ¼ E1 : r~ ¼ r~  tr r~ 1 i:e: 2G ð4Þ
E E >
> ~ ~H
tr r
: tr  ¼ ¼
3K K

~ H ¼ 13 tr r~ is the effective hydrostatic stress and E is isotropic Hooke’s tensor for virgin (undam-
aged) material.
(2) Effective stress
" #
 0 1 htr ri
12 0 12
r~ ¼ ð1  DÞ  r  ð1  DÞ þ  h tr ri 1 ð5Þ
3 1  13  tr D

where the material constant   1 is the hydrostatic sensitivity parameter (Lemaitre et al.,
2000).
(3) Damage criterion

f ¼ "^    0 ð6Þ
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P 2
where "^ ¼ hi i is the equivalent strain (Mazars, 1984). The consolidation function k is set as
a function of the trace of the damage tensor

  
tr D 0
 ¼ ðtr DÞ ¼ a  tan þ arctan ð7Þ
aA a

(4) Induced damage anisotropy governed by the positive – in terms of principal values – effective
strain tensor

_ ¼ h~
D _ iþ ~ ¼ E1 : r ð8Þ

In such a rate independent formulation, the damage multiplier _ satisfies Kuhn–Tucker load-
ing–unloading conditions f  0, _  0, _ f ¼ 0.

There are five material parameters introduced if  ¼ 3 is set: E, n for elasticity, k0 as damage
threshold and A and a as damage parameters.

Downloaded from ijd.sagepub.com by guest on July 26, 2016


822 International Journal of Damage Mechanics 25(6)

Mathematical and thermodynamics features


Let us first recall that the thermodynamics potential   (equation (2)) can be continuously differ-
entiated (Ladevèze, 1983; Ladevèze and Lemaitre, 1984; Desmorat and Cantournet (2008); Lemaitre
and Desmorat, 2005).
This mathematical property ensures the continuity of the model response in the nonproportional
loading cases. As a counterexample, for example, if hpiþ is the positive part of stress tensor in terms
of principal values and E~ damaged elasticity tensor, a term of the form hriþ : E~ 1 : hriþ cannot be
continuously differentiated and leads to discontinuities in multiaxial stress–strain response.
The mathematical differentiability feature can be illustrated from nonproportional numerical
tests (with rotating principal directions). The results for the Willam et al. (1987) test are given in
Figure 1. The loading is applied in two steps: (i) a first uniaxial tension is applied in direction z until
the peak in the stress–strain diagram and (ii) the second loading consists in a rotation of the strain
tensor principal axis by applying a non-proportional loading using biaxial-tension and shear (xz).
The applied increments for step (ii) are as follows, zz ¼ 12 xz and xx ¼ 34 xz . The major
principal directions angles for second-order tensors p,  and D are denoted by ’s, ’ and ’D,
respectively, with ’s ¼ ’ ¼ ’D ¼ 0 up to the end of step (i). Continuity of the stress components
(Figure 1(a)) as well as of the principal directions (Figures 1(b)) are obtained.
Lastly the damage evolution law is non-standard, i.e. it does not derive from a convex evolution
(dissipation) potential with respect to thermodynamics force Y. Therefore the proof of the positivity
of the intrinsic dissipation for the initial anisotropic damage model is to be derived and it is given in
Desmorat (2006), Desmorat et al. (2007) and Lemaitre and Chaboche (2009). The numerical
schemes for the computation of intrinsic dissipation due to anisotropic damage can be found in
Chambart et al. (2014).

Figure 1. Willam’s test results for initial the anisotropic damage model (Ragueneau et al., 2008). (a) For the con-
tinuous stresses and (b) for the continuous principal directions angles ’s, ’, ’D (angles in degree) for the stress,
strain and damage tensors, respectively.

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 823

Effective Hooke’s tensor


The elasticity law can be inverted in a closed form expression r ¼ E~ :  so that the effective
(damaged) elasticity tensor is
 
~ þ K~ 1  1 ¼ 2G ð1  DÞ1=2 ð1  DÞ1=2  ð1  DÞ  ð1  DÞ
E~ ¼ 2 G
3  tr D
  ð9Þ
1
þ K 1   tr D Hðtr Þ 1  1
3
where H(x) is the Heaviside function, and tensorial product  is defined by the identity
ðABÞ : C ¼ A  C  B for all symmetric tensors A, B, C.

Bulk modulus
One sees from equation (9) that the effective bulk modulus is
 
1
K~ ¼ K 1   tr D Hðtr Þ ¼ Kð1  DH Hðtr ÞÞ ð10Þ
3

It is constant equal to the virgin value K in compressive loadings at tr r ¼ 3K tr  5 0 (negative stress


triaxiality). This physically corresponds to a state of closed micro-cracks. At tr r ¼ 3K~ tr  4 0
(positive stress triaxiality) it is the same linear function K~ ¼ Kð1  DH Þ of the hydrostatic
damage DH ¼ 13 tr D¼ 13 ðD1 þ D þ D3 Þ whatever the stress triaxiality. It is therefore not a function
q2ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
of the quadratic norm kDk ¼ D21 þ D22 þ D33 nor of the infinite norm kDk1 ¼ maxDi, and this is for
any positive triaxiality.
The linearity of effective bulk modulus with respect to DH has been checked by Delaplace and
Desmorat (2007) – at least at low damage – for the following two states of micro-cracking:

. the left-hand curves of Figure 2 show that equation (10) is valid, perfectly for the
16  16  16 mm3 cube, up to DH ¼ 0.3; and
. the right-hand curves of Figure 2 show that K~ cannot be a function of kDk.

The uniaxial state of micro-cracking – not represented – is due to uniaxial tension loading applied
on the cubes of Figure 3 (considered as a representative volume element (RVE)), triaxial state of
micro-cracking – represented in Figure 3 – is due to equi-triaxial tension loading applied on the
cubes. Both monotonic loadings are applied numerically as discrete element computations on the
lattice medium (the cubes made of Voronoi cells and considered as the RVE). In such numerical tests
the material is described as a particles assembly representative of the material heterogeneity, the
particles being here linked by elastic–brittle beams (Herrmann and Roux, 1990; Schlangen and
Garboczi, 1997; Van Mier et al., 2002). Two sizes 8  8  8 mm3 and 16  16  16 mm3 of the
RVE of a micro-concrete were considered (Figure 3), the increase in size corresponding to an
increase in the number of particles and in the number of degrees of freedom: 512 particles and
4096 degrees of freedom for the 8-cube and 3072 particles and 24,576 degrees of freedom for the
16-cube. The crack patterns obtained at the end of the triaxial loading are those of Figures 3 for the
two samples. Note that the number of beams to break before failure varies from 1500 beams for
the 8-cube sample to 8000 for the 16-cube. The components of the damage tensor have been

Downloaded from ijd.sagepub.com by guest on July 26, 2016


824 International Journal of Damage Mechanics 25(6)

1 1

Triaxial loading Triaxial loading


Uniaxial loading Uniaxial loading
0.8 0.8

Modulus ratio K /K
Modulus ratio K /K

~
~

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 0.5 1 1.5 0 0.5 1 1.5
¬
√3 DH ||D||

1 1

Triaxial loading Triaxial loading


Uniaxial loading Uniaxial loading
0.8 0.8
Modulus ratio K /K

Modulus ratio K /K
~

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 0.5 1 1.5 0 0.5 1 1.5
¬
√3 DH ||D||

Figure
p ffiffiffi 2. Effective bulk modulus K~ from discrete element computations as a function of hydrostatic damage (of
3DH for comparison) and of quadratic norm kDk. (a) 8  8  8 mm3 cube, (b) 16  16  16 mm3 cube (from
Delaplace and Desmorat (2007)).

measured (for each mark of Figure (2) by means of repeated numerical elastic loading–unloading
sequences performed in uniaxial tension (even for the triaxial loading), using then the coupling of
elasticity with anisotropic damage given by equations (11) and (12) with one non-zero principal
stress si ¼ s, the two others sj6¼i ¼ 0.
~
The negative slope  in the K=K versus the DH diagram is found to be close to 1 (& 1.2). The
precise value is subjected to caution as it is obtained from numerical modeling (with no aggregates
for instance) but it can nevertheless be noticed that it is quite different from the slope  ¼ 3
obtained from the first set of material parameters identified for initial anisotropic damage model
(Desmorat, 2004) or for metals (Lemaitre et al., 2000).

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 825

Figure 3. Discrete element samples considered as Representative Volume Element ((a) 8  8  8 mm3 cube, (b):
16  16  16 mm3 cube) and crack patterns at the end of triaxial loading (high-damage level) for the two samples.

Note that considering  ¼ 1 in such an initial damage model means that the bulk modulus K~ fully
vanishes at tr D ¼ 3, i.e. at the maximum principal damage max Di larger than one in uniaxial and
equi-biaxial tension (K~ cannot vanish then as principal damages are bounded by 1!). This corres-
ponds to a quite high (spurious) elastic stiffness which is kept at rupture. On the other hand the
therefore preferred case  ¼ 3 leads to K~ ¼ 0 at tr D ¼ 1, i.e. at the maximum principal damage max
Di equal to 1/2 in equi-biaxial tension and equal to 1/3 in equi-triaxial tension: at values of shear
moduli far to be zero. Enforcing then K~ ¼ 0 but allowing still the damage tensor D to evolve up to
unit tensor 1 in an adequate procedure for the numerical control of rupture is a solution which leads
to numerical difficulties in finite element computations (Badel et al., 2007; Desmorat et al., 2007;
Ragueneau et al., 2008; Chambart, 2009; Leroux, 2012).
One will propose next an adequate shear-bulk coupling that coincides the full loss of both bulk
and shear stiffnesses K~ ¼ G~ ¼ 0 with no need of a procedure for the numerical control of rupture to
bound the damage eigenvalues to one.

Uniaxial and biaxial responses of initial model


In proportional loading, radial loading at constant stress triaxiality TX ¼ tr p/3seq ¼ sH/seq, the
stress, strain and damage tensors are colinear and closed form expressions can be derived.
In the principal framework the different tensors are all diagonal (of principal components si, i,
Di). The effective stress r~ is related to the stress and damage state by equation (5) and is also
diagonal. In the plane stress condition (s3 ¼ 0) its deviatoric part is given by
   
1 4 1 1 2 4 2 1
~ 10¼ þ þ   þ
9 1  D1 1  D2 1  D3 9 1  D1 1  D2 1  D3
   
0 1 2 4 1 2 1 4 1
~ 2 ¼   þ þ þ þ þ ð11Þ
9 1  D1 1  D2 1  D3 9 1  D1 1  D2 1  D3
   
0 1 1 2 2 2 1 2 2
~ 3 ¼   þ þ   þ þ
9 1  D1 1  D2 1  D3 9 1  D1 1  D2 1  D3

Downloaded from ijd.sagepub.com by guest on July 26, 2016


826 International Journal of Damage Mechanics 25(6)

when its hydrostatic part is given by

1 þ 2
~ H ¼ ð12Þ
3ð1  DH Þ

The elasticity law coupled with damage 0 ¼ r~ 0 =2G, tr  ¼ ~ H =K gives, for an in-plane stress loading
s1 ¼ cos , s2 ¼ sin interpreting  as an equivalent polar stress


"i ¼ Bi ð , Di Þ ð13Þ
E

with Bi a function of the parametric angle in the stress plane and of the principal damages Di given
in Appendix 1.
Similar but simpler calculations give the effective strain principal components as

 ~
"~i ¼ Bi ð Þ B~ i ð Þ ¼ Bi ð , Di ¼ 0Þ ð14Þ
E

_ ¼ _h~ iþ becomes
The damage evolution law D

_

D_ i ¼ _h"~i i ¼ hB~ i i ð15Þ
E

and can be integrated at constant in

h"~i i hB~ i i
Di ¼ tr D ¼ tr D ð16Þ
h"~1 i þ h"~2 i þ h"~3 i hB~ 1 i þ hB~ 2 i þ hB~ 3 i
The consistency condition
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

f ¼ ^   ¼ hB1 i2 þ hB2 i2 þ hB3 i2  ðtr DÞ ¼ 0 ð17Þ
E

allows the determination of the equivalent polar stress  ¼ ( , Di) as a function of angle and
principal damages Di.
Previous derivations analytically give the model response in proportional biaxial loading by
proceeding as follows.

(1) Consider a loading biaxiality through angle (constant for each proportional loading calcula-
tions) and any given value for the trace of the damage tensor tr D, starting from tr D ¼ 0.
(2) Calculate the principal damage components

hB~ i i
Di ¼ tr D ð18Þ
hB~ 1 i þ hB~ 2 i þ hB~ 3 i

as B~ i only depends on .

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 827
(3) Calculate the equivalent polar stress

E ðtr DÞ
 ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð19Þ
hB1 i þ hB2 i2 þ hB3 i2
2

with Bi ¼ Bi( , Di).


(4) Calculate the stress components

1 ¼  cos 2 ¼  sin ð20Þ

(5) Calculate the strain components


"i ¼ Bi ð21Þ
E

The model responses are plotted next. The material parameters are representative of a concrete
(peak stress in tension ut ¼ 3:5 MPa and peak stress in compression uc ¼ 30:5 MPa)

E ¼ 37, 000 MPa,  ¼ 0:2,  ¼ 3, 0 ¼ 5 105 , A ¼ 5000, a ¼ 3 103

Uniaxial tension
The case of tension – performed in direction 1 – corresponds to ¼ 0 (stress triaxiality TX ¼ 1/3). The
stress–strain response is given in Figure 4(a). The different curves s(1) and s(2) correspond to the

(a) (b)

Figure 4. Response of initial anisotropic damage model (a) in tension, (b) in equi-biaxial tension.

Downloaded from ijd.sagepub.com by guest on July 26, 2016


828 International Journal of Damage Mechanics 25(6)

tensile strain 1 and to the transverse strain 2 ¼ 3 abscissa. The damage state is simply D1  0,
D2 ¼ D3 ¼ 0.

Equi-biaxial tension
Equi-biaxial tension corresponds to ¼ p/4 (stress triaxiality TX ¼ 2/3). The stress s ¼ s1 ¼ s2 versus
the strain  ¼ 1 ¼ 2 curve (Figure 4(b)) exhibits a peak a bit lower than in tension (3.45 instead of
3.5 MPa) and a more brittle behavior, as expected for a quasi-brittle material. The damage state is
D1 ¼ D2  0, D3 ¼ 0.

Uniaxial compression
The stress strain responses s(i) and s(v) for compression in direction 1 ( ¼ p, stress triaxiality
TX ¼  1/3) are given in Figure 5, denoting v ¼ tr  ¼ 1 þ 22 the volumetric strain. As expected, the
s(v) response is found linear (s ¼ tr p ¼ Kv) due to the deviatoric/hydrostatic splitting with no
effect of the damage on bulk modulus in compressive states. The factor of around 10 between the
peak stresses in tension and in compression, usual for concrete, is obtained. The tension/compres-
sion dissymmetric behavior is mainly due to damage anisotropy: micro-cracks perpendicular to
loading direction in tension (D1 > 0, D2 ¼ D3 ¼ 0), micro-cracks parallel to loading direction in
compression D1 ¼ 0, D2 ¼ D3 > 0.

Equi-biaxial compression
The response in equi-biaxial compression ( ¼  3p/4, stress triaxiality TX ¼  2/3, see Figure 6) is
found to be much too brittle (a snapback is exhibited) with a much too low peak stress (6.5 MPa), as
for concrete it is usually larger than compression peak stress ut (Kupfer et al., 1969). This feature has
been pointed out in earlier works. Modeling improvements, not fully satisfactory due to their com-
plexity, have been proposed in Ragueneau et al. (2008) and Leroux (2012). The damage state is
D1 ¼ D2 ¼ 0, out of plane damage D3  0.

Figure 5. Response of initial anisotropic damage model in compression.

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 829

Figure 6. Response of initial anisotropic damage model in equi-biaxial compression.

(a) (b)

Figure 7. (a) Response of initial anisotropic damage model in shear and (b) microcracking pattern corresponding to
damage state D1  0, D2 ¼ 0 in shear.

Shear
The shear stress ¼ s1 versus shear strain 12 ¼ 1 curve is plotted in Figure 7 (case ¼  p/4, stress
triaxiality TX ¼ 0). It exhibits a peak stress of the same order of magnitude as the one in tension as
for the well established isotropic damage model (Mazars, 1984, 1986); but the present model does
not give any ductility at all in shear (one would have expected a plateau or a slow softening more
representative of friction-wear behavior).

Downloaded from ijd.sagepub.com by guest on July 26, 2016


830 International Journal of Damage Mechanics 25(6)
Discussion
This initial anisotropic damage model, dedicated to initially isotropic quasi-brittle materials such as
concrete, has the following nice properties.

(1) It represents the damage state encountered in tension (micro-cracking D1, D2 ¼ D3 ¼ 0, parallel
perpendicular to loading direction 1), in compression (micro-cracking D2 ¼ D3, D1 ¼ 0, parallel
to loading direction 1), in bicompression (out of plane micro-cracking D3, D1 ¼ D2 ¼ 0) and in
shear.
(2) The damage anisotropy is itself responsible for the dissymmetry of tension and compression
responses, as observed experimentally.
(3) There is only one damage variable, tensorial, for all loadings including tension and compression,
as requested by the status of a state variable of thermodynamics: D represents the micro-crack-
ing pattern present in the material whatever the loading sign.
(4) As a consequence the number of material parameters is quite low: five including elasticity par-
ameters if  ¼ 3 is reasonably set.
(5) It can be proven,thanks to the deviatoric/hydrostatic splitting in the definition of the effective
stress (5), that dissipation due to damage is positive in any case (Desmorat, 2006).
(6) The total softening of all stress components is obtained in three dimensions at positive stress
triaxiality (this property is again due to the deviatoric/hydrostatic splitting in the definition of
effective stress). At negative stress triaxiality bulk modulus is unchanged, tr p ¼ 3K tr  < 0, and
only (all) deviatoric stress components ij0 soften to zero.

There are of course drawbacks. The main one is the response in equi-bicompression which is
way too brittle. One can argue that the behavior in confined states shall not been modeled by
elasticity coupled with damage only and that the plasticity and permanent strains have to take
over (Burlion et al., 2002; Gatuingt et al., 2002; Govindjee et al., 1995; Grassl and Jirasek, 2006;
Meschke et al., 1998). Nevertheless an elasticity coupled with a damage model that includes
acceptable monotonic responses in confined stress states would be appreciated (see for instance
the work of Leroux (2012)). A second drawback is that the post-peak response has no tail.
Maximum principal damage reaches one at finite rupture strain, whose value is a bit small in
tensile cases. Brittleness is physical but it leads to costly numerical difficulties. One needs a specific
numerical control of rupture (Badel et al., 2007; Chambart, 2009; Desmorat et al., 2007) in order
to enforce the principal damages to remain bounded to one. In the case of plain concrete appli-
cations this works well but difficulties arise at concrete/bars sheared interfaces in reinforced con-
crete structures (Leroux, 2012). Related to this control of rupture procedure, there is the fact that
full softening at stress triaxiality larger than 1/3, such as in the equi-biaxial tension case, occurs at
principal damages Di < 1, strictly smaller than one. This is due to shear-bulk coupling considered
1
as damage D acts on both shear modulus (in a tensorial manner as ð1  DÞ2 terms, see equation
(9)) and bulk modulus as K~ ¼ Kð1  tr DÞ.
The next section is dedicated to the proposal of an anisotropic damage model that attempts to
keep the advantages of the initial model and to correct its drawbacks. The validity domain sought
still consists in monotonic applications at not too high triaxial confining pressures, so that one
allows us not to model the irreversible strains, neither the volumetric dilatancy in simple com-
pression, nor all complex cyclic effects, see for instance Bažant and Prat (1988a,b), Desmorat
(2004), Goidescu et al. (2015), Halm and Dragon (1998), Lebon (2011), Ragueneau et al. (2000),
Ramtani et al. (1992), Richard and Ragueneau (2013) and Souid et al. (2009) in the case of
tensorial damage.

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 831
Model with modified shear-bulk coupling
1
The change of second-order tensorial damage variable H ¼ ð1  DÞ2 is operated and the damage
dependency of deviatoric and of hydrostatic parts of novel effective stress (23) are both made homo-
geneous to H2. This allows the consideration of an unbounded damage evolution up to H ! 1 for any
loading case. The same damage variable H is chosen to act on both shear and bulk parts, represen-
tative of so-called shear-bulk coupling. There is no need to refer anymore to the standard damage
variable D but it is nevertheless convenient to still calculate it (as D ¼ 1  H2) as principal damages
Di 2 [0, 1] are more easily interpreted than unbounded principal values Hi 2 [1,1[ of tensor H. The ‘no
damage’ case corresponds to H ¼ 1 , D ¼ 0 or Hij ¼ dij , Dij ¼ 0 in terms of components.

Constitutive equations
The proposed anisotropic damage constitutive equations for quasi-brittle materials, such as con-
crete, are as follows.

(1) Initially isotropic elasticity coupled with damage

r~ 0 1
¼ þ tr r~ 1 ð22Þ
2G 9K

(2) Effective stress


 
0 1 1
r~ ¼ ðH  r0  HÞ þ tr H2 htr ri  h tr ri 1 ð23Þ
3 3

(3) Damage criterion


f ¼ "^    0 ð24Þ
where "^ is the Mazars equivalent strain.
(4) The consolidation function is taken to be linear in tr H as

 ¼ 0 þ SRs ðtr H  3Þ ð25Þ

with, as material parameters, the damage threshold k0, the triaxiality exponent s and the damage
strength S, enhanced by the (negative) stress triaxiality TX by means of the triaxiality function
Rn (Lemaitre and Chaboche (1985), see Appendix 2), here normed to unity in shear and possibly
bounded to the material constant B 4 Rc ¼ 3=2ð1 þ Þ in order to properly model biaxial
compression
 
9 1  2 2 H
R ¼ min 1 þ hTX i , B TX ¼ ð26Þ
2 1þ eq

(5) Induced damage anisotropy governed by the positive – in terms of principal values – effective
strain tensor h~iþ but written in terms of rate of damage tensor H

_ ¼ h~
H _  iþ ~ ¼ E1 : r ð27Þ

Downloaded from ijd.sagepub.com by guest on July 26, 2016


832 International Journal of Damage Mechanics 25(6)

The damage multiplier _ satisfies the Kuhn–Tucker loading–unloading conditions f  0, _  0,


_ f ¼ 0.

The stress triaxiality is TX ¼ 1/3 in tension leading to the value Rt ¼ 1 for the triaxiality function,
it is TX ¼ 2/3 in equi-biaxial tension so Rn ¼ 1, it is TX ¼ 0 in shear with Rn ¼ 1, TX ¼  1/3 in
compression with R ¼ Rc ¼ 3=2ð1 þ Þ 4 1 and TX ¼  2/3 in equi-biaxial compression in
which case R ¼ 3ð1  Þ=ð1 þ Þ 4 Rc 4 1. The standard definition for the stress triaxiality function
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
is preferred here to the definition of the equivalent strain reducing factor ¼ hriþ : hriþ =h tr ri
introduced by Mazars and coworkers in the criterion function f (La Borderie, 2003; Mazars, 1984;
Pontiroli, 1995).
There are five or six material parameters introduced, depending on whether the bounding value B
is introduced or not: E, n for elasticity, k0 as the damage threshold, the damage strength S, the
damage triaxiality exponent s and the bicompression parameter B. Their number is quite low for
constitutive equations attempting to properly model concrete multiaxial behavior with dissymmetry
tension/compression. Again the key point of modeling is the fact that such a dissymmetry is due to
1
damage anisotropy with a tensorial damage state represented by a single variable H ¼ ð1  DÞ2 .
There is no – thermodynamically inconsistent – use of a first damage variable for ‘tension’ and of a
second damage variable for ‘compression’.

Effective Hooke’s tensor


The elasticity law can be inverted in a closed form expression r ¼ E~ :  so that the effective
(damaged) elasticity tensor is at positive stress triaxiality TX (positive tr p and tr )

 
H2  H2 3K
E~ ¼ 2G H1 H1  þ 11 ð28Þ
tr H2 tr H2

It tends toward 0 in the limiting case H ! 1.


At negative TX (negative tr p and tr )
 
H2  H2
E~ ¼ 2G H1 H1  þK1  1 ð29Þ
tr H2

It tends toward K 1  1 in the limiting case H ! 1.

Bulk modulus
The bulk modulus K~ ¼ K remains unchanged at negative stress triaxiality TX (H ¼ 13 tr r 5 0, v ¼ tr
 < 0).
One sees from equation (28) that the effective (damaged) bulk modulus at positive stress triaxi-
ality (sH > 0, v > 0) is

3K 3K
K~ ¼ 2
¼ ð30Þ
tr H trð1  DÞ1

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 833

Figure 8. Effective bulk modulus K~ from shear-bulk coupling (23)–(30) as a function of hydrostatic damage DH and of
quadratic norm kDk with D ¼ 1  H2.

It decreases with damage as plotted in Figure 8, where the hydrostatic damage DH in terms of D has
been calculated from the knowledge of the damage tensor H

1 1
DH ¼ tr D ¼ 1  tr H2 ð31Þ
3 3

This gives a nonlinear variation in most cases including uniaxial tension H1  1, H2 ¼ H3 ¼ 1. Equi-
triaxial tension case corresponds to spherical damage tensors of principal components

 
1 1 1 3 2 1
Hi ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ tr H Di ¼ 1  ¼ 1  ¼ tr D ¼ DH ð32Þ
1  Di 3 H2i tr H 3

so that in such a particular case one recovers a linear loss of stiffness K~ ¼ Kð1  DH Þ which corres-
ponds to the hydrostatic sensitivity parameter  ¼ 1 in initial anisotropic damage modeling. The loss
of bulk modulus is found to be very similar to the one Figure 2 obtained from discrete element
computations.
Altogether vanishing shear and bulk stiffnesses are gained from proposed shear-bulk coupling (at
infinite strain as illustrated in the examples in the section on responses of the proposed anisotropic
model).

Responses of the proposed anisotropic model


Let us consider cases of proportional loading and again use the plane stress polar representation
s1 ¼ cos and s2 ¼ sin .

Downloaded from ijd.sagepub.com by guest on July 26, 2016


834 International Journal of Damage Mechanics 25(6)

Following the same formal derivation than for the initial anisotropic damage model one sets
for elasticity law coupledffi with damage (see Appendix 3 for functions Ci and C~ i ,
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Ceq ¼ hC1 i þ hC2 i2 þ hC3 i2 )
2

 
"i ¼ Ci ð , Hi Þ "^ ¼ Ceq ð , Hi Þ ð33Þ
E E

and for the effective strain ~ ¼ E1:p

 ~
"~i ¼ Ci ð Þ C~ i ¼ Ci ð , Hi ¼ 1Þ ð34Þ
E

_ ¼ h~
The damage evolution law H _ iþ becomes in proportional loading

_

H_ i ¼ _h"~i i ¼ hC~ i i ð35Þ
E

and can be integrated in

h"~i i hC~ i i
Hi ¼ 1 þ ðtr H  3Þ ¼ 1 þ ðtr H  3Þ ð36Þ
h"~1 i þ h"~2 i þ h"~3 i hC~ 1 i þ hC~ 2 i þ hC~ 3 i
The consistency condition then gives, with the triaxiality function Rn ¼ Rn(TX( ))

f ¼ "^  ð , tr HÞ ¼ 0 ð , tr HÞ ¼ 0 þ SRs ðtr H  3Þ ¼ 0 ð37Þ

In the same manner as for the initial anisotropic damage model, proceed as follows to calculate
the model response in proportional biaxial loading.

(1) Consider a loading biaxility through angle and any given value for tr H, starting from tr H ¼ 3.
The stress triaxiality is

1 1 þ 2 1 sin þ cos
TX ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ TX ð Þ ð38Þ
3  2    þ  2 3 1  sin cos
1 1 2 2

(2) Calculate the principal damage components

hC~ i i
Hi ¼ 1 þ ðtr H  3Þ ð39Þ
hC~ 1 i þ hC~ 2 i þ hC~ 3 i
as C~ i only depends on .
(3) Calculate the equivalent polar stress

E ð , tr HÞ
¼ ð40Þ
Ceq

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 835
due to Ceq ¼ Ceq( , Hi).
(4) Calculate the stress components

1 ¼  cos 2 ¼  sin ð41Þ

(5) Calculate the strain components



"i ¼ Ci ð42Þ
E

One cross-identifies the proposed anisotropic model on the initial tension and compression
responses of initial model, equations (4)–(8). As already mentioned the purpose is to avoid the
sudden softening to zero stress and to gain ductility, as a tail decreasing gently to zero stress in
stress–strain diagrams. The peak stresses are enforced identical for both models in tension
(ut ¼ 3:5 MPa) and in compression (uc ¼ 30:5 MPa).
The material parameters for concrete are

5
E ¼ 37, 000 MPa,  ¼ 0:2, 0 ¼ 9 105 , S ¼ 1:45 104 , s ¼ 4:9, B ¼
3

The responses of initial anisotropic damage model are reported as dashed lines in the next figures
(Figures 9 to 13).

Uniaxial tension
The response to the uniaxial tension, at polar angle ¼ 0 and stress triaxiality TX ¼ 1/3, is given in
Figure 9(a) and (b). The peak stress is ut ¼ 3:5 MPa as for initial damage model. A larger strain

(a) (b)

Figure 9. (a) Response of proposed anisotropic damage model in tension (dashed line: initial anisotropic damage
model). A larger strain scale is used in (b).

Downloaded from ijd.sagepub.com by guest on July 26, 2016


836 International Journal of Damage Mechanics 25(6)

scale is considered in Figure 9(b) in order to exhibit the gain in ductility and the tail obtained at high
softening (the stress nevertheless tends to zero at infinite strain).
The micro-cracking pattern characteristic of tension performed in direction 1 is recovered as
H2 ¼ H3 ¼ 1 (equivalent to D2 ¼ D3 ¼ 0) and

1 1 1
H1 ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ tr H  2 4 1 D1 ¼ 1  2
¼1 40 ð43Þ
1  D1 H1 ðtr H  2Þ2

Biaxial tension
The same peak stress as the initial damage model (still with more ductility, see Figure 10) is obtained
in the equi-biaxial tension case ¼ p/4, TX ¼ 2/3. The damage state is H1 ¼ H2 ¼ 12 ðtr H  1Þ 4 1,
H3 ¼ 1 equivalent to D1 ¼ D2 > 0, D3 ¼ 0.

Uniaxial compression
Compression in direction 1 is the case ¼ p and stress triaxiality TX ¼  1/3 (Figure 11).
Cross-identification on the compression response can be performed up to the softening stage at a
high-damage level (up to D2 ¼ D3 & 0.9). The fact that the stress has not softened by 90% at this
damage level is due both to damage anisotropy (damages D2 and D3 are not in the loading direction)
and to unchanged (undamaged) bulk modulus in compression K~ ¼ K. Such a last model feature is
reflected by a linear response in terms of the volumetric stress–strain curve s(v) ¼ Kv ¼ K tr  in
Figure 11. A softening tail is obtained for the s(1) response but its role in compression (as well as in

Figure 10. Response of proposed anisotropic damage model in equi-biaxial tension (dashed line: initial anisotropic
damage model).

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 837

Figure 11. Response of proposed anisotropic damage model in compression (dashed line: initial anisotropic damage
model).

bicompression) is much less important for most practical applications than at positive or zero stress
triaxiality.
The micro-cracking pattern in compression D1 ¼ 0, D2 ¼ D3 > 0 is obtained as

1
H1 ¼ 1 and H2 ¼ H3 ¼ ðtr H  1Þ 4 1 ð44Þ
2

corresponding to the standard damages


 2
1 2
D2 ¼ D3 ¼ 1  ¼ 1  40 ð45Þ
H22 tr H  1

Biaxial compression
Model response in equi-biaxial compression is plotted in Figure 12 for different values of bounding
material parameter B. Poisson’s ratio is n ¼ 0.2 so that the same response is obtained for any B  2
than for unbounded case B ! 1. One clearly sees that the parameter B allows any physical bicom-
pression response to be represented, depending on the concrete considered: it is named bicompres-
sion parameter. Be careful to just chose B > 3/2(1 þ n) ¼ 1.25 so that the response in uniaxial
compression remains unchanged (as well as for all the other model responses at larger stress triaxi-
ality TX   1/3 plotted in this paper). The choice of B ¼ 5/3 is found to be consistent with the
multiaxial experiments (Kupfer et al., 1969) exhibiting a ratio peak stress in bicompression/peak

Downloaded from ijd.sagepub.com by guest on July 26, 2016


838 International Journal of Damage Mechanics 25(6)

Figure 12. Response of proposed anisotropic damage model in equi-biaxial compression for different values of
bicompression parameter B (dashed line: initial anisotropic damage model, thick line: unbounded Rn case B ! 1 or
any case B  2).

stress in tension of approximately 1.15, i.e. here the peak stress is 35 MPa in bicompression.
The choice B ¼ 1.725 is consistent with a ratio of 1.35 obtained by Yin et al. (1989) (a peak stress
of 41 MPa in bicompression).
The physical damage state of the initial anisotropic model D3 > 0, D1 ¼ D2 ¼ 0 is recovered from
1
the principal out of plane damage H3 ¼ ð1  D3 Þ2 ¼ tr H  2 4 1 and in-plane damages
H1 ¼ H2 ¼ 1.

Shear
The shear response, ¼  p/4 and TX ¼ 0, is plotted in Figure 13. The peak stress satisfactory is of
the same magnitude for the peak stress in tension and the peak stress in shear, calculated from the
initial anisotropic damage model. The nice feature of a softening tail is obtained. Note that a non-
zero residual shear strain is still present (with value ¼ 0.1 MPa) at quite an important strain value,
12 ¼ 102.

Lode angle dependency


The general elasticity criterions for isotropic materials depends on the three stress invariants
I1 ¼ tr p, J2 ¼ 12 tr r0 2 ¼ eq
2
=3, J3 ¼ 13 tr r0 3 ¼ det r0 or in a equivalent manner of von Mises stress
seq, of stress triaxiality TX ¼ 13 tr r/seq and of Lode angle  2 0, 3 defined as

1 27 det r0
 ¼ arccos 3
ð46Þ
3 2 eq

not to be confused with the previous polar angle . The principal deviatoric stresses are then
i0 ¼ 23 eq cos i where 1 ¼ , 2 ¼   2p/3, 3 ¼  þ 2p/3, so that in principal basis diagonal

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 839
(a) (b)

Figure 13. (a) Response of proposed anisotropic damage model in shear (dashed curve: initial anisotropic damage
model). A larger strain scale is used in (b).

deviatoric stress is expressed as


2 3
cos 1 0 0
2 6 7
r0 ¼ eq nðÞ nðÞ ¼ 4 0 cos 2 0 5 ð47Þ
3
0 0 cos 3

with tensor n() deviatoric as tr n() ¼ 0. This exhibits the feature that Mazars initial elasti-
city surface, f ¼ "^  0 ¼ 0, built from the positive part of the principal strains within defin-
pffiffiffiffiffiffiffi
^ is a function of the second (von Mises) stress invariant eq ¼ 3J2
ition of equivalent strain ",
but also on both stress triaxiality and Lode angle. Mazars elasticity surface has the parametric
representation

2 E0 cos i
i ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð48Þ
3 he~1 ð, TX Þi2 þ he~2 ð, TX Þi2 þ he~3 ð, TX Þi2

in the principal stress space (s1, s2, s3).


In both anisotropic damage models studied the tensorial damage evolution is governed by the
effective positive strain tensor h~iþ of principal components

i0 tr r eq 2
h~i i ¼ þ ¼ he~i i e~i ¼ ð1 þ Þ cos i þ ð1  2ÞTX ð49Þ
2G 9K E 3

so that the rates of principal damages H_ i ¼ h


_ "~i i ¼ ð
_ eq =EÞhe~i i and even their mean value
_ are Lode angle dependent. This dependency for HH can be defined from
H_ H ¼ 13 tr H

Downloaded from ijd.sagepub.com by guest on July 26, 2016


840 International Journal of Damage Mechanics 25(6)

f ¼ "^   ¼ 0 as
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P
1X eq hei ð, TX , HÞi2  E0
HH ¼ Hi ¼ 1 þ ð50Þ
3 3ESRs
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P
due to "^ ¼ ðeq =EÞ hei ð, TX , HÞi2 if the ei are the principal values of tensor
 
2 1
e ¼ ð1 þ ÞðH  nðÞ  HÞ0 þ ð1  2Þ tr H2 hTX i  hTX i 1 ð51Þ
3 3

Nevertheless the Lode angle dependency defined here is not specific to the damage models con-
sidered in this paper: it is included in most anisotropic damage models, as stated in Desmorat (2012)
for a J2-plasticity model coupled with anisotropic damage.

Thermodynamics of second-order anisotropic damage


Let us now derive the thermodynamics framework for the novel anisotropic damage model.

Thermodynamics potential and state laws


The Gibbs free enthalphy density  , where  is the density, for the proposed anisotropic damage
model is a function of the stress and damage tensors, p and H, respectively, considered as a thermo-
dynamics state variables
 
1 1 1
 ?
ðr, DÞ ¼ trðH  r0  H  r0 Þ þ tr H2 htr ri2 þ h tr ri2 ð52Þ
4G 18K 3

The thermodynamics potential   can be continuously differentiated (Desmorat, 2000; Ladevèze


and Lemaitre, 1984; Lemaitre and Desmorat, 2005; Lemaitre et al., 2000). This ensures continuous
stress–strain response for any multiaxial nonproportional loading (as exhibited in the test by Willam
et al. (1987), shown in Figure 1) for the initial damage model).
The state laws are

@ ? @ ?
¼ Z¼ ð53Þ
@r @H

They lead to the following.

(1) The elasticity coupled with damage ((.)0 denotes deviatoric part)
 
1 0 0 1 1 2
¼ ðH  r  HÞ þ tr H htr ri  h tr ri 1 ð54Þ
2G 9K 3

This elasticity law coupled with damage defines the effective stress r~ from equation (22) and can
be recast as equations (22) and (23).

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 841
(2) The definition of thermodynamics force associated with damage H

1 0 1
Z¼ r  H  r0 þ htr ri2 H ð55Þ
2G 27K

Convexity with respect to the stress tensor and to the damage variable H
The second derivative of the thermodynamics potential with respect to the stress tensor is (where H
is the Heaviside function and the tensorial product  is HH : r ¼ H  p  H and the fourth-order
tensor HH is symmetric)
 
@2 ? 1 1 1 
E~ 1 ¼  ¼ HH  ðH  1 þ 1  HÞ þ tr H2 1  1
@r@r 2G 3 9
  ð56Þ
1 1
þ tr H2 Hðtr rÞ þ Hð tr rÞ 1  1
9K 3

It is the invert of the effective elasticity tensor. Convexity of thermodynamics potential   with
respect to the stress tensor is ensured – through deviatoric/hydrostatic splitting – by positivity of
E~ 1 ðXÞ for any non-zero symmetric second-order tensor X
 
1 0 1 1
X : E~ 1 : X ¼ X : ðH  X0  HÞ þ tr H2 htr Xi  h tr Xi tr X
2G 3 3
 
1 0 0 1 1 2 2 2
¼ trðX  H  X  HÞ þ tr H htr Xi þ h tr Xi ð57Þ
2G 27K 3
 
1 0 2 1 1 2 2 2
¼ trðX  HÞ þ tr H htr Xi þ h tr Xi 4 0
2G 27K 3

This proves convexity with respect to p as X : E~ 1 : X ¼ 2 ? ðX, HÞ 4 0 for any symmetric X 6¼ 0.


The convexity of   with respect to damage tensor H is also obtained as

@2 ? 1 0 0 1
 ¼ r r þ htr ri2 1 ð58Þ
@H@H 2G 27K

and for any X 6¼ 0

@2 ? 1 2 1
X: :X¼ trðr0  XÞ þ htr ri2 ðtr XÞ2 4 0 ð59Þ
@H@H 2G 27K

Convexity of free enthalpy with respect to the damage variable is not necessary (both state potentials
for initial anisotropic model and for novel anisotropic model are not convex with respect to damage
D). Convexity with respect to H prevents instabilities in the damage evolution driven by its asso-
ciated thermodynamics force (in the so-called standard generalized materials framework (Halphen

Downloaded from ijd.sagepub.com by guest on July 26, 2016


842 International Journal of Damage Mechanics 25(6)

and Nguyen, 1975)), whatever the loading intensity. Such a mathematical property was, for instance,
sought by Badel et al. (2007) in the case of anisotropic damage.

Non-standard anisotropic damage evolution law


The Mazars criterion function f ¼ "^   is used. It is not expressed as a function of Z, the thermo-
dynamics force associated with damage, and the damage evolution law (27) does not derive from an
evolution (pseudo-dissipation) potential whereas it does in the pioneering work by Ladevèze (1983),
who considered damage criterion fZ extended to anisotropy and expressed in terms of the thermo-
dynamics force Z

pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
fZ ¼ _ ¼ _ @fZ
trðZ  H  ZÞ þ k H : Z  k0  0 H ð60Þ
@Z

with k0, k1 as material parameter and where _ is the damage multiplier satisfying Khun–Tucker
loading–unloading conditions fZ  0, _  0 _ fZ ¼ 0. Applications to ceramics matrix composites
can be found in Ladevèze (1995) and Ladevèze et al. (1994).
The anisotropic damage model proposed uses the evolution law H _ ¼ h~
_ iþ , it is therefore non-
standard and the positivity of the dissipation is not guaranteed: it has to be proven (done in a further
section ‘‘Positivity of intrinsic dissipation’’). The damage multiplier _  0 is determined from con-
sistency condition f ¼ "^   ¼ 0 and f_ ¼ "_^  _ ¼ 0

tr H_ 1 d s 
_ ¼ þ ¼ R ^ ð61Þ
trh~i S trh~i dt 
þ

At positive stress triaxiality TX (i.e. at tr p  0, Rn ¼ 1) or in the case of compressive loading at


negative TX (tr p  0), but in the second case by neglecting the triaxiality change over a time incre-
ment, one has

1 _^ 1 hiþ : _
_ ¼ þ ¼ ð62Þ
SRs trh~i SRs ^ trh~iþ

so the damage law is defined as

_ ¼ 1 _^ 1 h~iþ hiþ : _


H þ ¼ ð63Þ
s
SR trh~i SRs trh~iþ ^

The tangent operator, such as r_ ¼ L : _ , can be derived using damaged elastic tensor (28) and
_  H1 and ðd=dtÞH2 ¼ H2  H  H
derivatives ðd=dtÞH1 ¼ H1  H _ þH_  H  H2

" Sym #
2
2 h~iþ h~iþ 1 H
L ¼ E~  r0   H1  trðr0  þH Þ
SRs trh~iþ trh~i tr H2
ð64Þ
2 H h~iþ hiþ
 htr ri 2
: þ 1
3S tr H trh~i ^

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 843
showing then that L ! 0 at high damage H ! 1 in the case of positive stress triaxiality TX (tr p  0),
but not at the negative stress triaxiality at which the bulk modulus remains constant in the effective
Hooke’s tensor E.~
The tangent operator is not symmetric, i.e. Lklij 6¼ Lijkl.

Positivity of intrinsic dissipation


The intrinsic dissipation has for expression D ¼ ð@ ? _ ¼Z:H
Þ=ð@HÞ : H _ (Desmorat, 2000, 2006;
Lemaitre and Chaboche, 2009) so that
 
1 0 1 _ ¼A:H
_ 0
D¼ r  H  r0 þ htr ri2 H : H ð65Þ
2G 27K

The term within the brackets is the sum of two positive symmetric second-order tensors
(1/2G)p0 Hp0 and (1/27K) htr pi2 H, as H is a positive symmetric tensor. The scalar product A : H _
of two positive tensors is positive (see Appendix 4) so is the dissipation, proving D  0 and fulfilling
the second principle of thermodynamics, and this for any nonproportional, possibly multiaxial,
loading.

Numerical implementation and nonlocal regularizations


Numerical scheme for finite element implementation
The time integration procedure for the full anisotropic damage model takes place at the Gauss
points of a finite element code and solves in a coupled manner the constitutive equations in the
sections on the model with modified shear-bulk coupling and constitutive equations. The strain
nþ1 ¼ (tnþ1) at time tnþ1, the stress triaxiality TX n and the damages Hn, Dn at time tn are the
inputs of the procedure. The outputs are the stresses pnþ1 and the damages Hnþ1, Dnþ1. In order
to integrate the damage model for loading paths more complex than those in the section on
Responses of proposed anisotropic model – for instance nonproportional loading – proceed as
the following steps.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P ffi
(1) Compute the equivalent strain "^nþ1 ¼ h"i nþ1 i2 from eigenstrains i nþ1.
(2) Make a test on the criterion function f ¼ "^nþ1  n , n ¼ 0 þ SRsn ðtr Hn  3Þ.
(a) If f  0, the material behaves elastically, then set Hnþ1 ¼ Hn, Dnþ1 ¼ Dn.
(b) If f > 0, the damage must be corrected by using the damage evolution law discretized as
H ¼ Hnþ1  Hn ¼  h~n iþ . with the damage direction evaluated at time step tn, ~ n ¼ E1:pn.
where E is the isotropic undamaged Hooke’s tensor. The expression for the damage multiplier
increment   is

tr Hnþ1  tr Hn "^nþ1  0
 ¼ where tr Hnþ1 ¼ 3 þ
trh~n iþ SRsn

and the actualization of the damage tensors H and D is

Hnþ1 ¼ Hn þ  h~n iþ Dnþ1 ¼ 1  H2


nþ1

Downloaded from ijd.sagepub.com by guest on July 26, 2016


844 International Journal of Damage Mechanics 25(6)

(2) Compute the stresses at the end of the increment pnþ1 by first using the elasticity law r~ nþ1 ¼ E:
nþ1 and then by inverting the equation (23) of effective stress
" #
H2 ~ nþ1 2
nþ1 : r 1 3htr r~ nþ1 i
rnþ1 ¼ H1
nþ1  r~ nþ1  H1
nþ1  Hnþ1 þ  h tr r~ nþ1 i 1
tr H2
nþ1
3 tr H2nþ1

There is no need for a specific procedure for the numerical control of rupture to the bound damage
tensor D to unity as the principal damages Di in novel model does not reach one at finite strain. For
a nonlocal implementation, simply replace the local Mazars strain "^nþ1 by its nonlocal averaging
"^nl
nþ1 at time tnþ1 (see next subsection), the test of step 2 then being made on the criterion function
f ¼ "^nl
nþ1  n and the damage multiplier reading

tr Hnþ1  tr Hn "^nl
nþ1  0
 ¼ wherenow tr Hnþ1 ¼ 3 þ
trh~n iþ SRsn

Nonlocal, nonlocal with internal time and eikonal nonlocal regularizations


Spurious mesh dependency is obtained in computations with local damage (softening) models. In
dynamics visco-damage framework allows the obtaining of mesh independency if the damage rate is
bounded (Allix, 2013; Allix and Deü, 1997). See Gatuingt (2008), Chambart (2009), Desmorat et al.
(2010b) and Leroux (2012) for the extension to anisotropic damage including the modeling of the
strain rate effect of concrete material.
In the present rate independent model formulation, a nonlocal averaging of the equivalent strain
is needed to ensure mesh independency, the criterion function being changed into

f ¼ "^nl    0 ð66Þ

with "^nl nonlocal equivalent strain. Different choices are possible, either taking advantage or not of
damage anisotropy.

(1) The use of standard nonlocal integral equivalent strain (Pijaudier-Cabot & Bažant, 1987),
Z  
nl 1 kx  nk
"^ ðxÞ ¼ 0 ^
"ðnÞdn ð67Þ
Vr ðxÞ V lc

with the normalizing factor


Z  
kx  nk
Vr ðxÞ ¼ 0 dn ð68Þ
V lc

^ is the local value of


with the integrals performed over the whole structure V. The quantity "(m)
equivalent strain at points m around position x and km  xk is the geometric distance. Two
1 2
frequently used weight functions are the Gaussian function 0 ðÞ ¼ e2 and the bell-shaped
2 2
polynomial function a0(z) ¼ h1  z i (Bažant and Jirásek, 2002).

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 845
(2) Such an approach considers a constant (isotropic) internal length lc whatever the damage level
and anisotropy.
(3) The use of gradient enhanced modeling (Aifantis, 1987; Peerlings et al., 1996), defining "^nl from
the implicit differential equation

"^nl  l2c r2 "^nl ¼ "^ ð69Þ

introducing an isotropic and constant internal length lc with the same remark on isotropy and
constancy of internal length than for standard nonlocal integral.
(4) The use of an internal time c as material parameter – instead of characteristic length lc – in
nonlocal integral averaging (Desmorat and Gatuingt, 2007, 2010)
Z   Z  
1 x x
"^nl ðxÞ ¼ 0 ^
"ðnÞdn Vr ðxÞ ¼ 0 dn ð70Þ
Vr ðxÞ V c V c

with xx the information time for an elastic wave to propagate from points x to m. This leads to
an evolving internal length, as suggested by Geers et al. (1998), Pijaudier-Cabot et al. (2004) and
Simone et al. (2004), but here damage is induced and anisotropic. This approach makes a highly
damaged zone equivalent to a crack. The main drawback of the nonlocal integral with internal
time framework is its prohibitive numerical cost in structural cases.
(5) The use of eikonal nonlocal integral framework (Desmorat et al., 2015)
Z ! Z !
nl 1 ‘~x ‘~x
"^ ðxÞ ¼ 0 ^
"ðnÞdn Vr ðxÞ ¼ 0 dn ð71Þ
Vr ðxÞ V lc V lc

where effective distances ‘~ ¼ ‘~x between points x and m are solution of anisotropic eikonal
equation

r‘~  H2  r‘~ ¼ r‘~  ð1  DÞ  r‘~ ¼ 1 ð72Þ

An anisotropic evolving internal length is obtained, dependent on both the damage level and
anisotropy. A highly damaged zone is found equivalent to a crack. In one dimension the eikonal
nonlocal integral framework is equivalent to the nonlocal framework with internal time theory.
(6) The use of eikonal gradient enhancement (Desmorat et al., 2015)

1 l2c 
"^nl  r  ðdet HÞ H2  r"^nl ¼ "^ ð73Þ
2 det H

Such an approach behaves as Geers et al. (1998) gradient enhancement but with an anisotropic
evolving internal length, dependent on both the damage level and anisotropy. For an uniaxial
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
state of damage H ¼ diag½H1 , 1, 1 ¼ diag½1= 1  D1 , 1, 1, setting x the normal to nucleated
crack, equation (73) gives
 
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffi @ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi @"^nl
"^nl  l2c 1  D1 1  D1 ¼ "^ ð74Þ
2 @x @x

Downloaded from ijd.sagepub.com by guest on July 26, 2016


846 International Journal of Damage Mechanics 25(6)

Both eikonal nonlocal integral and eikonal gradient enhancement end up as:
. a local behavior normally to the nucleated crack surface at D1 ! 1; and
. a nonlocal behavior parallel to the nucleated crack, as sought by Pijaudier-Cabot et al. (2007).

One does not enter here into the debate of the nonlocal integral versus the gradient enhancement,
both choices being often possible. It worth recalling the well-known (non-physical) feature of a
highly damaged spreading zone in nonlocal averaging. An evolving internal length (such as from a
eikonal nonlocal integral, see Desmorat et al. (2015) work) allows for the present anisotropic
damage modeling to localize the nucleated crack at a single point, where H1 ! 0, D1 ! 1.

Conclusion
An anisotropic damage model has been proposed for concrete materials, using the single Ladevèze
1
variable H ¼ ð1  DÞ2 for both coupling of deviatoric and hydrostatic elastic properties with
damage. The shear-bulk coupling obtained at positive stress triaxiality is found in accordance
with the numerical discrete element results when bulk modulus is not affected by damage in com-
pressive stress states. The model biaxial responses have been derived in a closed form parametric
representation in principal stresses space and advantageously compared to initial anisotropic
damage model responses. A key feature of the modeling is the proposed evolution law H _ ¼ h~
_ iþ
of damage anisotropy governed by the extensions, keeping second-order damage tensor H
unbounded and allowing for the sought property of gradual stress softening. The Lemaitre stress
triaxiality function has been introduced to enhance the performance of Mazars criterion and there-
fore the performance of the full anisotropic damage model in bicompression. The related thermo-
dynamics aspects have been discussed and both the proof of convexity of thermodynamics potential
with respect to stress p and the damage H, as well as the proof of the positivity of the intrinsic
dissipation, have been given.
The anisotropic damage models have been presented first in their local form. Different nonlocal
enhancements have finally been described; they simply replace the expression of local equivalent
strain "^ within Mazars criterion by its nonlocal averaging "^nl , either integral or of gradient type, with
a possibly evolving internal length, made anisotropic because of the tensorial nature of damage.
The modeling of permanent strains and of dilatancy in confined stress states is left for future
work. Suggested improvements in the considered anisotropic damage framework, for instance not
using plasticity theory, can be found in Desmorat (2004), Herrmann and Kestin (1988), La Borderie
(1991), La Borderie (2003) and Lebon (2011). The extension to cyclic loading is straightforward
from the concept of active damage (Desmorat et al., 2010b; Souid et al., 2009).

Declaration of Conflicting Interests


The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or pub-
lication of this article.

Funding
The author(s) received no financial support for the research, authorship, and/or publication of this article.

References
Aifantis E (1987) The physics of plastic deformation. International Journal of Plasticity 3: 211–247.

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 847
Allix O (2013) The bounded rate concept: A framework to deal with objective failure predictions in dynamic
within a local constitutive model. International Journal of Damage Mechanics 22: 808–828.
Allix O and Deü JF (1997) Delay-damage modelling for fracture prediction of laminated composites under
dynamic loading. Engineering Transactions 45: 29–46.
Badel P, Godard V and Leblond JB (2007) Application of some anisotropic damage model to the prediction of
the failure of some complex industrial concrete structure. International Journal of Solids and Structures 44:
5848–5874.
Bazant ZP and Gambarova PG (1984) Crack shear in concrete: Crack band microplane model. Journal of
Structural Engineering 110: 2015–2036.
Bažant ZP and Jirásek M (2002) Nonlocal integral formulations of plasticity and damage: Survey of progress.
Journal of Engineering Mechanics 128: 1119–1149.
Bažant ZP and Ozbolt J (1990) Nonlocal microplane model for fracture, damage and size effects in structures.
Journal of Engineering Mechanics 116: 2485–2505.
Bažant ZP and Prat PC (1988a) Microplane model for brittle-plastic material: I. Theory. Journal of Engineering
Mechanics 114: 1672–1688.
Bažant ZP and Prat PC (1988b) Microplane model for brittle-plastic material: II. Verification. Journal of
Engineering Mechanics 114: 1689–1702.
Billardon R and Pétry C (2005) Creep damage behaviour of a copper alloy on a large temperature range. In:
ASME/ASCE/SES conference on mechanics and materials (McMat2005), 1–3 June, Baton Rouge,
Louisiana, USA.
Brünig M (2003) An anisotropic ductile damage model based on irreversible thermodynamics. International
Journal of Plasticity 19: 1679–1714.
Burlion N, Gatuingt F, Pijaudier-Cabot G, et al. (2000) Compaction and tensile damage in concrete:
Constitutive modelling and application to dynamics. Computer Methods in Applied Mechanics and
Engineering 183: 291–308.
Carol I, Rizzi E and Willam K (2001) On the formulation of anisotropic elastic degradation. Part I: Theory
based on a pseudo-logarithmic damage tensor rate. Part II: Generalized pseudo-Rankine model for tensile
damage. International Journal of Solids and Structures 38(4): 491–518 (519–546).
Chaboche JL (1982) Le concept de contrainte effective appliqué à l’élasticité et à la viscoplasticité en présence
d’un endommagement anisotrope. Comportment Me´chanique des Solides Anisotropes. Dordrecht: Springer,
pp. 737–760.
Chaboche JL (1984) Anisotropic creep damage in the framework of continuum damage mechanics. Nuclear
Engineering and Design 79(3): 309–319.
Chaboche JL (1993) Development of continuum damage mechanics for elastic solids sustaining anisotropic and
unilateral damage. International Journal of Damage Mechanics 2: 311–329.
Chaboche J and Maire J (2000) New progress in micromechanics-based CDM models and their application to
CMCs. Composites Science and Technology 61(15): 2239–2246.
Chambart M (2009) Endommagement anisotrope et comportement dynamique des structures en béton armé
jusqu’à ruine. PhD ENS-Cachan, France.
Chambart M, Desmorat R and Gatuingt F (2014) Intrinsic dissipation of a modular anisotropic damage model:
Application to concrete under impact. Engineering Fracture Mechanics 127: 161–180.
Chen W, Maurel O, Reess T, et al. (2011) Modelling anisotropic damage and permeability of mortar under
dynamic loads. European Journal of Environmental and Civil Engineering 15: 727–742.
Cordebois J and Sidoroff F (1982) Anisotropic damage in elasticity and plasticity. Journal de Me´canique
The´orique et Applique´e, nume´ro spécial 45–60.
Cundall PA and Strack ODL (1979) A discrete numerical model for granular assemblies. Geotechnique 29:
47–65.
Delaplace A and Desmorat R (2007) Discrete 3D model as complimentary numerical testing for anisotropic
damage. International Journal of Fracture 148: 115–128.
Desmorat R (2000) Quasi-unilateral conditions in anisotropic elasticity. Comptes Rendus de l’Acade´mie des
Sciences 328(6): 445–450.

Downloaded from ijd.sagepub.com by guest on July 26, 2016


848 International Journal of Damage Mechanics 25(6)

Desmorat R (2004) Modèle d’endommagement anisotrope avec forte dissymétrie traction/compression. In: 5e`
journe´es du Regroupement Francophone pour la Recherche et la Formation sur le Be´ton (RF2B), 5–6 July,
Liège, Belgium.
Desmorat R (2006) Positivity of intrinsic dissipation of a class of nonstandard anisotropic damage models.
Comptes Rendus de l’Acade´mie des Sciences 334(10): 587–592.
Desmorat R (2012) Continuum approach in damage mechanics. In: Summer school hardening and damage of
materials under finite deformations, constitutive modeling and numerical implementation. IBZ TU Dortmund,
Germany, 3–7 September 2012. Available at: http://www.lem3.fr/summerschool/00-Files/RD.pdf.
Desmorat R and Cantournet S (2008) Modeling micro-defects closure effect with isotropic/anisotropic damage.
International Journal of Damage Mechanics 17: 65–96.
Desmorat R and Gatuingt F (2007) Introduction of an internal time in nonlocal integral theories. Internal report
LMT 268, LMT-Cachan, France, 2007. Available at: http://hal.archives-ouvertes.fr/hal-00200898/en/.
Desmorat R and Gatuingt F (2010) Introduction of an internal time in nonlocal integral theories. In: Bicanic N,
de Borst R, Mang HA, et al. (eds) Computational Modelling of Concrete Structures – EURO-C 2010.
London: Taylor & Francis, pp. 121–128.
Desmorat R and Otin S (2008) Cross-identification isotropic/anisotropic damage and application to anisother-
mal structural failure. Engineering Fracture Mechanics 75(11): 3446–3463.
Desmorat R, Gatuingt and Ragueneau F (2007) Nonlocal anisotropic damage model and related computa-
tional aspects for quasi-brittle materials. Engineering Fracture Mechanics 74(10): 1539–1560.
Desmorat R, Gatuingt F and Ragueneau F (2010a) Non standard thermodynamics framework for robust
computations with induced anisotropic damage. International Journal of Damage Mechanics 19(1): 53–73.
Desmorat R, Chambart M, Gatuingt F, et al. (2010b) Delay-active damage versus non-local enhancement for
anisotropic damage dynamics computations with alternated loading. Engineering Fracture Mechanics 77:
2294–2315.
Desmorat R, Gatuingt F and Jirasek M (2015) Nonlocal models with damage-dependent interactions moti-
vated by internal time. Preprint Engineering Fracture Mechanics 142: 255–275.
Goidescu C, Welemane H, Pantale O, et al. (2015) Anisotropic unilateral damage with initial orthotropy:
A micromechanics-based approach. International Journal of Damage Mechanics 24(3): 313–337.
Fichant S, Pijaudier-Cabot G and Laborderie C (1997) Continuum damage modelling: approximation of crack
induced anisotropy. Mechanics Research Communications 24: 109–114.
Gatuingt F and Pijaudier-Cabot G (2002) Coupled damage and plasticity modelling in transient dynamic
analysis of concrete. International Journal for Numerical and Analytical Methods in Geomechanics 26: 1–24.
Gatuingt F, Desmorat R, Chambart M, et al. (2008) Anisotropic 3D delay-damage model to simulate concrete
structures. Revue Europe´enne de Me´canique Nume´rique 17: 749–760.
Geers M, de Borst R, Brekelmans W, et al. (1998) Strain-based transient-gradient damage model for failure
analyses. Computer Methods in Applied Mechanics and Engineering 160: 133–153 (1998).
Govindjee S, Kay GJ and Simo JC (1995) Anisotropic modelling and numerical simulation of brittle damage in
concrete. International Journal for Numerical Methods in Engineering 38(21): 3611–3633.
Grassl P and Jirasek M (2006) Damage-plastic model for concrete failure. International Journal of Solids and
Structures 43: 7166–7196.
Halm D and Dragon A (1998) An anisotropic model of damage and frictional sliding for brittle materials.
European Journal of Mechanics – A/Solids 17: 439–460.
Halphen B and Nguyen QS (1975) Sur les matériaux standards généralisés. Journal de MÉcanique 14: 39–63.
Herrmann G and Kestin J (1988) On thermodynamics foundations of damage theory in elastic solids.
In: Mazars J and Bazant ZP (eds) Strain Localization and Size Effects due to damage and Cracking.
Cachan, France: CNRS-NSF, pp. 228–232.
Herrmann HJ and Roux S (1990) Statistical Models for the Fracture of Disordered Media. Elsevier: Amsterdam.
Jirasek M (1999) Comments on microplane theory. In: Pijaudier-Cabot G, Bittnar Z and Gérard B (eds)
Mechanics of Quasi-Brittle Materials and Structures. Paris: Hermès Science Publications, pp. 57–77.
Ju JW (1989) On energy-based coupled elastoplastic damage theories: Constitutive modeling and computa-
tional aspects. International Journal of Solids and Structures 25(7): 803–833.

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 849
Kattan P and Voyiadjis G (1990) A coupled theory of damage mechanics and finite strain elasto-plasticity – I.
Damage and elastic deformations. International Journal of Engineering Science 28: 421–435.
Kuhl E and Ramm E (1998) On the linearization of the microplane model. Mechanics of Cohesive-frictional
Materials 3: 343–364.
Kupfer H, Hilsdorf HK and Rusch H (1969) Behaviour of concrete under biaxial stresses. ACI Journal
Proceedings 66: 656–666.
La Borderie C (1991) Phe´nomènes unilate´raux dans un mate´riau endommageable: Mode´lisation et application à
l’Analyse de structures en be´ton. PhD Theis, Université Paris, France.
La Borderie C (2003) Strate´gies et Mode`les de Calculs pour les Structures en Be´ton, Habilitation Diriger des
Recherches. Université de Pau et des Pays de l’Adour, France.
Ladevèze P (1983) On an anisotropic damage theory. In: Boehler JP (ed.) Proceedings of the CNRS
International College on Failure Criteria of Structured Media. Vol. 351, France: Villars de Lens, pp. 355–363.
Ladevèze P and Lemaitre J (1984) Damage effective stress in quasi unilateral conditions. In: Niordson FI and
Olhoff N (eds) 16th International congress of theoretical and applied mechanics, Lyngby, Denmark, 19–25
August 1984. Amsterdam: Elsevier Science Publishers.
Ladevèze P, Gasser A and Allix O (1994) Damage mechanisms modelling for ceramic composites. Journal of
Engineering Materials and Technology 116: 331–336.
Ladevèze P (1995) Modeling and simulation of the mechanical behavior of CMCs. High-Temperature Ceramic-
Matrix Composites 47: 53–63.
Lebon G (2011) Analyse de l’endommagement des structures de ge´nie civil: Techniques de sous-structuration
hybride couple´es à un mode`le d’endommagement anisotrope. PhD Theis, ENS-Cachan, France.
Leckie F and Onat E (1981) Tensorial nature of damage measuring internal variables. In: Hult J and Lemaire J
(eds) Physical Non-linearities in Structural Analysis. Berlin: Springer, pp. 140–155.
Lemaitre J (1992) A Course on Damage Mechanics. Berlin: Springer.
Lemaitre J and Chaboche JL (1985) Me´canique des Mate´riaux Solides, Paris: Dunod. [Mechanics of Solid
Materials, Oxford: Oxford University Press, 1991 (english translation)].
Lemaitre J and Chaboche JL (2009) Me´canique des Mate´riaux Solides, 3rd ed. Paris: Dunod.
Lemaitre J and Sauzay M (1999) Evolution law for anisotropic damage. Comptes Rendus de l’Academie des
Sciences Series IIB Mechanics Physics Astronomy 327: 1231–1236.
Lemaitre J, Desmorat R and Sauzay M (2000) Anisotropic damage law of evolution. European Journal of
Mechanics - A/Solids 19: 187–208.
Lemaitre J and Desmorat R (2005) Engineering Damage Mechanics: Ductile, Creep, Fatigue and Brittle Failures.
Berlin: Springer.
Leroux A (2012) Mode`le multiaxial d’endommagement anisotrope: Gestion nume´rique de la rupture et application
à la ruine de structures en be´ton arme´ sous impacts. PhD Thesis, ENS-Cachan, France.
Maire J and Chaboche J (1997) A new formulation of Continuum Damage Mechanics (CDM) for composite
materials. Aerospace Science and Technology 1(4): 247–257.
Mazars J (1984) Application de la me`canique de l’endommagement au com- portement non line`aire et à la rupture
du be`ton de structure. Phd Thesis, Universitè Paris 6.
Mazars J (1986) A description of micro and micro scale damage of concrete structure. Journal of Engineering
Fracture of Mechanics 25(5–6): 729–737.
Menzel A and Steinmann P (2001) A theoretical and computational setting for anisotropic continuum damage
mechanics at large strains. International Journal of Solids and Structures 38: 9505–9523.
Menzel A, Ekh M, Steinmann P, et al. (2002) Anisotropic damage coupled to plasticity: Modelling based on the
effective configuration concept. International Journal for Numerical Methods in Engineering 54: 1409–1430.
Meschke G, Lackner R and Mang A (1998) An anisotropic elastoplastic-damage model for plain concrete.
International Journal for Numerical Methods in Engineering 42: 703–727.
Murakami S (1988) Mechanical modeling of material damage. Journal of Applied Mechanics 55: 280–286.
Papa E and Taliercio A (1996) Anisotropic damage model for the multiaxial static and fatigue behaviour of
plain concrete. Engineering Fracture Mechanics 55: 163–179.

Downloaded from ijd.sagepub.com by guest on July 26, 2016


850 International Journal of Damage Mechanics 25(6)

Peerlings R, de Borst R, Brekelmans W, et al. (1996) Gradient-enhanced damage model for quasi-brittle
materials. International Journal for Numerical Methods in Engineering 39: 391–403.
Pijaudier-Cabot G and Bažant Z (1987) Nonlocal damage theory. Journal of Engineering Mechanics 113:
1512–1533.
Pijaudier-Cabot G, Haidar K and Dubé J-F (2004) Nonlocal damage model with evolving internal length.
International Journal for Numerical and Analytical Methods in Geomechanics 28: 633–652.
Pijaudier-Cabot G, Krayani A and Dufour F (2007) Comments on boundary effects in non local damage based
models. In: Yuan H and Wittmann FH (eds) Nonlocal Modeling of Materials Failure. Aedificio Pubs.
Pontiroli C (1995)) Comportement au souffle des structures en be´ton arme´. PhD Thesis, ENS-Cachan, France.
Pröchtel P and Häußler-Combe U (2009) On the dissipative zone in anisotropic damage models for concrete.
International Journal of Solids and Structures 45(16): 4384–4406.
Ragueneau F, La Borderie C and Mazars J (2000) Damage model for concrete-like materials coupling cracking
and friction, contribution towards structural damping: First uniaxial applications. Mechanics of Cohesive-
Frictional Materials 5: 607–626.
Ragueneau F, Desmorat R and Gatuingt F (2008) Anisotropic damage modelling of biaxial behaviour and
ruprure of concrete structures. Computers and Concrete 5(4): 417–434.
Ramtani S, Berthaud Y and Mazars J (1992) Orthotropic behaviour of concrete with directional aspects:
modelling and experiments. Nuclear Engineering and Design 133: 97–111.
Richard B and Ragueneau F (2013) Continuum damage mechanics based model for quasi brittle materials
subjected to cyclic loadings: Formulation, numerical implementation and applications. Engineering Fracture
Mechanics 98: 383–406.
Schlangen E and Garboczi EJ (1997) Fracture simulations of concrete using lattice models: Computational
aspects. Engineering Fracture Mechanics 57(2–3): 319–332.
Simone A, Wells GN and Sluys LJ (2004) From continuous to discontinuous failure in a gradient enhanced
continuum damage model. Computer Methods in Applied Mechanics and Engineering 192(41–42): 4581–4607.
Souid A, Ragueneau F, Delaplace A, et al. (2009) Pseudodynamic testing and nonlinear substructuring of
damaging structures under earthquake loading. Engineering Structures 31: 1102–1110.
Steinmann P and Carol I (1998) A framework for geometrically nonlinear continuum damage mechanics.
International Journal of Engineering Science 36: 1793–1814.
Van Mier JGM, Van Vliet MRA and Wang TK (2002) Fracture mechanisms in particle composites: Statistical
aspects in lattice type analysis. Mechanics Materials 34: 705–724.
Willam K, Pramono E and Sture S (1987) Fundamental issues of smeared crack models. In: Shah |SP and
Schwartz |SE (eds) Proceedings of the SEM-RILEM International Conference. On Fracture of Concrete and
Rock. Society of Engineering Mechanics, 193–207. New York: Springer, pp. 142–157.
Yin WS, Su ECM, Mansur MA, et al. (1989) Biaxial Tests of Plain and Fiber Concrete. ACI Materials Journal
86(3): 236–243.

Appendix 1: Functions Bi

Bi ¼ B0i þ BH ¼ Bi ð , Di Þ
 
1þ 4 1 1
B01 ¼ þ þ cos
9 1  D1 1  D2 1  D3
 
1þ 2 2 1
 þ  sin
9 1  D1 1  D2 1  D3

Downloaded from ijd.sagepub.com by guest on July 26, 2016


Desmorat 851
 
1þ 2 2 1
B02 ¼  þ  cos
9 1  D1 1  D2 1  D3
 
1þ 1 4 1
þ þ þ sin
9 1  D1 1  D2 1  D3
 
1þ 2 1 2
B03 ¼   þ cos
9 1  D1 1  D2 1  D3
 
1þ 1 2 2
þ   sin
9 1  D1 1  D2 1  D3

1  2 1  2
BH ¼ hcos þ sin i  h cos  sin i
3  ðD1 þ D2 þ D3 Þ 3

Appendix 2: Triaxiality functions Rl


The strain energy density of virgin (undamaged) material is

2
1 eq R
we ¼  : E :  ¼
2 2E

where the concept of triaxiality function Rn has been introduced (Lemaitre, 1992), normed to one in
tension,

2 H 1 tr r
R ¼ ð1 þ Þ þ 3ð1  2ÞT2X TX ¼ ¼
3 eq 3 ð3 r0 : r0 Þ12
2

In the present work one uses a similar definition but for negative triaxiality only and normed to one
in pure shear (Rn ¼ 1 at TX  0)

9 1  2
R ¼ 1 þ hTX i2
2 1þ

and one bounds it to B.

Appendix 3: Functions Ci

Ci ¼ C0i þ CH ¼ Ci ð , Hi Þ

1þ  1þ 
C01 ¼ 4H21 þ H22 þ H23 cos  2H21 þ 2H22  H23 sin
9 9

Downloaded from ijd.sagepub.com by guest on July 26, 2016


852 International Journal of Damage Mechanics 25(6)

1þ  1þ 2 
C02 ¼  2H21 þ 2H22  H23 cos þ H1 þ 4H22 þ H23 sin
9 9
1þ  1þ 2 
C03 ¼  2H21  H22 þ 2H23 cos þ H1  2H22  2H23 sin
9 9
1  2  2  1  2
CH ¼ H1 þ H2 2
2 þ H3 hcos þ sin i  h cos  sin i
9 3

Appendix 4: Positive scalar product of two positive symmetric tensors


_ Spectral decomposition of A gives
Consider two positive symmetric second-order tensors A and H.
X
A¼ Ai ai  ai Ai  0 ð75Þ
i

with eigenvalues Ai and eigenvectors ai. The scalar product


X
_ ¼
A:H _  ai  0
Ai ai  H ð76Þ
i

is positive, as a sum of positive terms.

Downloaded from ijd.sagepub.com by guest on July 26, 2016


View publication stats

You might also like