Sunteți pe pagina 1din 29

Available online at www.sciencedirect.

com

International Journal of Plasticity 24 (2008) 2192–2220


www.elsevier.com/locate/ijplas

Three dimensional combined


fracture–plastic material model for concrete
Jan Červenka a, Vassilis K. Papanikolaou b,*
a
Červenka Consulting, Predvoje 22, 16200 Prague 6, Czech Republic
b
Laboratory of Reinforced Concrete and Masonry Structures, Civil Engineering Department,
Aristotle University of Thessaloniki, P.O. Box 482, Thessaloniki, 54124, Greece

Received 30 July 2007; received in final revised form 23 January 2008


Available online 5 February 2008

Abstract

This paper describes a combined fracture–plastic model for concrete. Tension is handled by a
fracture model, based on the classical orthotropic smeared crack formulation and the crack band
approach. It employs the Rankine failure criterion, exponential softening, and it can be used as a
rotated or a fixed crack model. The plasticity model for concrete in compression is based on the
Menétrey–Willam failure surface, the plastic volumetric strain as a hardening/softening parameter
and a non-associated flow rule based on a nonlinear plastic potential function. Both models use a
return-mapping algorithm for the integration of constitutive equations. Special attention is given
to the development of an algorithm for the combination of the two models. The suggested combi-
nation algorithm is based on a recursive substitution, and it allows for the two models to be devel-
oped and formulated separately. The algorithm can handle cases when failure surfaces of both
models are active, but also when physical changes such as crack closure occur. The model can be
used to simulate concrete cracking, crushing under high confinement and crack closure due to crush-
ing in other material directions. The model is integrated in a general finite element package ATENA
and its performance is evaluated by comparisons with various experimental results from the
literature.
Ó 2008 Elsevier Ltd. All rights reserved.

Keywords: A. Fracture; B. Concrete; B. Constitutive behavior; B. Elastic–plastic material; C. Finite elements

*
Corresponding author. Tel.: +30 2310995662; fax: +30 2310995614.
E-mail address: billy@civil.auth.gr (V.K. Papanikolaou).

0749-6419/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijplas.2008.01.004
J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2193

Nomenclature

a attraction parameter of the plastic potential function


A first coefficient of the plastic potential function
B second coefficient of the plastic potential function
C third coefficient of the plastic potential function
c(j) softening function
dk plastic and fracture multiplier
D, Dijkl elasticity matrix
Ds secant constitutive matrix
0
D cr cracked stiffness matrix
e eccentricity parameter of the Menétrey–Willam failure surface
ep plastic deviatoric strain
Ec concrete elastic modulus
ff rankine failure surface (fracture)
fp Menétrey–Willam failure surface (plasticity)
fc uniaxial compressive concrete strength
ft uniaxial tensile concrete strength
g plastic potential function
Gf fracture energy
H hardening modulus
k(j) hardening function
ko hardening parameter defining the onset of plastic flow
lij stress return direction
Lt characteristic length
m friction parameter of the Menétrey–Willam failure surface
n order of the plastic potential function
nk eigenvector defining direction k
n1 first parameter of softening function
n2 second parameter of softening function
r(h, e) elliptic function of the Menétrey–Willam failure surface
rg crack shear stiffness coefficient
t slope parameter of the softening function
T transformation matrix
w crack opening
wo crack opening corresponding to zero tensile stress
af fracture contribution in the combination algorithm
ap plasticity contribution in the combination algorithm
b relaxation factor in the combination algorithm
e, eij total strain vector and tensor
ee, eeij elastic strain vector and tensor
ep, epij plastic strain vector and tensor
ef, efij fracture strain vector and tensor
^efk maximum fracture strain in direction k
2194 J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220

epv plastic volumetric strain


epv;t plastic volumetric strain at uniaxial concrete strength (threshold value)
h lode angle of the stress vector in Haigh–Westergaard stress space
j hardening/softening parameter
kt scaling factor for the tensile strength ft of the plasticity model
m concrete Poisson’s ratio
n hydrostatic length of the stress vector in Haigh–Westergaard stress space
n0 hydrostatic length of the plastic strain vector
q deviatoric length of the stress vector in Haigh–Westergaard stress space
q0 deviatoric length of the plastic strain vector
r, rij stress vector and tensor
rco concrete stress defining the onset of plastic flow
v convergence rate of the combination algorithm
w inclination of the plastic strain vector

1. Introduction

This paper describes a three dimensional constitutive material model for concrete,
which combines plasticity with fracture. Fracture is modeled by an orthotropic smeared
crack model based on the Rankine tensile criterion. A hardening/softening plasticity
model based on the Menétrey and Willam (1995) three-parameter failure surface is used
to simulate concrete crushing. Although many papers have been published on plasticity
models for concrete (e.g. Pramono and Willam, 1989; Etse, 1992; Feenstra, 1993; Menét-
rey et al., 1997; Feenstra et al., 1998; Grassl et al., 2002) or smeared crack models (e.g.
Rashid, 1968; Červenka and Gerstle, 1971; Bažant and Oh, 1983; De Borst, 1986; Rots
and Blaauwendraad, 1989), there are not many descriptions of their successful combina-
tion in the literature. Owen et al. (1983) presented a combination of cracking and visco-
plasticity. Comprehensive treatise of the problem was provided also by De Borst (1986),
and several works have been published on the combination of damage and plasticity
(e.g. Simo and Ju, 1987; Meschke et al., 1988; Bielger and Mehrabadi, 1995; Lee and Fen-
ves, 1998; or more recent works by Grassl and Jirásek (2006), Mohamad-Hussein and
Shao (2007), Contrafatto and Cuomo (2006), Jason et al. (2006), Cicekli et al. (2007)
and Chiarelli et al. (2003) for rock material). Various concrete models have been proposed
in the literature that are based on different approaches such as for instance the class of
microplane models (Bažant et al., 2000) or models based on concrete micromechanics
(Mattei et al., 2007), which are derived using the theory of granular materials (Christoffer-
sen et al., 1981; Oda et al., 1982; Mehrabadi et al., 1982).
The plastic-damage models are usually formulated within the concept of thermodynam-
ics, and with the exception of the works by Meschke et al. (1988) and Cicekli et al. (2007),
they usually consider an isotropic damage formulation, which neglects the anisotropic nat-
ure of cracked concrete behavior. In the presented model, the cracked concrete is modeled
as an orthotropic material and it considers the problematic of physical changes, like for
instance crack closure as it is the case of the model proposed by Cicekli et al. (2007). In
addition, it considers the shear behavior of cracked concrete and rotated as well as fixed
crack formulation. Also within the proposed approach it is possible to formulate both
J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2195

models (i.e. plastic and fracture) entirely separately, and their combination can be pro-
vided in a separate algorithm. From a programming point of view, such an approach is
well suited for object oriented programming.
The suggested model is an improved version of a previously published model by Čer-
venka et al. (1998) that includes various enhancements in both fracturing and plasticity
parts. The fracturing part of the model have been extended to address the mode II and
III (i.e. shear) crack propagation. The plasticity part is extended to include the formu-
lations suggested by Papanikolaou and Kappos (2007), which describe the increased
strength and deformation capacity of concrete under multiaxial compression and can
properly handle problems including confinement effects. The major difference between
the present model and the one suggested by the above writers is that the former (using
a special algorithm that combines plasticity with fracture) can efficiently handle all pos-
sible loading and unloading paths (tension, compression or combinations of both) and
describe crack opening and closure as well. Another advantage of the present model
is that it was successfully integrated in a general finite element package (ATENA,
Červenka et al., 2007) and can be directly applied to the analysis of complex reinforced
concrete structures, rather than simulating only multiaxial loading of plain concrete
(using a constitutive driver).
The approach of strain decomposition introduced by De Borst (1986) is used to com-
bine fracture and plasticity models together. Both models are developed within the frame-
work of the return-mapping algorithm (Wilkins, 1964). This approach guarantees the
solution for all magnitudes of strain increment. From an algorithmic point of view, the
problem is then transformed into finding an optimal return point on the active failure sur-
faces. The combined algorithm must determine the separation of strains into plastic and
fracturing components, while it must preserve stress equivalence in both models. The pro-
posed algorithm is based on a recursive iterative scheme. It can be shown that such a recur-
sive algorithm cannot reach convergence in certain cases such as, for instance, softening
and dilating materials. For this reason the recursive algorithm is extended by a relaxation
method, in order to stabilize convergence.
In the first part of the paper, the constitutive equations of the fracture and plastic model
are presented. This part also contains a description of the recursive algorithm for the com-
bination of the two material models. In the subsequent section, the numerical behavior of
this algorithm is demonstrated under several selected loading histories, along with com-
parisons with experimental results. The last section demonstrates the model performance
in practical engineering applications.

2. Material model formulation

The material model formulation assumes small strains and is based on the strain
decomposition into elastic ðeeij Þ, plastic ðepij Þ and fracture ðefij Þ components (De Borst,
1986), which can be written in the rate form as
e_ ij ¼ e_ eij þ e_ pij þ e_ fij ð1Þ
The stress development can be then described by the following rate equations describing
the progressive degradation (concrete cracking) and plastic yielding (concrete crushing):
r_ ij ¼ Dijkl  ð_ekl  e_ pkl  e_ fkl Þ ð2Þ
2196 J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220

where the fracture strain rate ð_efij Þ and plastic strain rate ð_epij Þ are evaluated from the frac-
ture and plasticity models respectively. The constitutive equations of the both models can
be summarized as follows:
Flow rule governs the evolution of plastic and fracturing strains:
ogp
Plastic model : e_ pij ¼ k_ p  mpij ; mpij ¼ ð3Þ
orij
ogf
Fracture model : e_ fij ¼ k_ f  mfij ; mfij ¼ ð4Þ
orij
where k_ p is the plastic multiplier rate and gp is the plastic potential function. Following the
unified theory of elastic degradation of Carol et al. (1994) it is possible to define analogous
quantities for the fracturing model, i.e. k_ f is the inelastic fracturing multiplier, respectively,
and gf is the potential defining the direction of inelastic fracturing strains in the fracturing
model. The consistency conditions can be than used to evaluate the change of the plastic
and fracturing multipliers.
of p
f_ p ¼ npij  r_ ij þ H p  k_ p ¼ 0; npij ¼ ð5Þ
orij
of f
f_ f ¼ nfij  r_ ij þ H f  k_ f ¼ 0; nfij ¼ ð6Þ
orij
This represents and system of two equations for the two unknown multiplier rates k_ p
and k_ f , and is analogous to the problem of multi-surface plasticity (Simo et al., 1988).
The combination of the two models is described in detail in Section 2.3.

2.1. Fracture model for concrete cracking

The Rankine criterion is used for concrete cracking. For each direction (k = 1, 2, 3), it is
expressed as
fkf ¼ t rij  nki  nkj  ft 6 0 ð7Þ
or in Haigh–Westergaard coordinates (Fig. 1):

Fig. 1. Rankine failure surface represented in Haigh–Westergaard coordinates: (a) 3D stress space, (b) Rendulic
plane and (c) deviatoric plane.
J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2197
pffiffiffi pffiffiffi
ff ¼ n  2  q  cos h  3  ft 6 0 ð8Þ
The Rankine criterion (7) represents actually three distinct planes forming a pyramid in
stress space, as depicted in Fig. 1. It is assumed that strains and stresses are converted into
material directions given by the eigenvectors nk, which in the case of rotated crack model
correspond to the instantaneous principal strain directions, and in the case of fixed crack
model are given by the principal strain directions at the onset of cracking. Therefore, trij
identifies with the trial stress and ft with the concrete tensile strength. The trial stress state
is computed by the elastic predictor:
t
rij ¼ n rij þ Dijkl  dekl ð9Þ
t
If the trial stress ( rij) does not satisfy the Rankine failure criterion (Eq. (7)), the incre-
ment of fracture strain in direction k can be calculated using the assumption that the final
stress state must satisfy the following equation:
fkf ¼ nþ1 rij  nki  nkj  ft ¼ t rij  nki  nkj  Dijmn  defmn  nki  nkj  ft ¼ 0 ð10Þ
This equation can be further simplified under the assumption that the increment of frac-
turing strain is normal to the Rankine failure surface, and that always only one direction is
being checked each time. For failure surface k, the fracturing strain increment has the fol-
lowing form (associated form, i.e. gf = f f):
ofkf
de0fk k
ij ¼ dk  ¼ dkk  nki  nkj ð11Þ
orij
After substituting Eq. (11) in Eq. (10), a formula for the increment of the fracturing mul-
tiplier is recovered:
t
rij  nki  nkj  ft ðwk Þ
dkk ¼ ð12Þ
Dijmn  nki  nkj  nkm  nkn

where
wk ¼ Lt  ð^efk þ dkk Þ ð13Þ
The system of Eqs. (12) and (13) must be solved iteratively since for softening materials
the value of the current tensile strength ft (wk) is a function of the crack opening (wk) which
is based on the following empirical formula suggested by Hordijk (1991):
"  3 #
r w c  w w
¼ 1 þ c1   e 2 wo   ð1 þ c31 Þ  ec2 ð14Þ
ft wo wo

where: r is the tensile concrete stress normal to crack, ft is the concrete tensile strength,
c1 = 3, c2 = 6.93 and wo = 5.14  Gf/ft. (Gf is the fracture energy of the material, provided
as a model parameter, see Table 1.)
The crack opening (w) is computed from the total accumulated value of fracturing
strain ð^efk Þ in direction k, plus the current increment of fracturing strain (dk), and this
sum is multiplied by the characteristic length (Lt). The characteristic length as a crack
band size was introduced by Bažant and Oh (1983). Various methods were proposed
for the crack band size calculation in the framework of the finite element method. Feenstra
(1993) suggested a method based on integration point volume, which is not well suited for
2198
J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220
Table 1
Suggested parameters for fracture and plasticity models
fc (MPa) 20 30 40 50 60 70 80 90 100 110 120
Ec (MPa) 24377 27530 30011 32089 33893 35497 36948 38277 39506 40652 41727
m 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2
ft (MPa) 1.917 2.446 2.906 3.323 3.707 4.066 4.405 4.728 5.036 5.333 5.618
kt 1.043 1.227 1.376 1.505 1.619 1.722 1.816 1.904 1.986 2.063 2.136
e 0.5281 0.5232 0.5198 0.5172 0.5151 0.5133 0.5117 0.5104 0.5092 0.5081 0.5071
fco (MPa) 4.32 9.16 15.62 23.63 33.14 44.11 56.50 70.30 85.48 102.01 114.00
epv;t 4.92  104 6.54  104 8.00  104 9.35  104 1.06  103 1.18  103 1.30  103 1.41  103 1.52  103 1.62  103 1.73  103
t 1.33  103 2.00  103 2.67  103 3.33  103 4.00  103 4.67  103 5.33  103 6.00  103 6.67  103 7.33  103 8.00  103
A 7.342177 5.436344 4.371435 3.971437 3.674375 3.43856 3.245006 3.082129 2.942391 2.820644 2.713227
B 8.032485 6.563421 5.73549 5.430334 5.202794 5.021407 4.871993 4.745867 4.637358 4.542587 4.458782
C 3.726514 3.25626 3.055953 2.903173 2.797059 2.719067 2.659098 2.611426 2.572571 2.540158 2.512681
n 3 3 3 3 3 3 3 3 3 3 3
Gf (MN/m) 4.87  105 6.47  105 7.92  105 9.26  105 1.05  104 1.17  104 1.29  104 1.40  104 1.50  104 1.61  104 1.71  104
J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2199

distorted elements. A consistent and rather complex approach was proposed by Olivier
(1989). In the present model, the crack band size is calculated as a width or size of the ele-
ment projected into the direction k (Fig. 2). Červenka et al. (1995) showed that this
approach is satisfactory for low order linear finite elements, which are used throughout
this study. They also proposed a modification, which accounts for cracks that are not
aligned with element edges. The crack band approach assures that the energy dissipation
does not depend on the finite element size. Other methods have been proposed in the lit-
erature based on nonlocal averaging (Bažant and Pijaudier-Cabot, 1987) or gradient
approaches (De Borst and Mühlhaus, 1992), and a recent work by Hashiguchi and Tsuts-
umi (2007).
The system of Eqs. (12) and (13) can be solved by recursive substitutions. It is possible to
show by expanding ft(wk) into a Taylor series that this iteration scheme converges as long as
 
 oft ðwk Þ Dijmn  nki  nkj  nkm  nkn
 < ð15Þ
 ow  Lt
Eq. (15) is violated for softening materials only when snap back is observed in the
stress–strain relationship, which can occur if large finite elements are used. In the standard
displacement based finite element method, the strain increment is given and therefore, a
snap back on the constitutive level cannot be captured. This means that the critical region,
a snap back on the softening curve will be skipped in a real calculation, which physically
means that the energy dissipated by the system will be overestimated. This is of course
minðD  Þ
undesirable, and finite elements smaller than Lt < of ð0Þkkkk  should be used, where
 ow
t

 
oft ð0Þ
 
 ow  denotes the initial slope of the crack softening curve.
It is important to distinguish between the total fracturing strain ð^efk Þ, which corresponds
to the maximum fracturing strain at material direction k reached during the loading pro-
cess, and the current fracturing strain ðefij Þ, which can be smaller due to crack closure. The
local fracturing strains e0fij in the local coordinate system given by the material directions nk
can be calculated using the following equation derived by Rots and Blaauwendraad (1989)
in the matrix form:
1
e0f ¼ ðD þ D0cr Þ De ð16Þ

Fig. 2. Tensile strength function with respect to crack width (Hordijk, 1991).
2200 J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220

where D0cr
ijkl is the stiffness inside the crack zone defined as
0f
r0ij ¼ D0cr
ijkl  ekl ð17Þ
The fourth order crack tensor D0cr
represents the cracking stiffness in the local material
ijkl
directions. In the current formulation, it is assumed that there is no interaction between
normal and shear components. Thus, the nonzero crack tensor elements are given by
the following formulas: Mode I (normal) crack stiffness:
ft ðwk Þ
D0cr
kkkk ¼ ðno index summationÞ ð18Þ
^efk
Mode II and III (shear) crack stiffness:
D0cr 0cr 0cr
ijij ¼ minðDiiii ; Djjjj Þ  r
g
ðno index summation and i 6¼ jÞ ð19Þ
g
where r is the crack shear stiffness coefficient, which is an input parameter. It represents
the ratio of the shear crack stiffness to the normal crack stiffness. The recommended value
is within the range of 1–10. In Eqs. (18) and (19), it is necessary to handle the special cases
before the onset of cracking, when the expressions approach infinity. Large penalty num-
ft ð0Þ
bers are used for normal crack stiffness in these cases: D0cr
kkkk ¼ e , where e is a small num-
ber (also ^efk is initialized to e at the beginning).
The secant constitutive matrix in the material directions can be derived from Eq. (2)
using Eq. (16):
1
D0s ¼ D  D  ðD0cr þ DÞ D ð20Þ
Strain vector transformation matrix T (i.e. global to local strain transformation matrix)
can be used to transform the local secant stiffness matrix to the global coordinate system.
A detailed flow chart of the fracture model is shown in Fig. 3.
Ds ¼ TT  D0s  T ð21Þ

2.2. Plasticity model for concrete crushing

The new stress state in the plastic model is computed using the predictor-corrector
formula:
nþ1
rij ¼ n rij þ Dijkl  ðdekl  depkl Þ ¼ t rij  Dijkl  depkl ¼ t rij  rpij ð22Þ

The plastic corrector ðrpij Þ is computed directly from the yield function by the return-map-
ping algorithm:
f p ðt rij  rpij Þ ¼ f p ðt rij  dk  lij Þ ¼ 0 ð23Þ
The crucial aspect is the definition of the return direction (lij) which can be defined
as
ogðt rkl Þ
lij ¼ Dijkl  ð24Þ
orkl
where g(trij) is the plastic potential function, which derivative is evaluated at the predictor
stress state (trij) to determine the return direction.
J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2201

Fig. 3. Flow chart of the fracture model for concrete cracking.


2202 J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220

Fig. 4. Menétrey–Willam failure criterion represented in Haigh–Westergaard coordinates: (a) 3D stress space, (b)
Rendulic plane and (c) deviatoric plane.

In the present plasticity model, the Menétrey and Willam (1995) three parameter failure
surface is used (Fig. 4):

 2 !
p
pffiffiffiffiffiffiffi q q n
f ðn;q;hÞ ¼ 1:5 þ m  pffiffiffi rðh;eÞ þ pffiffiffi  cðjÞ ¼ 0
kðjÞ  fc 6  kðjÞ  fc 3  kðjÞ  fc
ð25Þ

where:

ðkðjÞ  fc Þ2  ðkt  ft Þ2 e
m¼3  is the cohesion parameter of the material and ð26Þ
kðjÞ  fc  kt  ft eþ1
4ð1  e2 Þ cos2 h þ ð2e  1Þ2
rðh;eÞ ¼ is an elliptic function
2ð1  e2 Þ cos h þ ð2e  1Þ½4ð1  e2 Þcos2 h þ 5e2  4e1=2
ð27Þ
In the above equations, (n, q, h) are the Haigh–Westergaard coordinates and fc and ft
are the compressive strength and tensile strength respectively. Parameter kt > 1 is a scaling
value for the tensile concrete strength in order to provide intersection between the Rankine
(fracture) and the Menétrey–Willam (plasticity) failure surfaces during the combination
procedure (Fig. 5). Parameter e 2 (0.5, 1.0) defines the roundness of the Menétrey–Willam
failure surface, with a recommended value e = 0.52 (Menétrey and Willam, 1995) leading
to equibiaxial concrete strength equal to fbc = 1.14  fc, very similar to the experimental
work by Kupfer et al. (1969).
The position of the Menétrey–Willam failure surface is not fixed but it can expand and
move along the hydrostatic axis (simulating the hardening and softening stages), based on
the value of the hardening/softening parameter (j). In the current model, this parameter
identifies with the volumetric plastic strain (Grassl et al., 2002):

dj ¼ depv ¼ dep1 þ dep2 þ dep3 ð28Þ


J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2203

Fig. 5. (a) Intersection of Rankine and Menétrey–Willam failure surfaces for kt = 2 and (b) no intersection for
kt = 1.

The instantaneous shape and location of the loading surface during hardening is
defined by a hardening function (k), which depends on the hardening/softening parameter
(j). This function is directly incorporated in the Menétrey–Willam failure surface Eqs. (25)
and (26), operating as a scaling factor on the compressive concrete strength (fc). It has the
following elliptic form (Červenka et al., 1998):
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p 2
ev;t  epv
kðjÞ ¼ kðepv Þ ¼ k o þ ð1  k o Þ  1  ð29Þ
epv;t

where epv;t is the plastic volumetric strain at uniaxial concrete strength (onset of softening)
and ko is the value that defines the initial yield surface that bounds the initial elastic regime
(onset of plasticity). At the end of the hardening process, the hardening function retains a
constant value of unity and the material enters the softening regime, which is controlled by
the softening function (c). This function simulates the material decohesion by shifting the
2204 J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220

loading surface along the negative hydrostatic axis. It is assumed that it follows the soften-
ing function originally proposed by Van Gysel and Taerwe (1996) for uniaxial compression:
0 12
B 1 C
cðjÞ ¼ cðepv Þ ¼ @  2 A ð30Þ
n1 1
1þ n2 1

where:
n1 ¼ epv =epv;t ð31Þ
n2 ¼ ðepv;t þ tÞ=epv;t ð32Þ
Parameter t in Eq. (32) controls the slope of the softening function and the outmost
square is necessary due to the quadratic nature of the loading surface. The softening func-
tion value starts from unity and complete material decohesion is attained at c = 0. The
evolution of both hardening and softening functions with respect to the hardening/soften-
ing parameter is schematically shown in Fig. 6.
The plasticity model incorporates a non-associated flow rule using a polynomial plastic
potential function (g), with Lode angle (h) dependency and adjustable order (n):
 n 
q 1 q n
g ¼A pffiffiffi þ C þ ðB  CÞð1  cos 3hÞ  pffiffiffi þ pffiffiffi a
k  c  fc 2 k  c  fc k  c  fc
ð33Þ
Parameters A, B and C define the shape of the plastic potential function in stress space and
their calibration is based on the assumption that the inclination (w) of the incremental
plastic strain vector identifies with the inclination of the total plastic strain vector at three
distinct stress states, namely the uniaxial, equibiaxial and triaxial compressive concrete

k(κ ) / c(κ )
c k
1.0

k
0.8

0.6
c

0.4

0.2
ko

0.0
p κ = εpv
ε v,t

Fig. 6. Evolution of hardening (k) and softening (c) functions with respect to the plastic volumetric strain.
J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2205

Fig. 7. Direction (w) of the incremental (a) and total (b) plastic strain vectors.

strength (Fig. 7). The presented model adopts a regular flow rule in a sense that the plastic
flow direction depends solely on the location on the yield surface, and is not affected by the
loading direction (Nicot and Darve, 2007). A detailed calibration scheme for the plasticity
model parameters, based on an extensive experimental database can be found in Papa-
nikolaou and Kappos (2007) and suggested values (including the fracture model parame-
ters) for various uniaxial compressive concrete strengths (fc) are shown in Table 1.
For the integration of the constitutive equations, an implicit backward-Euler return-
mapping algorithm was applied, incorporating a regula-falsi/secant iterative scheme (Čer-
venka et al., 1998). In order to maintain consistency with constitutive equations, the above
scheme iteratively updates the following variables: (1) the plastic multiplier (dk) which
yields both the stress corrector ðdkDijkl mkkl Þ and the additional plastic strains ðþdkmkij Þ,
(2) the return direction ðmkij Þ of the stress vector towards the Menétrey–Willam failure sur-
face and (3) the hardening/softening functions (k(j), c(j)), which control the position and
the size of the Menétrey–Willam failure surface (f) and the plastic potential surface (g) in
the stress space. The input data (step n) are the current stress tensor (nrij), the plastic strain
tensor ðn epij Þ, describing the deformation history (load path dependency) of the material,
and the trial total strain increment (deij). The output data (step n + 1) are the updated non-
linear stress tensor (n + 1rij) and the updated plastic strain tensor ðnþ1 epij Þ. The suggested
algorithm is numerically stable, independent of the load step size and does not require
the differentiation of the Menétrey–Willam failure surface. A detailed flow chart of the
procedure is shown in Fig. 8 and a schematic depiction of the stress return process is
shown in Fig. 9.

2.3. Combination of plasticity and fracture models

The objective is to combine the above models into a single model so that the plasticity
model is used for concrete crushing and the fracture model for cracking. This problem can
be generally stated as a simultaneous solution of the two following inequalities:
2206 J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220

Fig. 8. Flow chart of the backward-Euler return-mapping algorithm.

Fig. 9. Schematic depiction of the return process.


J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2207

f p ðn rij þ Dijkl  ðdekl  defkl  depkl ÞÞ 6 0 ! solve for depkl ð34Þ


f n
f ð rij þ Dijkl  ðdekl  depkl  defkl ÞÞ 6 0 ! solve for defkl ð35Þ
Each of the above expressions depends on the output from the other, and therefore the
following iterative scheme is applied:
fði1Þ corði1Þ pðiÞ pðiÞ
Step 1 : f p ðn rij þ Dijkl  ðde
kl dekl þ b  dekl  dekl ÞÞ 6 0 ! solve for dekl
ð36Þ
pðiÞ fðiÞ fðiÞ
Step 2 : f f ðn rij þ Dijkl  ðdekl  dekl  dekl ÞÞ 6 0 ! solve for dekl ð37Þ
corðiÞ fðiÞ fði1Þ
Step 3 : deij ¼ deij  deij ð38Þ
Iterative correction of the strain norm between two subsequent iterations is expressed as
corðiÞ corði1Þ
kdeij k ¼ ð1  bÞ  af  ap  kdeij k ð39Þ
where:
fðiÞ fði1Þ
kdeij  deij k
af ¼ pðiÞ pði1Þ
ð40Þ
kdeij  deij k
pðiÞ pði1Þ
kdeij  deij k
ap ¼ corði1Þ
ð41Þ
kdeij k
and b is a relaxation factor, which is introduced in order to guarantee convergence. Its cal-
culation is based on the runtime values of af and ap, so that the convergence of the iterative
scheme can be assured. The parameters af and ap characterize the mapping properties of
each model (i.e. plastic and fracture). It is possible to consider each model as an operator,
which maps strain increment on the input into a fracture or plastic strain increment on the
output. The product of the two mappings must be contractive in order to guarantee con-
vergence. The necessary condition for convergence is
ð1  bÞ  af  ap < 1 ð42Þ
f p
It can be shown that the values of a and a are directly proportional to the softening rate
in each model. Since the softening model remains usually constant for a material model
and a finite element, their values do not change significantly between iterations. It is pos-
sible to select the scalar b, such that Eq. (42) is always satisfied at the end of each iteration,
based on the current values of af and ap. There are three possible scenarios, which must be
handled, for the appropriate calculation of b:
ð1Þ jaf  ap j 6 v ð43Þ
where v is related to the requested convergence rate. For a linear rate it can be set to v = 1/2.
In this case the convergence is satisfactory and b is set to zero.
ð2Þ v < jaf  ap j < 1 ð44Þ
jaf  ap j
then the convergence would be slow. In this case b should be calculated as b ¼ 1  v ,
in order to increase the convergence rate.
ð3Þ 1 6 jaf  ap j ð45Þ
v
then the algorithm is diverging. In this case b should be calculated as b ¼ 1  f p to
ja  a j
stabilize the iterations.
2208 J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220

This approach guarantees convergence as long as the parameters af and ap do not


change drastically between iterations which is satisfied for smooth and correctly formu-
lated finite element models. The rate of convergence depends on material brittleness, dila-
tion rate and finite element size. It is advantageous to further stabilize the algorithm by
smoothing the relaxation factor (b) during the iterative process:
bi þ bi1
b¼ ð46Þ
2
where the superscripts i and i  1 denote values from two subsequent iterations. This will
eliminate problems due to possible oscillation of the relaxation factor. An important con-
dition for the convergence of the above algorithm is that the failure surfaces of the two
models (Rankine and Menétrey–Willam) are intersecting each other in all possible posi-
tions during hardening and softening, which is provided by parameter kt (Fig. 5). Addi-
tional constraints are used in the iterative algorithm; if the stress state at the end of the
first step violates the Rankine criterion, the order of the first two steps of the algorithm
(36 and 37) is reversed. Also in reality concrete crushing in one direction has an effect
on the cracking in other directions. It is assumed that after the plasticity yield criterion
is violated, the tensile strength in all material directions is gradually reduced to zero.
The proposed algorithm for the combination of plastic and fracture models is schemat-
ically shown in Fig. 10. When both surfaces are activated, the behavior is quite similar to
the multi-surface plasticity (Simo et al., 1988). However, contrary to the multi-surface
plasticity algorithm, the proposed method is more general in the sense that it covers all
loading regimes including physical changes such as crack closure. Currently, it is devel-
oped only for two interacting models, and its extension to multiple models is not
straightforward.

Fig. 10. Schematic description of the iterative process shown in two dimensions for clarity.
J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2209

2.4. Model verification

A constitutive driver was developed for testing various stress–strain loading histories in
order to investigate the behavior of the proposed model. In this section, the constitutive
driver is used to verify the numerical behavior of the proposed algorithm for the combi-
nation of the fracture and plastic models. Three distinct loading scenarios were considered:

(1) Uniaxial compression beyond the maximum compressive concrete strength (crush-
ing), then unloading into tension and then reloading again into compression. In
Fig. 11 it is shown that during unloading, a crack opens in the loading direction 3
just after the attainment of concrete tensile strength (which has been reduced due
to earlier crushing) and eventually closes when the loading is reversed. The material
response thereafter follows the same uniaxial compression load path, just before
unloading took place.
(2) Crack closure due to crushing in another material direction (Fig. 12). In this example
a crack is first introduced in direction 1. After the crack has fully opened and no
more stresses can be transferred across the crack, displacements are constrained in
direction 1 and compressive strains are applied in direction 3. After crushing starts
in direction 3, the plastic strains in direction 1 will ultimately close the crack. After
the crack in direction 1 closes, a biaxial compression state is recovered. At this point,
stress r3 should further increase until it reaches a (second) softening branch, which
corresponds to the appropriate biaxial stress state.
(3) In the third example, loading is applied in all three material directions (Fig. 13).
Equivalent magnitudes of tensile strain increments are first applied in directions 1
and 2, introducing two cracks respectively. Displacements are then constrained in
directions 1 and 2 and compressive strains are applied in direction 3. This means that
all three surfaces and both models are active. After crushing in direction 3, both
cracks will close and a triaxial compression state will be recovered.

fc= 30 MPa ft= 2.45 MPa


-35
3 fc
-30

-25

-20

-15

-10
reduced tensile
-5 strength due to
crack opens crushing
0
ft 3
5
0 -0.001 -0.002 -0.003 -0.004 -0.005 -0.006

Fig. 11. Example of concrete uniaxial compression-tension-compression load path.


2210 J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220

fc= 30 MPa ft= 2.45 MPa


-35
biaxial stress state
3 fc
-30

crack closes
-25

-20

-15

-10 First load stage : direction 1 (x)

-5 Second load stage : direction 3 (z)


crack opens
0
ε1 ft 3
5
0.002 0 -0.002 -0.004 -0.006 -0.008

Fig. 12. Example of concrete biaxial stress state, introduced by crack closure due to concrete crushing.

fc = 30 MPa ft = 2.45 MPa


-90
3 trixial stress state
-80

-70
First load stage : directions 1 and 2 (x and y)
-60
Second load stage : direction 3 (z)
-50

-40
fc
-30
both cracks close
-20

-10
both cracks open
0
1 = 2 ft 3
10
0.003 0.002 0.001 0.000 -0.001 -0.002 -0.003 -0.004 -0.005 -0.006

Fig. 13. Example of concrete triaxial stress state, introduced by crack closure due to concrete crushing.

The successful treatment of the above complex loading scenarios by the combination
algorithm validates its numerical stability in the context of the present material model.
In order to assess the accuracy of its performance, the model was further evaluated against
several experimental results from the literature on uniaxial, biaxial and triaxial tests (Figs.
14–20). The correlation is reasonable, which renders the model applicable to large scale
finite element analysis of plain and reinforced concrete structures.
J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2211

-120

3 (MPa) fc = 105.7 MPa

-100
93.9

-80

65.0
-60
50.3

-40
31.7

21.7
-20

3
0
0.000 -0.001 -0.002 -0.003 -0.004 -0.005 -0.006

Fig. 14. Comparison between numerical and experimental results on uniaxial compression (Dahl, 1992).

1 (MPa) ft = 2.6 MPa


3.2
( 2 = 0.55 1)
( 2 = 0.55 1)
3 ( 2 = 0.55 1)
2.8 1
3 2, 3 1, 2 1
( 2 = 0) 2 ( 2= 1)

2.4
( 2= 1) ( 2 = 0)
2.0

1.6
1 : tensile
2 : tensile
1.2
3 =0

0.8
2 = 0 : uniaxial extension
2= 1 : equibiaxial extension
0.4

(-) (+)
0.0
-0.00005 0.00000 0.00005 0.00010

Fig. 15. Comparison between numerical and experimental results on uniaxial and biaxial tension (Kupfer et al.,
1969).

3. Applications in finite element analysis

The suggested model was integrated in the general finite element package ATENA
(Červenka et al., 2007) and was applied on three dimensional analysis of reinforced
2212 J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220

3 (MPa) fc = 32.0 MPa


-45
1 2 3
-40
( 2 = 0.52 1)

-35 1
2, 3
( 2= 1)
-30
1, 2
-25
3

-20 ( 2 = 0)
1 =0
2 : compressive
-15
3 : compressive

-10
2 = 0 : uniaxial compression
2= 1 : equibiaxial compression
-5
(+) (-)
0
0.005 0.003 0.001 -0.001 -0.003 -0.005

Fig. 16. Comparison between numerical and experimental results on biaxial tension-compression (Kupfer et al.,
1969).

3 (MPa) fc = 32.0 MPa


-35
1, 2 1 : tensile 3 ( 1 = 0)
2 =0
-30
3 : compressive

-25

3
-20
1
( 1/ 3 = -0.1)
2
-15
1 2 3
( 1/ 3 = -0.2)
-10

-5

(+) (-)
0
0.0015 0.0010 0.0005 0.0000 -0.0005 -0.0010 -0.0015 -0.0020 -0.0025

Fig. 17. Comparison between numerical and experimental results on biaxial compression (Kupfer et al., 1969).

concrete structures. In the first example, the well known Leonhardt’s shear beam (Leon-
hardt and Walther, 1962) was modeled using three dimensional solid elements for concrete
and embedded linear truss elements for steel. This structural element was selected because
- according to experimental evidence - it exhibits both tensile and compressive stress states
J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2213

3 (MPa) fc = 28.6 MPa


-120

1= 2 = 21 Pa
-100

14.7 Pa
-80

8.4 Pa
-60

4.2 Pa
-40

2.1 Pa

-20 1.05 Pa

1= 2 3
0
0.04 0.03 0.02 0.01 0 -0.01 -0.02 -0.03 -0.04 -0.05

Fig. 18. Comparison between numerical and experimental results on triaxial compression of normal strength
concrete (Imran, 1994).

3 (MPa) fc = 60.2 MPa


-200

1 = 2 = 29.3 Pa
23.3 Pa
-160
20.3 Pa

14.3 Pa
-120
11.3 Pa

8.3 Pa
-80
5.3 Pa

2.3 Pa
-40

0 Pa 3
0
0 -0.005 -0.01 -0.015 -0.02 -0.025 -0.03 -0.035 -0.04

Fig. 19. Comparison between numerical and experimental results on triaxial compression of high strength
concrete (Xie et al., 1995).

together with shear failure, which is expected to activate both parts (fracture and plastic-
ity) of the present combined constitutive model. Fig. 21 shows the model geometry, mate-
rial properties and solution parameters. Due to symmetry, half length of the beam was
modeled using proper boundary conditions on the symmetry plane. Fig. 22 shows a com-
parison between numerical and experimental results (applied vertical force versus vertical
deflection at midspan), the crack development (starting from small diffused flexural cracks
2214 J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220

near the midspan which concludes to a large diagonal crack between loading and support
points) and the embedded reinforcement axial stresses. It is observed that correlation is
very reasonable concerning both the load–displacement response and the failure mode
(shear failure).
The second example exhibits the confinement sensitivity of the present model
under compressive loading. A confined square reinforced concrete column, tested

3 (MPa) fc = 60.6 MPa


-140

1= 2 = 12 Pa
-120

-100 8 MPa

-80
4 Pa

-60

0 MPa
-40

-20

1= 2 3
0
0.015 0.01 0.005 0 -0.005 -0.01 -0.015 -0.02

Fig. 20. Comparison between numerical and experimental results on triaxial compression of high strength
concrete (Candappa et al., 2001).

Concrete strength : 28.5 MPa

Number of brick elements : 5416


Number of truss elements : 2

Prescribed displacements with


Newton-Raphson solution

Model geometry (half length due to symmetry) Finite Element Mesh

Fig. 21. Three dimensional finite element model of the shear beam (Leonhardt and Walther, 1962).
J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2215

80 Failure at 0.0033 m
Applied force (kN) (d)
70
(c)
60

50
(b)
40
(a)
30 Analysis (a) Max principal strains and cracks @ 0.0005 m
Experiment
20

10
Vertical deflection @ midspan (m)
0
0 0.001 0.002 0.003 0.004

Applied force - vertical deflection at midspan


(b) Max principal strains and cracks @ 0.001 m

Experimental shear failure mode (c) Max principal strains and cracks @ 0.002 m

Embedded reinforcement axial stresses @ 0.0033 m (d) Max principal strains and cracks @ 0.0033 m

Fig. 22. Nonlinear response and crack patterns of the shear beam.

experimentally by Sheikh and Uzumeri (1980) is modeled using solid elements for con-
crete and embedded truss elements for longitudinal and transverse reinforcement (Papa-
nikolaou and Kappos, 2005). Due to double symmetry, one quarter of the column was
modeled, the appropriate boundary conditions were applied on the symmetry planes and
concentric compressive displacements were prescribed. Fig. 23 shows the concrete finite
element mesh, the embedded reinforcement bars and the model loading and boundary
conditions. Fig. 24 shows a comparison between numerical and experimental results.
Reasonable correlation is observed concerning both the maximum strength (2.4% differ-
ence) and the post-peak response. Moreover, the experimentally observed cracking of the
concrete cover was successfully captured by the analysis. It has to be noted here that it is
necessary that the finite element size is smaller than concrete cover thickness in order to
successfully capture the effect of cover cracking. Finally, a parametric study using
2216 J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220

Fig. 23. Three dimensional finite element model of the confined concrete column (Sheikh and Uzumeri, 1980).

Sheikh and Uzumeri (1980) - Column 2A1-1 fc = 31.88 MPa


4000
P(kN)
3500

3000

2500

2000

1500
Analysis
Core
Concrete cover concrete Experiment
1000

500

0
0 0.001 0.002 0.003 0.004 0.005 0.006

Fig. 24. Load–displacement response, axial stress distribution and cover cracking of the reinforced concrete
column.

different transverse reinforcement arrangements of increasing complexity (Fig. 25) clearly


shows that the present constitutive model successfully captures the expected increase in
strength and deformation capacity (reducing slope of the post-peak response curve) due
to the passive confinement effect.
J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2217

fc = 27.97 MPa
6000
P(kN)

5000
C

*Arrangement C
4000

3000

2000
Arrangement B
Longitudinal reinforcement
1000 Plain concrete

0
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.008 0.009 0.01

Sheikh and Uzumeri (1980) - Column 2C5-17 * Arrangement A

Fig. 25. Increase in strength and deformation capacity of the reinforced concrete column, for increasing
complexity of transverse reinforcement arrangements.

4. Conclusions

A combined material model for concrete was presented, including an orthotropic


smeared crack model for concrete cracking based on the Rankine failure criterion and a
plasticity model for concrete crushing based on the Menétrey–Willam criterion. The hard-
ening/softening parameter for the plasticity model was related to the plastic volumetric
strain, interacting with a nonlinear plastic potential function. Both models are formulated
separately and a combination algorithm is developed to iteratively determine the separa-
tion of the strain increment into the fracturing and plastic part. The behavior of this algo-
rithm was verified against several loading histories and reasonable correlation with
experimental results was generally observed. The suggested model was successfully inte-
grated in the ATENA finite element software (Červenka et al., 2007) and was applied
on the analysis of reinforced concrete beams under flexure and confined reinforced con-
crete columns under concentric compression. The comparison with experimental results
was reasonable and it is therefore applicable for further applications on finite element
analysis of plain and reinforced concrete structures.
The proposed model could be further improved in the following areas: It is docu-
mented in the work by Collins (1978) that the compressive strength of cracked concrete
should be reduced and should depend on the crack opening in other directions. This
assumption may improve the accuracy in the prediction of the shear strength of rein-
forced concrete beams. Another important aspect is the shear behavior after cracking.
The current model intuitively relates the shear stiffness of the cracked material to the
normal crack opening law. There are experimental results in the literature, for instance
Walraven (1980), which could be used to improve the shear behavior of the model.
Finally, the plasticity part of the model could be further enhanced to include a cap
2218 J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220

surface for hydrostatic compression and handle the initial plastic compaction of concrete
that is observed experimentally.

Acknowledgements

The contribution of the first author was supported by the research Grant number
1ET409870411 from the Czech Academy of Sciences. The financial support is greatly
appreciated. The contribution of the second author was performed within the framework
of the research project ‘‘ASProGe: Seismic Protection of Bridges”, funded by the General
Secretariat of Research and Technology (GGET) of Greece.

References

Bažant, Z.P., Oh, B.H., 1983. Crack band theory for fracture of concrete. Materials and Structures, RILEM 16
(3), 155–177.
Bažant, Z.P., Pijaudier-Cabot, G., 1987. Nonlocal continuum damage, localization instability and convergence.
Journal of Applied Mechanics, ASME 55 (2), 287–293.
Bažant, Z.P., Caner, F.C., Carol, I., Adley, M.D., Akers, S.A., 2000. Microplane model M4 for concrete: I
Formulation with work-conjugate deviatoric stress. Journal of Engineering Mechanics, ASCE 126 (9), 944–
961.
Bielger, M.W., Mehrabadi, M.M., 1995. An energy-based constitutive model for anisotropic solids subject to
damage. Mechanics of Materials 19 (2–3), 151–164.
Candappa, D.C., Sanjayan, J.G., Setunge, S., 2001. Complete triaxial stress–strain curves of high-strength
concrete. Journal of Materials in Civil Engineering, ASCE 13 (3), 209–215.
Carol, I., Rizzi, E., Willam, K., 1994. A unified theory of elastic degradation and damage based on a loading
surface. International Journal of Solids and Structures 31 (30), 2835–2865.
Červenka, V., Gerstle, K., 1971. Inelastic analysis of reinforced concrete panels Part I: Theory. Publication
I.A.B.S.E. 31 (11), 32–45.
Červenka, V., Pukl, R., Ozbolt, J., Eligehausen, R., 1995. Mesh sensitivity effects in smeared finite element
analysis of concrete structures. In: Proceedings of the 2nd International Conference on Fracture Mechanics of
Concrete Structures – FraMCoS 2, 1995, pp. 1387–1396.
Červenka, J., Červenka, V., Eligehausen, R., 1998. Fracture–plastic material model for concrete. Application to
analysis of powder actuated anchors. In: Mihashi, H., Rokugo, K. (Eds.), Proceedings of the 3rd
International Conference on Fracture Mechanics of Concrete Structures – FraMCoS 3, Gifu, Japan, vol. 2.
Aedificatio Publishers, Freiburg, Germany, pp. 1107–1116.
Červenka, V., Jendele, L., Červenka, J., 2007. ATENA Program Documentation. Part 1: Theory. Červenka
Consulting, Prague, Czech Republic.
Chiarelli, A.S., Shao, J.F., Hoteit, N., 2003. Modeling of elastoplastic damage behavior of a claystone.
International Journal of Plasticity 19 (1), 23–45.
Christoffersen, J., Mehrabadi, M.M., Nemat-Nasser, S., 1981. A micromechanical description of granular
material behavior. Journal of Applied Mechanics, ASME 48 (2), 339–344.
Cicekli, U., Voyiadjis, G.Z., Abu Al-Rub, R.K., 2007. A plasticity and anisotropic damage model for plain
concrete. International Journal of Plasticity 23 (10–11), 1874–1900.
Collins, M.P., 1978. Towards a rational theory for RC members in shear. Journal of the Structural Division,
ASCE 104 (4), 649–666.
Contrafatto, L., Cuomo, M., 2006. A framework of elastic–plastic damaging model for concrete under multiaxial
stress states. International Journal of Plasticity 22 (12), 2272–2300.
Dahl, K.K.B., 1992. A constitutive model for normal and high-strength concrete. ABK Report No. R287,
Department of Structural Engineering, Technical University of Denmark.
De Borst, R., 1986. Non-linear analysis of frictional materials. Ph.D. Thesis, Delft University of Technology, The
Netherlands.
De Borst, R., Mühlhaus, H.B., 1992. Gradient dependant plasticity: formulation and algorithmic aspects.
International Journal for Numerical Methods in Engineering 35 (3), 521–539.
J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220 2219

Etse, G., 1992. Theoretische und numerische untersuchung zum diffusen und lokalisierten versagen in beton.
Ph.D. Thesis, University of Karlsruhe.
Feenstra, P.H., 1993. Computational aspects of biaxial stress in plain and reinforced concrete. Ph.D. Thesis, Delft
University of Technology, The Netherlands.
Feenstra, P.H., Rots, J.G., Amesen, A., Teigen, J.G., Hoiseth, K.V., 1998. A 3D constitutive model for concrete
based on co-rotational concept. In: Proceedings of EURO-C, 1998, vol. 1, pp. 13–22.
Grassl, P., Jirásek, M., 2006. Damage-plastic model for concrete failure. International Journal of Solids and
Structures 43 (22–23), 7166–7196.
Grassl, P., Lundgren, K., Gylltoft, K., 2002. Concrete in compression: a plasticity theory with a novel hardening
law. International Journal of Solids and Structures 39 (20), 5205–5223.
Hashiguchi, K., Tsutsumi, S., 2007. Gradient plasticity with the tangential-subloading surface model and the
prediction of shear-band thickness of granular materials. International Journal of Plasticity 23 (5), 767–797.
Hordijk, D.A., 1991. Local approach to fatigue of concrete. Ph.D. Thesis, Delft University of Technology, The
Netherlands.
Imran, I., 1994. Applications of non-associated plasticity in modeling the mechanical response of concrete. Ph.D.
Thesis, Department of Civil Engineering, University of Toronto.
Jason, L., Huetra, A., Pijaudier-Cabot, G., Ghavamian, S., 2006. An elastic–plastic damage formulation for
concrete: application to elementary tests and comparison with an isotropic damage model. Computer
Methods in Applied Mechanics and Engineering 195 (52), 7007–7092.
Kupfer, H., Hilsdorf, K., Rusch, H., 1969. Behavior of concrete under biaxial stresses. ACI Journal 66 (8), 656–
666.
Lee, J., Fenves, G.L., 1998. Plastic-damage model for cyclic loading of concrete structures. Journal of
Engineering Mechanics, ASCE 124 (8), 892–900.
Leonhardt, F., Walther, R., 1962. Schubversuche an einfeldrigen stahlbetonbalken mit und ohne schubbeweh-
rung. Deutscher Ausschuss für Stahlbeton, 151, Berlin, West Germany.
Mattei, N.J., Mehrabadi, M.M., Huaning, Z., 2007. A micromechanical constitutive model for the behavior of
concrete. Mechanics of Materials 39 (4), 357–379.
Mehrabadi, M.M., Nemat-Nasser, S., Oda, M., 1982. On statistical description of stress and fabric in granular
materials. International Journal for Numerical and Analytical Methods in Geomechanics 6 (1), 95–108.
Menétrey, P., Willam, K.J., 1995. Triaxial failure criterion for concrete and its generalization. ACI Structural
Journal 92 (3), 311–318.
Menétrey, P., Walther, R., Zimmermann, T., Willam, K.J., Regan, P.E., 1997. Simulation of punching failure in
reinforced-concrete structures. Journal of Structural Engineering, ASCE 123 (5), 652–659.
Meschke, G., Lackner, R., Mang, H.A., 1988. An anisotropic elastoplastic-damage model for plain concrete.
International Journal for Numerical Methods in Engineering 42 (4), 703–727.
Mohamad-Hussein, A., Shao, J.F., 2007. Modelling of elastoplastic behaviour with non-local damage in concrete
under compression. Computers and Structures 85 (23–24), 1757–1768.
Nicot, F., Darve, F., 2007. Basic features of plastic strains: from micro-mechanics to incrementally nonlinear
models. International Journal of Plasticity 23 (9), 1555–1588.
Oda, M., Nemat-Nasser, S., Mehrabadi, M.M., 1982. A statistical study of fabric in a random assembly of
spherical granules. International Journal for Numerical and Analytical Methods in Geomechanics 6 (1), 77–
94.
Olivier, J., 1989. A consistent characteristic length for smeared cracking models. International Journal for
Numerical Methods in Engineering 28 (2), 461–474.
Owen, J.M., Figueiras, J.A., Damjanic, F., 1983. Finite element analysis of reinforced and prestressed concrete
structures including thermal loading. Computer Methods in Applied Mechanics and Engineering 41 (3), 323–
366.
Papanikolaou, V.K., Kappos, A.J., 2005. Modelling confinement in concrete columns and bridge piers through
3D nonlinear finite element analysis. In: fib Symposium Keep Concrete Attractive, Budapest, Hungary, May
2005, pp. 488–495.
Papanikolaou, V.K., Kappos, A.J., 2007. Confinement-sensitive plasticity constitutive model for concrete in
triaxial compression. International Journal of Solids and Structures 44 (21), 7021–7048.
Pramono, E., Willam, K.J., 1989. Fracture energy-based plasticity formulation of plain concrete. Journal of
Engineering Mechanics, ASCE 115 (6), 1183–1204.
Rashid, Y.R., 1968. Analysis of prestressed concrete pressure vessels. Nuclear Engineering and Design 7 (4), 334–
344.
2220 J. Červenka, V.K. Papanikolaou / International Journal of Plasticity 24 (2008) 2192–2220

Rots, J.G., Blaauwendraad, J., 1989. Crack models for concrete: discrete or smeared? Fixed, multi-directional or
rotating? Heron 34 (1).
Sheikh, S.A., Uzumeri, S.M., 1980. Strength and ductility of tied concrete columns. Journal of the Structural
Division, ASCE 108 (4), 929–950.
Simo, J.C., Ju, J.W., 1987. Strain and stress-based continuum damage models: I Formulation, II Computational
aspects. International Journal of Solids and Structures 23 (7), 821–869.
Simo, J.C., Kennedy, J.G., Govindjee, S., 1988. Non-smooth multisurface plasticity and viscoplasticity loading/
unloading conditions and numerical algorithms. International Journal for Numerical Methods in Engineering
26 (10), 2161–2185.
Van Gysel, A., Taerwe, L., 1996. Analytical formulation of the complete stress–strain curve for high strength
concrete. Materials and Structures, RILEM 29 (193), 529–533.
Walraven, J.C., 1980. Aggregate Interlock: A Theoretical and Experimental Analysis. Delft University of
Technology, The Netherlands.
Wilkins, M.L., 1964. Calculation of elastic–plastic flow. Methods of Computational Physics, vol. 3. Academic
Press, New York, pp. 211–263.
Xie, J., Elwi, A.E., MacGregor, J.G., 1995. Mechanical properties of three high-strength concretes containing
silica fume. ACI Materials Journal 92 (2), 135–145.

S-ar putea să vă placă și