You are on page 1of 24

Comparative study on nite elements with embedded

discontinuities
Milan Jirasek
1
Laboratory of Structural and Continuum Mechanics, Department of Civil Engineering, Swiss Federal Institute of Technology (EPFL),
CH-1015 Lausanne, Switzerland
Received 29 January 1998
Abstract
The recently emerged idea of incorporating strain or displacement discontinuities into standard nite element interpolations has
triggered the development of powerful techniques that allow ecient modeling of regions with highly localized strains, e.g. of fracture
process zones in concrete or shear bands in metals or soils. Following the pioneering work of Ortiz, Leroy, and Needleman, a number
of studies on elements with embedded discontinuities were published during the past decade. It was demonstrated that local enrich-
ments of the displacement and/or strain interpolation can improve the resolution of strain localization by nite element models. The
multitude of approaches proposed in the literature calls for a comparative study that would present the diverse techniques in a unied
framework, point out their common features and dierences, and nd their limits of applicability. There are many aspects in which
individual formulations dier, such as the type of discontinuity (weak/strong), variational principle used for the derivation of basic
equations, constitutive law, etc. The present paper suggests a possible approach to their classication, with special attention to the type
of kinematic enhancement and of internal equilibrium condition. The dierences between individual formulations are elucidated by
analyzing the behavior of the simplest nite element the constant-strain triangle (CST). The sources of stress locking (spurious stress
transfer) reported by some authors are analyzed. It is shown that there exist three major classes of models with embedded disconti-
nuities but only one of them gives the optimal element behavior. 2000 Elsevier Science S.A. All rights reserved.
Keywords: Finite elements; Strain discontinuity; Displacement discontinuity; Cracking; Shear bands; Localization; Variational
principles; Enhanced assumed strain; Stress locking
1. Introduction
Concrete cracking has been traditionally modeled either by the discrete approach, which lumps the
additional deformation due to cracking into a displacement discontinuity (opening of a localized crack), or
by the smeared approach, which treats this deformation as cracking strain distributed over a certain ma-
terial volume. Both techniques have their advantages and drawbacks, often discussed at international
conferences and workshops. The recently emerged concept of strain or displacement discontinuities em-
bedded into standard nite elements combines the strong points of both approaches.
As observed already by Rots [1] and other researchers, traditional smeared-crack models for concrete
fracture suer by stress locking, i.e., by spurious stress transfer across a widely open crack. For xed crack
models with a nonzero retention factor, locking is mainly due to shear stresses generated by a rotation of
the principal strain axes after the crack initiation. However, locking is observed even for rotating crack
models, which keep the principal axes of strain and stress aligned, so that stresses tangential to the crack
www.elsevier.com/locate/cma
Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330
1
Tel.: 41-21-6932413; fax: 41-21-693676.
E-mail address: milan@1scsun5.ep.ch (M. Jirasek).
0045-7825/00/$ - see front matter 2000 Elsevier Science S.A. All rights reserved.
PII: S 0 0 4 5 - 7 8 2 5 ( 9 9 ) 0 0 1 5 4 - 1
cannot arise. The source of this phenomenon was analyzed in [2]. It was shown that the spurious stress
transfer is caused by a poor kinematic representation of the discontinuous displacement eld around a
macroscopic crack. Unless the direction of the macroscopic crack (represented by a band of cracking el-
ements) happens to be parallel to element sides, the directions of maximum principal strain determined
from the nite element interpolation at individual Gauss points deviate from the normal to the macroscopic
crack. The lateral principal stress has a nonzero projection on the crack normal, which generates cohesive
forces acting across the crack even at very late stages of the cracking process when the crack should be
completely stress-free.
The remedy proposed in [3] was based on the observation that a scalar damage model does not exhibit
this type of locking because all stress components tend to zero as the crack opens wide. As locking appears
only at late stages of the damage process, it is natural to start with a constitutive model that captures the
anisotropy induced by cracking, e.g., with the rotating crack model, and switch to a scalar damage for-
mulation only when the crack opening reaches a certain critical value. The examples presented in [3,4]
demonstrated that the combined rotating crack model with transition to scalar damage (RCSD) can re-
produce the entire experimentally determined load-displacement diagram for a number of fracture tests
reported in the literature.
Alternatively, stress locking can be avoided by improving the kinematic representation of highly
localized strains. This is the usual approach adopted by a number of researchers during the past decade
[646]. Techniques based on this idea were presented under dierent names but all of them incorporate a
discontinuity (of strains or displacements) into the interior of a nite element. The aim of the present study
is to compare various versions of the embedded discontinuity approach reported in the literature, evaluate
their performance, and identify a method suitable for practical application in fracture simulations. A
combination of the smeared and embedded crack approaches and an extension of the combined model to a
nonlocal formulation, briey outlined in [47], will be addressed in detail in a separate paper [48].
The present publication emphasizes applications to fracture problems, but the theoretical comparison
and classication of models is fully general and applies to modeling of shear bands and slip lines equally well.
2. Literature survey
Already in 1980, Johnson and Scott [5] realized that conventional nite element methods (displacement
methods) for plasticity problems are based on using continuous trial functions and thus are not particularly well
adapted to the nature of the true solution. They proposed to modify the shape functions such that dis-
placement discontinuities can be captured. However, their work was limited to perfect plasticity and they
presented only a simple one-dimensional example.
An early attempt to improve the kinematic description of a cracking element is due to Droz [6]. He
considered a perfectly brittle material, for which the normal stress drops down to zero immediately after the
formation of a crack. The elastic stressstrain law can be written as
r = D
e
e; (1)
where r is the column matrix of stress components, e the column matrix of strain components, and D
e
the
elastic stiness matrix, for plane stress given by
D
e
=
E
1 m
2
1 m 0
m 1 0
0 0
1m
2
_
_
_
_
(2)
in which E is Young's modulus and m is Poisson's ratio. After crack initiation, D
e
is replaced by a reduced
secant stiness matrix, D
f
. In a local coordinate system with axes aligned with the crack normal and crack
tangent, the secant stiness of a cracked element is in [6] given by
D
f
=
0 0 0
0 E 0
0 0
aE
2(1m)
_
_
_
_
; (3)
308 M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330
where a is the so-called shear retention factor (by other authors usually denoted as b), assuming a value
between 0 and 1. In early studies, the shear retention factor was supposed to have a constant nonzero value,
which inevitably leads to stress locking due to shear. Later improvements treated the retention factor as a
function of the crack opening, decaying to zero as the crack opens wide. The rotating crack model com-
pletely discards the shear retention factor because the shear stiness coecient is uniquely dened by the
assumption that the axes of principal stress and strain coincide.
Droz [6] noted that even if the shear retention factor in (3) is set to zero, the element stiness matrix in
general still contains terms that link nodes on opposite sides of the crack. To avoid stress locking, he
suggested to use discontinuous shape functions. The approach of Droz is inspiring but it can be directly
applied only in the `perfectly brittle' case, when the stress transmitted by the crack completely vanishes
immediately after crack initiation. Material laws with gradual softening call for a more rened model.
In 1987, a seminal paper was published by Ortiz et al. [7]. In an eort to improve the resolution of shear
bands by quadrilateral nite elements, they proposed to enrich the strain eld such that a weak disconti-
nuity
2
can be captured. The added strain mode was determined by local bifurcation analysis based on the
acoustic tensor, and its shape was kept constant throughout the analysis (only the magnitude was growing).
Only one weak discontinuity line could cross an element, and so another weak discontinuity line in the
neighboring element was needed to model the entire localization band; see Fig. 1(a). The authors applied
the HellingerReissner principle and presented their technique in the B-bar format.
The idea was improved and further developed by Belytschko et al. [8]. They proposed to embed a lo-
calization zone into the nite element, i.e., an element could contain a band of localized strain bounded by
two parallel weak discontinuity lines; see Fig. 1(b). The width of the band thus became independent of the
element size and could be considered as a material parameter. The direction of the band was also deter-
mined by local bifurcation analysis but, in contrast to [7], the ratio of the opening and sliding component of
the added strain mode was not kept xed. This made the element much more exible and capable of re-
producing nonproportional processes. Fish and Belytschko extended the approach to large deformation
problems [9] and suggested a technique that renes the approximation of a shear band prole in a visco-
plastic solid [10]. Belytschko and coworkers [1114] developed a spectral overlay method, which uses
spectral interpolants superimposed on nite element interpolants. This technique seems to be very powerful
in capturing the shape of the strain eld and its peak value in the localization zone.
A strong discontinuity (embedded localization line) was considered by Dvorkin et al. [15,16]; see
Fig. 1(c). They used the principle of virtual work (PVW) with an added term representing the work of
cohesive tractions on the displacement discontinuity (crack opening). The `parent' elements were MITC
quadrilaterals. A similar modication of the standard bilinear quadrilateral [17] and of the constant-strain
triangle (CST) [1821] was then developed by Klisinski, Oloson, and coworkers. Due to the simplicity of
these elements, no variational principle was needed and the basic equations were derived from physical
considerations.
2
Across a weak discontinuity, certain components of the strain eld have a jump while the displacement eld remains continuous. A
discontinuity in the displacements is classied as strong.
Fig. 1. Elements with (a) one weak discontinuity, (b) two weak discontinuities and (c) one strong discontinuity.
M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330 309
A fully consistent variational derivation of an element with a strong discontinuity (called `embedded
crack' by the authors) was demonstrated by Lot and Shing [2224]. They started from the HuWashizu
principle with an added term representing the work of cohesive tractions and presented their model in the
enhanced assumed strain (EAS) format.
The relationship between softening stressstrain laws for a continuum and stress-separation laws for an
interface was studied in detail by Simo et al. [2527]. They showed that a strong discontinuity can be
described by a continuum-type constitutive law with softening modulus having a distributional character.
Instead of postulating an interface law, it is possible to use the continuum law with a regular function
approximating the Dirac distribution. The authors developed an EAS element with a strong discontinuity
and proposed a nonsymmetric version of the formulation that leads to an improved element performance.
Applications of this technique were presented by Armero and Garikipati [28,29] and by Oliver [30,31]. The
approach was extended to multiplicative large-strain plasticity [32] and to large-strain isotropic damage
[33]. Recently, Oliver et al. [34,35] suggested a model with gradual transition from a weak to a strong
discontinuity.
Regularized displacement discontinuities were also studied by Larsson and Runesson. Initially, the
discontinuities were allowed only at element interfaces [3638], later they were embedded in nite elements
[39,40]. The approach was extended to undrained soils considered as a mixture of a solid skeleton and uid-
lled pores, with discontinuities in displacements and pore pressure [41]. Elements with embedded bands of
localized strain, similar to those of Belytschko et al. [8], were applied to modeling of concrete fracture and
of shear bands in soils by Sluys, Berends and de Borst [4246]. In the most recent implementations, these
authors allowed a rotation of the discontinuity direction, similar to [20,21].
All authors mentioned so far eliminated parameters representing the discontinuity by condensation on
the element level. In contrast to that, Bolzon and Corigliano [49] proposed to model the discontinuity by
additional degrees of freedom treated as global unknowns, similar to the nodal displacements. It is also
worth noting that Li [50] tried to extend the original approach of Droz [6] to strain-softening materials. He
proposed a gradual transition from the standard shape functions to discontinuous ones.
Techniques enriching standard nite element interpolations by strain or displacement discontinuities
were proposed in the literature under dierent names (see Table 1) and derived in dierent ways, e.g., from
simple physical considerations, from the extended principle of virtual work, from the HellingerReissner or
HuWashizu variational principles, using an EAS format or a B-bar format. There is an obvious need for a
comparative study that would present the diverse techniques in a unied framework, point out their
common features and dierences, and nd their limits of applicability. Even though it is hardly possible to
meet all these goals in the present paper, we will attempt a basic evaluation with special attention to
applications in modeling of highly localized damage and fracture.
Table 1
Techniques from the literature
Ortiz et al. [7] Finite element for localized failure
Belytschko et al. [8] Embedded localization zone
Dvorkin et al. [15] Embedded localization line
Klisinski et al. [17] Inner softening band
Lot and Shing [23] Embedded crack
Simo et al. [25] Element with a strong discontinuity
Simo and Oliver [27]
Armero and Garikipati [28]
Oliver [31]
Larsson and Runesson [36] Discontinuous displacement approximation
Larsson et al. [39] Embedded cohesive crack
Larsson and Runesson [40] Embedded localization band
Berends [42] EAS element for fracture
Berends et al. [43] Discontinuous modeling of mode-I failure
Sluys [44] Discontinuous modeling of shear banding
310 M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330
3. Basic framework
3.1. Three-eld variational principle
We start from the general HuWashizu variational principle [51,52], which deals with three independent
elds: the displacement eld, u, the strain eld, e, and the stress eld, r. These elds are dened in a domain
V whose boundary consists of two parts, S
u
and S
t
, with prescribed displacements and tractions, respec-
tively. Except for the kinematic boundary conditions,
u = u on S
u
(4)
and for certain restrictions on regularity, the assumed elds are completely arbitrary and mutually inde-
pendent. All other governing equations can be replaced by the variational equality
_
V
de
T
~ r(e) dV d
_
V
r
T
(u e) dV =
_
V
du
T

bdV
_
S
t
du
T

t dS; (5)
which must hold for any admissible variations du; de, and dr. In (5), the symbol d denotes variation, ~ r(e) is
the stress computed from the assumed strain e using the constitutive equations,
3
is the kinematic op-
erator transforming displacements into strains (the engineering-notation counterpart of the symmetric
gradient operator),

b are the prescribed body forces and

t are the prescribed tractions (surface forces).
Expanding the variation of the second integral, applying the Green theorem to the term containing du,
and taking into account the independence of variations du, de, and dr, we can derive from (5) the strain-
displacement equations
u = e; (6)
constitutive equations
~ r(e) = r; (7)
equilibrium equations

+
r =

b; (8)
and static boundary conditions
nr =

t on S
t
(9)
in which
+
is the adjoint operator of and n is the outward unit normal to the boundary.
3.2. Finite element discretization
Variational statement (5), representing the weak form of Eqs. (6)(9), can be exploited when discretizing
the problem. To this end, we interpolate the unknown elds as
u ~ Nd N
c
d
c
; (10)
e ~ Bd Ge; (11)
r ~ Ss; (12)
where N is the standard displacement interpolation matrix (containing the usual shape functions), B is the
standard strain interpolation matrix (containing the derivatives of the shape functions), N
c
and G are
matrices containing some enrichment terms for displacements and strains, respectively, S is a stress
3
In general, the constitutive operator ~ r depends also on the previous strain history, which may enter through some internal variables.
For simplicity, this dependence is not explicitly marked here.
M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330 311
interpolation matrix, and vectors
4
d, d
c
, e, and s collect the degrees of freedom corresponding to nodal
displacements, enhanced displacement modes, enhanced strain modes, and stress parameters, respectively.
The present interpolation is fairly general and covers some special techniques such as the EAS method [53]
or the B-bar approach [54].
Substituting approximations Eqs. (10)(12) into the variational identity (5) and taking into account that
(Nd) = Bd we obtain
dd
T
_
V
B
T
~ r(Bd Ge) dV de
T
_
V
G
T
~ r(Bd
_
Ge) Ss
_
dV
ds
T
_
V
S
T
(B
c
d
c
Ge) dV dd
T
c
_
V
B
T
c
Ss dV = dd
T
f
ext
dd
T
c
f
c
;
(13)
where B
c
is a strain interpolation matrix that would correspond to the displacement interpolation matrix N
c
(i.e., B
c
is dened by the identity (N
c
d) = B
c
d),
f
ext
=
_
V
N
T

bdV
_
S
t
N
T

t dS (14)
is the vector of (standard) external forces, and
f
c
=
_
V
N
T
c

bdV
_
S
t
N
T
c

t dS (15)
is the vector of nonstandard external forces. For simplicity we will assume that the loads are applied outside
the region with enhanced interpolation, in which case f
c
= 0.
Taking into account the independence of variations, we obtain the discretized equations
_
V
B
T
~ r(Bd Ge) dV = f
ext
; (16)
_
V
G
T
~ r(Bd Ge) dV
_
V
G
T
SdV s = 0; (17)
_
V
S
T
B
c
dV d
c

_
V
S
T
G dV e = 0; (18)
_
V
B
T
c
SdV s = 0: (19)
In order to linearize the dependence of ~ r on parameters d and e we switch to the rate (incremental) for-
mulation and dierentiate (16)(19) with respect to time.
5
For a given state, any incrementally linear
stressstrain equations can formally be written in the rate form
_
~ r = D_ e ~ D(B
_
d G_ e); (20)
where D = o~ r=oe is the tangential stiness matrix of the material. Substituting (20) into (16)(19) we obtain
a set of linear equations
_
V
B
T
DB B
T
DG 0 0
G
T
DB G
T
DG G
T
S 0
0 S
T
G 0 S
T
B
c
0 0 B
T
c
S 0
_

_
_

_
dV
_
d
_ e
_ s
_
d
c
_

_
_

_
=
_
f
ext
0
0
0
_

_
_

_
: (21)
4
In the present context, `vector' is a short name for `column matrix'.
5
Time plays here only the role of a formal parameter controlling the loading process; the model is rate-independent.
312 M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330
The interpolations of stress and strain can be discontinuous, and so we can select the interpolation
functions such that each stress or strain parameter is associated with only one nite element. The same
holds for the enhanced displacement parameters. Parameters e, s and d
c
can therefore be eliminated on the
local (element) level, so that the global equations contain only the standard displacement degrees of
freedom, d.
From now on, we consider Eqs. (16)(19) and (21) written for one nite element occupying a certain
volume V
e
. Of course, the external force vector, f
ext
, is then replaced by the contribution of the current
element to the internal forces, f
e
int
.
3.3. Basic types of enhancement
Let us introduce three basic techniques known from the literature and present them as particular cases or
modications of the general formulation Eqs. (16)(19).
1) Eqs. (18) and (19) are linear already in the total formulation, and so it is natural to inspect them rst.
Suppose that we do not introduce any displacement enhancement terms, i.e., we enrich only the strain
interpolation. Then all terms containing d
c
, N
c
, or B
c
can be deleted from the formulation. Eq. (19) then
does not exist at all (it has been derived by taking the variation of d
c
), and Eq. (18) is reduced to
_
V
e
S
T
G dV e = 0: (22)
In order to pass the generalized patch test, the element must be able to reproduce a constant stress eld
exactly, and so the minimum choice for the stress interpolation matrix is S = I = unit matrix. The com-
patibility conditions (22) now read
_
V
e
G dV e = 0: (23)
The matrix G can always be constructed such that
_
V
e
G dV = 0: (24)
If this condition of zero mean was not satised, we could modify G by subtracting from each entry its mean
value over the element. As the standard strain approximation e = Bd always contains all constant-strain
modes, this modication does not aect the space of functions described by the enhanced approximation
e = Bd Ge.
We can therefore right away assume that the enhanced strain interpolation matrix G satises condition
(24). Then, Eq. (18) is satised for any e. Moreover, as the matrix multiplying s in (17) is now a zero matrix,
the stress parameters s completely disappear from the formulation. We end up with a set of nonlinear
equations
_
V
e
B
T
~ r(Bd Ge) dV = f
e
int
; (25)
_
V
e
G
T
~ r(Bd Ge) dV = 0; (26)
which, at a given state, can be linearized into
_
V
e
B
T
DB B
T
DG
G
T
DB G
T
DG
_ _
dV
_
d
_ e
_ _
=
_
f
int
0
_ _
: (27)
Note that if D is symmetric then the linearized system of Eqs. (27) is also symmetric.
2) Alternatively, we could construct some suitable enhanced displacement interpolation matrix N
c
,
evaluate the corresponding matrix B
c
for which (N
c
d) = B
c
d, and set G = B
c
. Eq. (18) can be satised
M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330 313
(independently of the choice of S) by setting d
c
= e. Eq. (19) now imposes some restrictions on the stress
parameters s but the important point is that by adding (19) to (17) we eliminate the stress parameters and
construct exactly the same equations as before, i.e., (25) and (26). Thus both approaches lead to the same
form of governing equations, with the only dierence that the matrix G either has to satisfy condition (24),
or has to be constructed by applying the strain-displacement operator to a suitable displacement
enhancement.
3) Simo and Oliver [27] proposed a modication of the governing equations that does not follow from a
variational principle; conceptually it corresponds to a weak form with test functions dierent from trial
functions, similar to the PetrovGalerkin method. Their approach represents a compromise between the
preceding two cases because they use G = B
c
in the strain interpolation (11) while replacing G
T
in (26) by a
matrix G
+
that is not the transpose of B
c
but satises the condition of zero mean (24). Later it will be
explained why this choice leads to an improvement of the element performance. The resulting linearized
equations
_
V
e
B
T
DB B
T
DG
G
+
DB G
+
DG
_ _
dV
_
d
_ e
_ _
=
_
f
int
0
_ _
(28)
are in general nonsymmetric even if the material stiness matrix D is symmetric.
In Section 5 we will show that elements derived from the rst abovementioned formulation nicely satisfy
the traction continuity condition but they cannot properly represent the kinematics of a displacement or
strain discontinuity. On the other hand, elements derived from the second formulation nicely reect the
discontinuity but they lead to an awkward approximation of the traction continuity condition. The above
cases 1 and 2 will be referred to as the statically optimal symmetric (SOS) formulation and the kinematically
optimal symmetric (KOS) formulation, respectively. The third approach has the potential of representing
both the kinematic and the static aspects properly, and so it will be called the statically and kinematically
optimal nonsymmetric (SKON) formulation.
3.4. Assumed-strain format
For any of the three formulations introduced in the preceding subsection, the governing equations can
be written as
K
bb
K
bg
K
gb
K
gg
_ _
_
d
_ e
_ _
=
_
f
int
0
_ _
; (29)
where
K
bb
=
_
V
e
B
T
DBdV ; (30)
K
bg
=
_
V
e
B
T
DG dV ; (31)
K
gb
=
_
V
e
G
+
DBdV ; (32)
K
gg
=
_
V
e
G
+
DG dV : (33)
For the SOS formulation, G is a matrix with zero mean over the element and G
+
= G
T
. For the KOS
formulation, G = B
c
and G
+
= B
T
c
. Finally, for the SKON formulation, G = B
c
and G
+
is a matrix with
zero mean.
314 M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330
Assuming that K
gg
is regular, we can eliminate _ e and write
_ e = K
1
gg
K
gb
_
d; (34)
_
f
int
= K
bb
_
d K
bg
K
1
gg
K
gb
_
d = K
_
d; (35)
where
K = K
bb
K
bg
K
1
gg
K
gb
(36)
is the condensed element stiness matrix to be used in the assembly process. This is the usual procedure
exploited by EAS methods; see [53].
3.5. B-bar format
The symmetric formulations (SOS and KOS) can easily be cast into the B-bar format [54]. Substituting
(34) into the rate form of the strain approximation (11) we obtain
_ e ~ B
_
d G_ e = B
_
d GK
1
gg
K
gb
_
d = (B GK
1
gg
K
gb
)
_
d =

B
_
d; (37)
where

B = B GK
1
gg
K
gb
(38)
is the so-called B-bar matrix. A straightforward calculation shows that the element stiness matrix (36) can
alternatively be dened as
K =
_
V
e

B
T
D

BdA (39)
and that
_
V
e

B
T
_
~ rdA =
_
V
e
B
T
_
~ rdA (40)
so that the equivalent internal forces
f
e
int
=
_
V
e
B
T
~ rdA (41)
that (after assembly) have to be in equilibrium with the external forces can alternatively be computed as
f
e
int
=
_
t
0
_
V
e

B
T
_
~ rdAds: (42)
Note that the B-bar matrix dened by (38) changes throughout the calculation (because it depends on
the current tangential stiness of the material) while the standard B-matrix remains constant.
4. Particular formulations
4.1. Localization band in two dimensions
For simplicity, we focus on a plane problem, discretized by triangular or quadrilateral elements. Suppose
that a certain element, occupying an area A
e
, is crossed by a localization band
6
of a constant width k. The
element can be divided into a region of localized strain L and its complement N, which usually consists of
6
By a localization band we mean a region of localized strain, bounded by two parallel weak discontinuities; see also Fig. 1(b).
M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330 315
two disjoint regions, N

and N

; see Fig. 2(a). When dealing with a single element, we can work in a local
coordinate system (x; y) aligned with the localization band to be captured. Axis x is assumed to be normal
to the band and axis y parallel to it. If A
L
is the area of the localization band, we can dene an equivalent
length of the band l = A
L
=k; see Fig. 2(b). It is also useful to dene an equivalent `element width' (char-
acteristic size) h = A
e
=l (Fig. 2(c)), so that A
L
=A
e
= kl=hl = k=h. The area of the nonlocalized region is
A
N
= A
e
A
L
= (h k)l.
4.2. Statically optimal symmetric formulation
4.2.1. Element with a localization band
The approach of Belytschko et al. [8] is recovered from the SOS formulation if we use a strain en-
richment that allows a constant jump in strains e
x
and c on the boundaries of the localization band. This can
be achieved by setting
G =
1
kA
e
(A
N
v
L
A
L
v
N
)P; (43)
where
P =
1 0
0 0
0 1
_
_
_
_
(44)
and v
L
and v
N
are the characteristic functions
7
of the localization band and of its complement, respectively.
Matrix G is thus piecewise constant and can be presented as
G = P
L
v
L
P
N
v
N
; (45)
here
P
L
=
A
N
kA
e
P; P
N
=
A
L
kA
e
P: (46)
The factors at v
L
and v
N
in (43) are chosen such that G satises the condition of zero mean (24). Indeed,
_
A
e
(A
N
v
L
A
L
v
N
) dA = A
N
A
L
A
L
A
N
= 0 (47)
and so
_
A
e
G dA = 0. The common scaling factor 1=kA
e
is chosen such that the components of e have the
meaning of normal and shear strain dierence between the localized and nonlocalized region multiplied by
the width of the band. In the limit for a localization band width approaching zero, they represent the
opening and sliding component of a strong (displacement) discontinuity.
Fig. 2. Element with a localization band.
7
The characteristic function v
M
of a subdomain M is equal to 1 inside M and equal to 0 everywhere else.
316 M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330
According to (25), the internal forces are evaluated as
f
e
int
=
_
A
e
B
T
~ rdA =
_
A
L
B
T
r
L
dA
_
A
N
B
T
r
N
dA; (48)
where r
L
is the stress in the localization band, computed from the strain
e
L
= Bd P
L
e = Bd
A
N
kA
e
Pe (49)
and r
N
is the stress outside the localization band, computed from the strain
e
N
= Bd P
N
e = Bd
A
L
kA
e
Pe: (50)
The stresses must satisfy the internal equilibrium condition (26), which for the matrix G dened by (43)
leads to
A
N
kA
e
_
A
L
P
T
r
L
dA
A
L
kA
e
_
A
N
P
T
r
N
dA = 0: (51)
After a simple rearrangement we obtain
1
A
L
_
A
L
P
T
r
L
dA =
1
A
N
_
A
N
P
T
r
N
dA: (52)
Note that the multiplication by P
T
selects the rst and third component from r
L
or r
N
, i.e., the normal
stress r
x
and the shear stress s
xy
. Eq. (52) can therefore be interpreted as a weak stress continuity condition,
stating that r
x
and s
xy
averaged over the localization band must be the same as r
x
and s
xy
averaged over the
nonlocalized region.
4.2.2. Element with a discontinuity line
Let us now consider the limit case when the thickness of the localization band tends to zero and the band
collapses into a curve S
L
. Denition (43) of the enhanced strain interpolation matrix can be rewritten as
G =
1
khl
(h [ k)lv
L
klv
N
[P =
h k
hk
v
L
_

1
h
v
N
_
P: (53)
As k 0, the term v
L
=k tends to the Dirac distribution d
L
, and G tends to
G
0
= d
L
_

1
h
_
P: (54)
The contribution of the localization band to the internal forces vanishes and (48) reads
f
e
int
=
_
A
e
B
T
rdA; (55)
where we write r instead of r
N
because the nonlocalized region N now extends over the entire element with
the exclusion of the discontinuity curve, S
L
. For the same reason, we will drop the subscript at e
N
.
In Eq. (52), the product P
T
r
L
has to be replaced by the cohesive tractions t, and the resulting internal
equilibrium condition
1
l
_
S
L
t dS =
1
A
e
_
A
e
P
T
rdA (56)
means that the tractions averaged over the discontinuity line S
L
must be equal to the rst and third
component (r
x
and s) of the stress averaged over the bulk of the element. This is the weak form of the
condition of traction continuity across the localization line.
M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330 317
4.2.3. Constant-strain triangle
To gain further insight into the structure of the discretized equations, consider the simplest element a
CST with a nite localization band. Strains e
L
inside the localization band are constant, and strains e
N
outside the band are also constant (but in general dierent from e
L
). Consequently, the stresses are also
piecewise constant and the internal forces can be evaluated as
f
e
int
= B
T
(A
L
r
L
A
N
r
N
): (57)
The internal equilibrium condition (52) for a CST element reads
P
T
r
L
= P
T
r
N
: (58)
Due to the fact that the stresses are piecewise constant, the traction continuity condition on the boundary
between L and N is enforced in the strong sense.
The structure of the governing equations is depicted in the diagram in Fig. 3(a). The top part corre-
sponds to the equilibrium equations, the middle part to the constitutive relations, and the bottom part to
the kinematic equations. Dashed arrows indicate that the source is added to the target. For example, the
strain e
N
is given by the sum of of Bd and P
N
e. Symbol ~ r denotes the constitutive operator, i.e., r
N
= ~ r(e
N
)
and r
L
= ~ r(e
L
).
In the limit case for k 0 we obtain a CST element with an embedded displacement discontinuity line,
described by the equations schematically shown in Fig. 3(b). Symbol ~s denotes the constitutive operator for
the interface. Solid arrows mean that the source must be equal to the target. For example, the tractions t are
not the sum of P
T
r and ~s(e) but they are equal to either of these expressions, i.e., t = P
T
r = ~s(e).
Fig. 3. Structure of the equations describing the SOS formulation (a) with a localization band and (b) with a localization line.
318 M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330
4.3. Kinematically optimal symmetric formulation
4.3.1. Element with a localization band
The approach of Lot and Shing [2224] is recovered from the KOS formulation if we use an enhanced
displacement interpolation matrix
N
c
=
N
c
0
0 N
c
_ _
(59)
with a special shape function
N
c
(x) = H
L
(x)

N
i=1
H
L
(x
i
)N
i
(x); (60)
where N is the number of nodes per element; N
i
, i = 1; 2; . . . ; N, are the standard shape functions; x
i
,
i = 1; 2; . . . ; N, are the nodal coordinate vectors; and H
L
is a ramp function assuming a constant value
H
L
= 0 in N

and a constant value H


L
= 1 in N

, with a linear transition from 0 to 1 in L. Function H


L
supplies the actual enrichment of the displacement interpolation while the corrective term

H
L
(x
i
)N
i
(x) is
only a linear combination of the standard shape functions. This correction makes sure that the value of the
nonstandard shape function N
c
at every node is zero, so that the standard degrees of freedom d keep the
meaning of nodal displacements. Indeed, taking into account that N
i
(x
j
) = d
ij
= Kronecker delta, we get
N
c
(x
j
) = H
L
(x
j
)

N
i=1
H
L
(x
i
)N
i
(x
j
) = H
L
(x
j
)

N
i=1
H
L
(x
i
)d
ij
= 0: (61)
Applying the strain-displacement operator to the enhanced displacement interpolation we obtain the
enhanced strain interpolation matrix
B
c
=
H
L;x
0
0 H
L;y
H
L;y
H
L;x
_
_
_
_

N
i=1
H
L
(x
i
)
N
i;x
0
0 N
i;y
N
i;y
N
i;x
_
_
_
_
: (62)
Subscript after a comma indicates partial dierentiation, e.g., H
L;x
= oH
L
=ox. If the coordinate system is
aligned with the localization band (see Fig. 2(a)), we have H
L;x
= v
L
=k and H
L;y
= 0. For isoparametric nite
elements, each of the matrices
B
i
=
N
i;x
0
0 N
i;y
N
i;y
N
i;x
_
_
_
_
: (63)
is a certain submatrix of the standard strain interpolation matrix
B = B
1
B
2
. . . B
N
[ [: (64)
Therefore, the enhanced strain interpolation matrix from (62) can be presented in a compact form
B
c
=
1
k
v
L
P BH
L
; (65)
where P is the Boolean matrix dened by (44) and
H
L
=
H
L
(x
1
) 0
0 H
L
(x
1
)
.
.
.
.
.
.
H
L
(x
N
) 0
0 H
L
(x
N
)
_

_
_

_
: (66)
M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330 319
Formula (48) for the internal forces remains valid but this time the stress r
L
is computed from the strain
e
L
= Bd
1
k
P
_
BH
L
_
e = B(d H
L
e)
1
k
Pe = Bd
N

1
k
Pe (67)
and the stress r
N
is computed from the strain
e
N
= Bd BH
L
e = B(d H
L
e) = Bd
N
; (68)
where
d
N
= d H
L
e: (69)
For the present choice G = B
c
, the internal equilibrium condition (26) reads
_
A
L
1
k
P
_
BH
L
_
T
r
L
dA
_
A
N
BH
L
( )
T
r
N
dA = 0; (70)
which can be transformed into
H
T
L
_
A
L
B
T
r
L
dA
_

_
A
N
B
T
r
N
dA
_
=
1
k
_
A
L
P
T
r
L
dA: (71)
The expression in parentheses on the left-hand side is according to (48) equal to the vector of internal
forces. Multiplication by H
T
L
is equivalent to summing the internal forces corresponding to the nodes that
are located in N

and adding a weighted contribution of the nodes that happen to be in the localization
band L. Thus, rewriting (71) as
H
T
L
f
e
int
=
l
A
L
_
A
L
P
T
r
L
dA (72)
we obtain a condition with the following interpretation: the average values of stress components r
x
and s
xy
in the localization band multiplied by the length of the band must be equal to the sum of internal forces
acting on the `positive' side of the band.
4.3.2. Element with a discontinuity line
In the limit for k 0, the internal forces are evaluated according to (55), and the weak stress continuity
condition (72) reads
H
T
L
f
e
int
=
_
S
L
t dS: (73)
This means that the sum of internal forces acting on the `positive' side of the band must be equal to the
cohesive tractions integrated along the discontinuity line. In other words, if we cut the element along the
discontinuity line and replace the interaction between the separated parts by the cohesive tractions, either of
the separated parts of the element must satisfy the conditions of force equilibrium (but not necessarily the
condition of moment equilibrium); see Fig. 4.
4.3.3. Constant-strain triangle
The structure of the governing equations for a CST element based on the KOS formulation with either a
nite localization band or a discontinuity line is shown in the diagrams in Fig. 5. Symbol I denotes the
identity operator, i.e., d
N
is equal to the sum of Id = d and H
L
e.
4.4. Statically and kinematically optimal nonsymmetric formulation
The approach of Klisinski et al. [17,18] is recovered from the SKON formulation if we use the enhanced
strain interpolation matrix B
c
given by (65) but in the internal equilibrium condition (26) we use the
320 M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330
transpose of the matrix G from (43). Consequently, the stresses r
L
and r
N
are calculated from the strains e
L
and e
N
given by (67) and (68), same as for the elements due to Lot and Shing, but the stress continuity
condition (52) or (56) is taken from the approach due to Belytschko et al.
The SKON formulation of a CST with a discontinuity line leads to the element constructed by Olofsson
et al. [18] based on simple physical considerations. They decomposed the nodal displacements into a part
due to uniform strain in the bulk of the element and a part due to crack opening. In our notation, these
parts correspond to d
N
and H
L
e, respectively. Olofsson et al. related the strains e = Bd
N
to stresses r using a
linear elastic law, and the crack opening e to the cohesive tractions t using a plasticity-type interface model.
Finally, a natural traction continuity requirement t = P
T
r was imposed, and the internal forces were
calculated as f
e
int
= A
e
B
T
r. The structure of the governing equations for this element is shown in Fig. 6(b).
Fig. 6(a) presents a similar element with a nite localization band.
The earliest implementation of an element combining the advantages of the SOS and KOS formulations
was presented by Dvorkin et al. [15] for quadrilaterals. A general version of the SKON formulation for any
Fig. 5. Structure of the equations describing the KOS formulation (a) with a localization band and (b) with a localization line.
Fig. 4. Equilibrium of an element split by a discontinuity line.
M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330 321
type of parent element was proposed by Simo and Oliver [27]. For the CST their approach gives exactly the
same element as the one due to Olofsson et al. [18]. The general technique was then applied by Armero and
Garikipati [28,32] to mixed triangular elements.
4.5. Derivation from the PVW
Additional insight is provided by the derivation of the symmetric formulations from the PVW. When
looking at a single element, we may consider the forces f
e
int
as external ones and express the external virtual
work as
dW
ext
= f
e T
int
dd: (74)
For a model with a strong discontinuity, the internal virtual work is done by the stresses in the continuous
part on the corresponding virtual strains and by the tractions transmitted by the discontinuity on the virtual
opening and sliding components of the displacement jump. Hence we may write
dW
int
=
_
A
e
r
T
de dA
_
S
L
t
T
dedS: (75)
According to the PVW, the internal virtual work must be equal to the external one for any virtual change
of the kinematic state. By denition, the virtual change must be kinematically admissible. If we postulate a
certain kinematic assumption, e.g.,
e = B(d H
L
e) (76)
then the virtual change is uniquely described by dd and de, which may be regarded as independent pa-
rameters that determine the virtual strain
de = B(dd H
L
de): (77)
Fig. 6. Structure of the equations describing the SKON formulation (a) with a localization band and (b) with a localization line.
322 M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330
Substituting (77) into (75) and setting the resulting expression equal to (74) we obtain
_
A
e
r
T
B(dd H
L
de) dA
_
S
L
t
T
dedS = f
e T
int
dd: (78)
This equality holds for arbitrary dd and de if and only if
_
A
e
B
T
rdA = f
e
int
(79)
and

_
A
e
H
T
L
B
T
rdA
_
S
L
t dS = 0: (80)
The rst condition is the standard formula for the evaluation of equivalent nodal forces while the second
one can be rewritten as
_
S
L
t dS = H
T
L
f
e
int
: (81)
This is condition (73) that enforces (in the weak sense) equilibrium between the tractions and the nodal
forces acting on the solitary
8
node. It was derived from the requirement that, if the strains in the con-
tinuous part of the element are kept constant and the displacement discontinuity is subjected to a virtual
change, the virtual work done by the nodal forces must be equal to the virtual work done by the tractions
across the discontinuity. So a derivation based on the PVW that starts from a `natural' kinematic as-
sumption (76) automatically leads to static condition (81) that is work-conjugate with the kinematic one but
diers from the `natural' condition of traction continuity.
In an analogous manner, it is possible to proceed from a static assumption to the work-conjugate ki-
nematic condition derived from the complementary PVW, which states that the work done by virtual nodal
forces on the actual displacements,
dW
+
ext
= df
e T
int
d (82)
must be equal to the work done by virtual stresses on the actual strains and by virtual tractions on the
actual displacement discontinuity,
dW
+
int
=
_
A
e
dr
T
e dA
_
S
L
dt
T
edS (83)
for an arbitrary virtual state that is statically admissible. The postulated static assumptions consist of the
standard relation between stresses and nodal forces,
_
A
e
B
T
rdA = f
e
int
(84)
and a condition that links the stresses and tractions across the discontinuity. It is natural to require that the
tractions be equal to the projected stress components,
t = P
T
r: (85)
Relations analogous to (84) and (85) must also hold for any virtual change of the static state. We might
therefore regard the virtual stress dr as independent and express the virtual nodal forces and virtual
tractions as df
e
int
=
_
A
e
B
T
drdA and dt = P
T
dr, respectively. Substituting this into (82) and (83) and ap-
plying the complementary PVW we obtain
8
The solitary node is the one separated from the other nodes (of the same element) by the discontinuity line.
M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330 323
_
A
e
dr
T
e dA
_
S
L
dr
T
PedS =
_
A
e
dr
T
BdAd: (86)
The virtual stress eld dr is not completely arbitrary but has to be self-equilibrated inside A
e
. This
condition holds in particular for constant stress elds. In this case, we might take dr out of the integrals,
and by standard arguments we arrive at
_
A
e
e dA
_
S
L
PedS =
_
A
e
Bd dA: (87)
For the special case of a CST, this may be further simplied to
e = Bd
l
A
e
Pe; (88)
which is the kinematic condition work-conjugate with the `natural' condition of traction continuity;
cf. Eq. (50) with e
N
e and A
L
=k l.
5. Discussion
Having demonstrated the three fundamental formulations, we can classify the models from the literature
that were briey described in the historical overview. There are many aspects in which individual models
dier, e.g., the type of parent element, type of discontinuity (weak/strong), constitutive law, etc.; see Table 2.
A possible classication is suggested by the diagrams in Figs. 3, 5 and 6. The top part of each diagram
corresponds to the equilibrium equations, the middle part to constitutive equations, and the bottom part to
kinematic equations. We will not pay too much attention to the constitutive aspects, even though they are
certainly an important ingredient of the model. The principal aspect studied here is the type of kinematic
enhancement and of the internal equilibrium condition.
In the model of Ortiz et al. [7], only one weak discontinuity line could cross an element, and so the width
of the localization band was not independent of the element size. Belytschko et al. [8] allowed two parallel
weak discontinuity lines in a single element, so that the element could contain a band of localized strain.
They developed the rst element with an embedded localization band, based on what we now call the SOS
formulation. Recently, this formulation has been applied to the CST by a number of authors; see Table 2.
The only KOS formulation published in the literature seems to be the one due to Lot and Shing [24],
based on Lot's dissertation [22]. The lack of popularity of this approach can be attributed to the cum-
bersome form of the internal equilibrium condition (81); see the discussion in Section 5.2.
To the best of the author's knowledge, the earliest implementation of an element combining the ad-
vantages of the SOS and KOS formulations was presented by Dvorkin et al. [15] for standard bilinear
quadrilaterals (Q4) and for quadrilaterals with a mixed interpolation of tensorial components (QMITC).
However, the true nature of that model was somewhat obscured by its complicated derivation. A very
similar element was constructed by Klisinski et al. [17] based on simple and instructive physical consid-
erations. In a later paper [18], the same technique was applied to a CST.
A general version of the SKON formulation for an arbitrary type of parent element was proposed in a
short paper by Simo and Oliver [27] and fully described by Oliver [31]. For the CST their approach gives
exactly the same kinematic and static equations as those due to Oloson et al. [18].
5.1. Kinematic equations
The dierences between the symmetric formulations are best explained by scrutinizing a CST with a
strong discontinuity line. For simplicity we assume that the continuous material is linear elastic and all
nonlinear processes take place in the discontinuity line. The standard displacement interpolation would
lead to a constant strain e = Bd. Due to the presence of a discontinuity, a part of the strain is relaxed by
crack opening and sliding. The SOS formulation starts from the standard strain and partially relaxes the
324 M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330
T
a
b
l
e
2
O
v
e
r
v
i
e
w
o
f
f
o
r
m
u
l
a
t
i
o
n
s
F
o
r
m
u
-
l
a
t
i
o
n
A
u
t
h
o
r
s
P
a
r
e
n
t
e
l
e
m
e
n
t
s
D
i
s
c
o
n
t
i
n
u
i
t
y
M
a
t
e
r
i
a
l
l
a
w
A
p
p
l
i
c
a
t
i
o
n
e
x
a
m
p
l
e
s
S
O
S
B
e
l
y
t
s
c
h
k
o
e
t
a
l
.
[
8
]
Q
u
a
d
r
i
l
a
t
e
r
a
l
s
W
e
a
k
J
2
-
p
l
a
s
t
i
c
i
t
y
S
h
e
a
r
b
a
n
d
s
L
a
r
s
s
o
n
a
n
d
R
u
n
e
s
s
o
n
[
4
0
]
C
S
T
R
e
g
u
l
a
r
i
z
e
d
s
t
r
o
n
g
P
l
a
s
t
i
c
i
t
y
M
i
x
e
d
-
m
o
d
e
f
r
a
c
t
u
r
e
L
a
r
s
s
o
n
e
t
a
l
.
[
4
1
]
C
S
T
R
e
g
u
l
a
r
i
z
e
d
s
t
r
o
n
g
P
l
a
s
t
i
c
i
t
y
S
h
e
a
r
b
a
n
d
s
i
n
s
o
i
l
s
B
e
r
e
n
d
s
[
4
2
]
C
S
T
W
e
a
k
S
m
e
a
r
e
d
c
r
a
c
k
C
r
a
c
k
i
n
g
A
r
m
e
r
o
[
2
9
]
C
S
T
S
t
r
o
n
g
D
a
m
a
g
e
C
r
a
c
k
i
n
g
o
f
c
o
n
c
r
e
t
e
S
l
u
y
s
[
4
4
]
C
S
T
W
e
a
k
P
l
a
s
t
i
c
i
t
y
S
h
e
a
r
b
a
n
d
s
S
l
u
y
s
a
n
d
B
e
r
e
n
d
s
[
4
5
]
C
S
T
R
o
t
a
t
i
n
g
w
e
a
k
S
m
e
a
r
e
d
c
r
a
c
k
,
p
l
a
s
t
i
c
i
t
y
C
r
a
c
k
i
n
g
K
O
S
L
o
t

[
2
2
]
,
L
o
t

a
n
d
S
h
i
n
g
[
2
4
]
C
S
T
,
Q
4
,
Q
E
5
,
Q
M
6
S
t
r
o
n
g
C
o
h
e
s
i
v
e
c
r
a
c
k
C
r
a
c
k
i
n
g
o
f
c
o
n
c
r
e
t
e
S
K
O
N
D
v
o
r
k
i
n
e
t
a
l
.
[
1
5
]
Q
4
,
Q
M
I
T
C
S
t
r
o
n
g
C
o
h
e
s
i
v
e
c
r
a
c
k
C
r
a
c
k
i
n
g
o
f
c
o
n
c
r
e
t
e
K
l
i
s
i
n
s
k
i
e
t
a
l
.
[
1
7
]
Q
4
S
t
r
o
n
g
C
o
h
e
s
i
v
e
c
r
a
c
k
C
r
a
c
k
i
n
g
S
i
m
o
a
n
d
O
l
i
v
e
r
[
2
7
]
C
S
T
R
e
g
u
l
a
r
i
z
e
d
s
t
r
o
n
g
D
a
m
a
g
e
,
p
l
a
s
t
i
c
i
t
y
S
h
e
a
r
b
a
n
d
s
O
l
o

s
o
n
e
t
a
l
.
[
1
8
]
C
S
T
S
t
r
o
n
g
P
l
a
s
t
i
c
i
n
t
e
r
f
a
c
e
C
r
a
c
k
i
n
g
A
r
m
e
r
o
a
n
d
G
a
r
i
k
i
p
a
t
i
[
2
8
]
M
i
x
e
d
t
r
i
a
n
g
l
e
s
S
t
r
o
n
g
P
l
a
s
t
i
c
i
t
y
S
h
e
a
r
b
a
n
d
s
i
n
m
e
t
a
l
s
A
r
m
e
r
o
a
n
d
G
a
r
i
k
i
p
a
t
i
[
3
2
]
M
i
x
e
d
t
r
i
a
n
g
l
e
S
t
r
o
n
g
L
a
r
g
e
s
t
r
a
i
n
p
l
a
s
t
i
c
i
t
y
S
h
e
a
r
b
a
n
d
s
i
n
m
e
t
a
l
s
O
l
i
v
e
r
[
3
1
]
C
S
T
R
e
g
u
l
a
r
i
z
e
d
s
t
r
o
n
g
D
a
m
a
g
e
,
p
l
a
s
t
i
c
i
t
y
C
r
a
c
k
i
n
g
,
s
h
e
a
r
b
a
n
d
s
O
h
l
s
s
o
n
a
n
d
O
l
o

s
o
n
[
1
9
]
C
S
T
S
t
r
o
n
g
P
l
a
s
t
i
c
i
n
t
e
r
f
a
c
e
M
i
x
e
d
-
m
o
d
e
f
r
a
c
t
u
r
e
O
l
i
v
e
r
e
t
a
l
.
[
3
5
]
C
S
T
W
e
a
k

s
t
r
o
n
g
P
l
a
s
t
i
c
i
t
y
S
h
e
a
r
b
a
n
d
s
i
n
s
o
i
l
s
O
l
i
v
e
r
a
n
d
P
u
l
i
d
o
[
3
3
]
C
S
T
R
e
g
u
l
a
r
i
z
e
d
s
t
r
o
n
g
L
a
r
g
e
s
t
r
a
i
n
d
a
m
a
g
e
S
h
e
a
r
b
a
n
d
s
,
n
e
c
k
i
n
g
T
a
n
o
e
t
a
l
.
[
2
1
]
C
S
T
R
o
t
a
t
i
n
g
s
t
r
o
n
g
P
l
a
s
t
i
c
i
n
t
e
r
f
a
c
e
M
i
x
e
d
-
m
o
d
e
f
r
a
c
t
u
r
e
M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330 325
strain component normal to the discontinuity line and the shear component, while the KOS formulation rst
subtracts the contribution of the displacement jump from the nodal displacements and only then applies the
kinematic operator. An important nding is that the former approach in general cannot reproduce full stress
relaxation around a widely open crack. To explain that, consider a displacement eld given by u = 0 on one
side of the discontinuity line (in N

) and by u = e = e
n
; e
s

T
on the other side of the discontinuity line (in
N

). Suppose that a certain element is crossed by the discontinuity such that nodes 1 and 2 are in N

and
node 3 is in N

; see Fig. 7(a) and (b). The SOS formulation leads to the following strain interpolation:
e = Bd
1
h
Pe =
1
2A
e
(y
1
y
2
)e
n
(x
2
x
1
)e
s
(x
2
x
1
)e
n
(y
1
y
2
)e
s
_

_
_

l
A
e
e
n
0
e
s
_

_
_

_
=
1
2A
e
(y
1
y
2
2l)e
n
(x
2
x
1
)e
s
(x
2
x
1
)e
n
(y
1
y
2
2l)e
s
_

_
_

_
: (89)
It is obvious that all strain components vanish only if x
1
= x
2
and y
1
y
2
= 2l, i.e., if side 12 is parallel to
the discontinuity line and twice as long as the intersection of the discontinuity line with the element under
consideration. In all other situations, some of the strain components are nonzero, which must lead to a
spurious stress transfer.
In contrast to the locking SOS formulation, the KOS formulation subtracts the crack opening com-
ponents e
n
and e
s
from the nodal displacements u
3
and v
3
, which leads to zero nodal displacements
Fig. 7. Element crossed by a stress-free crack.
326 M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330
d
N
= d H
L
e and, consequently, to zero strains e = Bd
N
in the continuous part of the material. Spurious
stress transfer is thus avoided.
To gain further insight, we rewrite the kinematic equations of the KOS formulation as
e = B(d H
L
e) = Bd BH
L
e: (90)
If node 3 is solitary, as in Fig. 7, we have
H
L
=
0 0 0 0 1 0
0 0 0 0 0 1
_ _
T
(91)
and
BH
L
= B
3
=
1
2A
e
y
1
y
2
0
0 x
2
x
1
x
2
x
1
y
1
y
2
_

_
_

_ =
l
12
2A
e
y
1
y
2
l
12
0
0
x
2
x
1
l
12
x
2
x
1
l
12
y
1
y
2
l
12
_

_
_

_
=
1
h
12
cos a 0
0 sin a
sin a cos a
_

_
_

_; (92)
where l
12
=

(x
2
x
1
)
2
(y
1
y
2
)
2
_
is the length of the side connecting nodes 1 and 2 (opposite to the
solitary node), h
12
= 2A
e
=l
12
is the distance of the solitary node from this side, and cos a = (y
1
y
2
)=l
12
and
sin a = (x
2
x
1
)=l
12
are the components of the unit vector normal to this side, a being the angle between the
side and the discontinuity line; see Fig. 7(c).
If the side opposite to the solitary node is parallel to the discontinuity line, then h
12
= h, cos a = 1,
sin a = 0, and
BH
L
=
1
h
1 0
0 0
0 1
_
_
_
_
=
1
h
P: (93)
This means that in this particular case the kinematic equations of the KOS and SOS formulations are
equivalent. In a general case, the product BH
L
has the form (93) in a coordinate system (x; y) aligned with
the side opposite to the solitary node; see Fig. 7(c). Of course, in such a coordinate system the components
of e do not have the meaning of opening and sliding but they can be easily transformed into these physical
components by a simple rotation. The important point is that the KOS formulation relaxes strain com-
ponents e
xx
and c
x y
, instead of the components e
xx
and c
xy
relaxed by the SOS formulation. Since e
y y
is not
aected by the relative displacement of the solitary node with respect to the opposite side, a full strain
relaxation is possible in the KOS formulation, and locking does not occur.
5.2. Static equations
According to the preceding analysis, the KOS formulation is clearly superior from the kinematic point of
view. Let us now look at the static aspects, especially at the internal equilibrium condition. For the CST, the
SOS formulation leads to the natural traction continuity condition (85) on the boundary of the localization
band. On the other hand, the KOS formulation leads to the condition that the tractions along the dis-
continuity be equal to the nodal force on the solitary node divided by the length of the band; cf. Eq. (81).
The natural traction continuity condition can be severely violated. Consequently, the criteria for the onset
of localization written in terms of stresses in the bulk are no longer equivalent to the same criteria written in
terms of the tractions on the discontinuity line. For example, if we use the Rankine criterion in the stress
space, we would certainly like the normal traction at the onset of localization to be equal to the tensile
strength and the shear traction to be zero. This naturally follows from the continuity condition but is not
properly reproduced by the `smeared nodal forces' corresponding to the KOS formulation. Suppose that
the body is in a state of stress characterized by r
x
= f
t
= tensile strength, r
y
< f
t
, s
xy
= 0. Nodal forces at
the solitary node 3 are calculated as
M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330 327
F
3x
= A
e
1
2A
e
(y
1
_
y
2
)r
x
(x
1
x
2
)s
xy

=
y
1
y
2
2
f
t
; (94)
F
3y
= A
e
1
2A
e
(y
1
_
y
2
)s
xy
(x
1
x
2
)r
y

=
x
2
x
1
2
r
y
(95)
and the smearing procedure leads to ctitious tractions
t
n
=
F
3x
l
=
y
1
y
2
2l
f
t
; (96)
t
s
=
F
3y
l
=
x
2
x
1
2l
r
y
; (97)
that are in general dierent from the actual tractions, t
n
= f
t
and t
s
= 0. The natural traction continuity
condition is recovered only in the special case when l = (y
1
y
2
)=2 and x
1
= x
2
. These are once again the
conditions that were necessary for the SOS formulation to avoid locking.
6. Conclusion
A number of techniques enriching the standard nite element interpolation by additional terms corre-
sponding to a displacement or strain discontinuity have been presented within a unied framework and
critically evaluated. It has been shown that there exist three major classes of these models, called here SOS,
KOS and SKON. The SOS formulation cannot properly reect the kinematics of a completely open crack
but it gives a natural traction continuity condition, while the KOS formulation describes the kinematic
aspects satisfactorily but it leads to an awkward relationship between the stress in the bulk of the element
and the tractions across the discontinuity line.
These properties of the symmetric formulations were the driving force behind the development of the
nonsymmetric SKON formulation, which combines the optimal static and kinematic equations and leads to
an improved numerical performance. This formulation deals with a very natural traction continuity con-
dition and is capable of properly representing complete separation at late stages of the fracturing process,
without any locking eects (spurious stress transfer). The price to pay is the loss of symmetry of the
tangential stiness matrix. Results reported in the literature show that the nonsymmetric model can be used
with success in numerical simulations of localized cracking and shear banding. It is also worth noting that
the SKON formulation does not require any specication of the `length' of the localization band. This is an
important advantage because such length is in general not an objective quantity and its value depends on
the (partially ambiguous) rule for the positioning of the discontinuity inside the element.
Numerical implementation of nite elements with a localization line (embedded crack) based on the
SKON formulation and on a damage-type traction-separation law for the discontinuity shall be presented
in a separate paper [48].
Acknowledgements
Financial support of the Swiss Committee for Technology and Innovation under project CTI.3201.1 is
gratefully acknowledged. Thanks are also due to the director of the project, Dr. Thomas Zimmermann, and
to Prof. Javier Oliver of the Technical University of Catalonia for stimulating discussions.
References
[1] J.G. Rots, Computational Modeling of Concrete Fracture, Ph.D. thesis, Delft University of Technology, Delft, The Netherlands,
1988.
[2] M. Jirasek, Th. Zimmermann, Analysis of rotating crack model, J. Eng. Mech. ASCE 124 (1998) 842851.
[3] M. Jirasek, Th. Zimmermann, Rotating crack model with transition to scalar damage, J. Eng. Mech. ASCE 124 (1998) 277284.
328 M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330
[4] M. Jirasek, Th. Zimmermann, Rotating crack model with transition to scalar damage: I. Local formulation, II. Nonlocal
formulation and adaptivity, LSC Internal Report, 97/01, January 1997.
[5] C. Johnson, R. Scott, A nite element method for problems in perfect plasticity using discontinuous trial functions, in:
W. Wunderlich, E. Stein, K.-J. Bathe, (Eds.), Nonlinear Finite Element Analysis in Structural Mechanics, Springer, Berlin, 1981,
pp. 307324.
[6] P. Droz, Modele numerique du comportement non-lineaire d'ouvrages massifs en beton non arme, Ph.D. thesis No. 682, Swiss
Federal Institute of Technology at Lausanne (EPFL), 1987.
[7] M. Ortiz, Y. Leroy, A. Needleman, A nite element method for localized failure analysis, Comput. Meths. Appl. Mech. Eng. 61
(1987) 189214.
[8] T. Belytschko, J. Fish, B.E. Engelmann, A nite element with embedded localization zones, Comput. Meths. Appl. Mech. Eng. 70
(1988) 5989.
[9] J. Fish, T. Belytschko, Elements with embedded localization zones for large deformation problems, Comput. Struct. 30 (1988)
247256.
[10] J. Fish, T. Belytschko, A nite element with a unidirectionally enriched strain eld for localization analysis, Comput. Meths. Appl.
Mech. Eng. 78 (1990) 181200.
[11] T. Belytschko, J. Fish, A. Bayliss, The spectral overlay on nite elements for problems with high gradients, Comput. Meths. Appl.
Mech. Eng. 81 (1990) 7189.
[12] T. Belytschko, Y.-Y. Lu, Globallocal nite elementspectral-boundary element techniques for failure analysis, Comput. Struct.
37 (1990) 133140.
[13] T. Belytschko, Y.-Y. Lu, A curvilinear spectral overlay method for high gradient problems, Comput. Meths. Appl. Mech. Eng. 95
(1992) 383396.
[14] T. Belytschko, Y.-Y. Lu, A spectral overlay method for quasi-static viscoplasticity: Applications to anti-plane crack and
comparison with closed-form solution, Comput. Struct. 51 (1994) 719728.
[15] E.N. Dvorkin, A.M. Cuiti~ no, G. Gioia, Finite elements with displacement interpolated embedded localization lines insensitive to
mesh size and distortions, Int. J. Num. Meth. Eng. 30 (1990) 541564.
[16] E.N. Dvorkin, A.P. Assanelli, 2D nite elements with displacement interpolated embedded localization lines: The analysis of
fracture in frictional materials, Comput. Meths. Appl. Mech. Eng. 90 (1991) 829844.
[17] M. Klisinski, K. Runesson, S. Sture, Finite element with inner softening band, J. Eng. Mech. ASCE 117 (1991) 575587.
[18] Th. Olofsson, M. Klisinski, P. Nedar, Inner softening bands: A new approach to localization in nite elements, in: H. Mang,
N. Bicanic, R. de Borst (Eds.), Computational Modelling of Concrete Structures, Pineridge, Swansea, 1994, pp. 373382.
[19] U. Ohlsson, Th. Olofsson, Mixed-mode fracture and anchor bolts in concrete: Analysis with inner softening bands, J. Eng. Mech.
ASCE 123 (1997) 10271033.
[20] R. Tano, Localization Modelling with Inner Softening Band Finite Elements, Licentiate thesis 1997:26, Lulea University of
Technology, Sweden, August 1997.
[21] R. Tano, M. Klisinski, Th. Olofsson, Stress locking in the inner softening band method: A study of the origin and how to reduce
the eects, in: R. de Borst, N. Bicanic, H. Mang, G. Meschke (Eds.), Computational Modelling of Concrete Structures, Balkema,
Rotterdam, 1998, pp. 329335.
[22] H.R. Lot, Finite element analysis of fracture in concrete and masonry structures, Ph.D. thesis, University of Colorado, Boulder,
1992.
[23] H.R. Lot, P.B. Shing, Analysis of concrete fracture with an embedded crack approach, in: H. Mang, N. Bicanic, R. de Borst
(Eds.), Computational Modelling of Concrete Structures, Pineridge, Swansea, 1994, pp. 343352.
[24] H.R. Lot, P.B. Shing, Embedded representation of fracture in concrete with mixed nite elements, Int. J. Num. Meth. Eng. 38
(1995) 13071325.
[25] J.C. Simo, J. Oliver, F. Armero, An analysis of strong discontinuities induced by strain softening in rate-independent inelastic
solids, Computational Mech. 12 (1993) 277296.
[26] J. Oliver, J.C. Simo, Modelling strong discontinuities in solid mechanics by means of strain softening constitutive equations, in:
H. Mang, N. Bicanic, R. de Borst (Eds.), Computational Modelling of Concrete Structures, Pineridge, Swansea, 1994, pp. 363
372.
[27] J.C. Simo, J. Oliver, A new approach to the analysis and simulation of strain softening in solids, in: Z.P. Bazant et al. (Eds.),
Fracture and Damage in Quasibrittle Structures, E. and F.N. Spon, London, 1994, pp. 2539.
[28] F. Armero, K. Garikipati, Recent advances in the analysis and numerical simulation of strain localization in inelastic solids, in:
D.R.J. Owen, E. O~ nate, (Eds.), Computational Plasticity: Fundamentals and Applications, Proc. COMPLAS IV, International
Center for Numerical Methods in Engineering, Barcelona, 1995, pp. 547561.
[29] F. Armero, Localized anisotropic damage of brittle materials, in: D.R.J. Owen, E. O~ nate, E. Hinton (Eds.), Computational
Plasticity: Fundamentals and Applications, Proc. COMPLAS V, International Center for Numerical Methods in Engineering,
Barcelona, 1997, pp. 635640.
[30] J. Oliver, Continuum modelling of strong discontinuities in solid mechanics, in: D.R.J. Owen, E. O~ nate (Eds.), Computational
Plasticity: Fundamentals and Applications, Proc. COMPLAS IV, International Center for Numerical Methods in Engineering,
Barcelona, 1995, pp. 455479.
[31] J. Oliver, Modelling strong discontinuities in solid mechanics via strain softening constitutive equations. Part 1: Fundamentals.
Part 2: Numerical Simulation, Int. J. Num. Meth. Eng. 39 (1996) 35753624.
[32] F. Armero, K. Garikipati, An analysis of strong discontinuities in multiplicative nite strain plasticity and their relation with the
numerical simulation of strain localization in solids, Int. J. Solids Struct. 33 (1996) 28632885.
M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330 329
[33] J. Oliver, D.G. Pulido, On the use of strain-softening damage constitutive equations to model cracking of concrete, in: R. de Borst,
N. Bicanic, H. Mang, G. Meschke (Eds.), Computational Modelling of Concrete Structures, Balkema, Rotterdam, 1998, pp. 363
372.
[34] J. Oliver, M. Cervera, O. Manzoli, On the use of J2 plasticity models for the simulation of 2D strong discontinuities in solids, in:
D.R.J. Owen, E. O~ nate, E. Hinton (Eds.), Computational Plasticity: Fundamentals and Applications, Proc. COMPLAS V,
International Center for Numerical Methods in Engineering, Barcelona, 1997, pp. 3855.
[35] J. Oliver, M. Cervera, O. Manzoli, On the use of strain-softening models for the simulation of strong discontinuities in solids, in:
R. de Borst, E. van der Giessen (Eds.), Material Instabilities in Solids, Wiley, Chichester, 1998.
[36] R. Larsson, K. Runesson, Discontinuous displacement approximation for capturing plastic localization, Int. J. Num. Meth. Eng.
36 (1993) 20872105.
[37] R. Larsson, A generalized ctitious crack model based on plastic localization and discontinuous approximation, Int. J. Num.
Meth. Eng. 38 (1995) 31673188.
[38] R. Larsson, K. Runesson, Cohesive crack models for semibrittle materials derived from localization of damage coupled to
plasticity, Int. J. Fracture 69 (1995) 101122.
[39] R. Larsson, K. Runesson, M.

Akesson, Embedded cohesive crack models based on regularized discontinuous displacements, in:
F.H. Wittmann (Ed.), Fracture Mechanics of Concrete Structures, Proceedings of FraMCoS-2, Aedicatio Publishers, Freiburg,
Germany, 1995, pp. 899911.
[40] R. Larsson, K. Runesson, Element-embedded localization band based on regularized displacement discontinuity, J. Eng. Mech.
ASCE 122 (1996) 402411.
[41] R. Larsson, K. Runesson, S. Sture, Embedded localization band in undrained soil based on regularized strong discontinuity
theory and FE-analysis, Int. J. Solids Struct. 33 (1996) 30813101.
[42] A. Berends, Enhanced assumed strain element for fracture, Report on graduation project, Delft University of Technology, Delft,
The Netherlands, 1996.
[43] A.H. Berends, L.J. Sluys, R. de Borst, Discontinuous modeling of mode-I failure, in: D.R.J. Owen, E. O~ nate, E. Hinton (Eds.),
Computational Plasticity: Fundamentals and Applications, Proc. COMPLAS V, International Center for Numerical Methods in
Engineering, Barcelona, 1997, pp. 751758.
[44] L.J. Sluys, Discontinuous modeling of shear banding, in: D.R.J. Owen, E. O~ nate, E. Hinton (Eds.), Computational Plasticity:
Fundamentals and Applications, Proc. COMPLAS V, International Center for Numerical Methods in Engineering, Barcelona,
1997, pp. 735744.
[45] L.J. Sluys, A.H. Berends, 2D/3D modelling of crack propagation with embedded discontinuity elements, in: R. de Borst, N.
Bicanic, H. Mang, G. Meschke (Eds.), Computational Modelling of Concrete Structures, Balkema, Rotterdam, 1998, pp. 399408.
[46] L.J. Sluys, A.H. Berends, Discontinuous failure analysis for mode-I and mode-II localization problems, Int. J. Solids Struct. 35
(1998) 42574274.
[47] M. Jirasek, Embedded crack models for concrete fracture, in: R. de Borst, N. Bicanic, H. Mang, G. Meschke (Eds.),
Computational Modelling of Concrete Structures, Balkema, Rotterdam, 1998, pp. 291300.
[48] M. Jirasek, Th. Zimmermann, Embedded crack model: I. Basic formulation, II. Combination with smeared cracks, Int. J. Num.
Meth. Eng. (in press).
[49] G. Bolzon, A. Corigliano, An interface variables formulation for embedded-crack nite elements, in: D.R.J. Owen, E. O~ nate,
E. Hinton (Eds.), Computational Plasticity: Fundamentals and Applications, Proc. COMPLAS V, International Center for
Numerical Methods in Engineering, Barcelona, 1997, pp. 16171624.
[50] Y.-J. Li, Mesh-bias study by using discontinuous shape functions, in: R. de Borst, N. Bicanic, H. Mang, G. Meschke (Eds.),
Computational Modelling of Concrete Structures, Balkema, Rotterdam, 1998, pp. 373379.
[51] H.C. Hu, On some variational principles in the theory of elasticity and plasticity, Scintia Sinica 4 (1955) 3354.
[52] K. Washizu, On the variational principles of elasticity and plasticity, Technical Report 25-18, Aeroelastic and Structures Research
Laboratory, MIT Press, Cambridge, 1955.
[53] J.C. Simo, M.S. Rifai, A class of mixed assumed strain methods and the method of incompatible modes, Int. J. Num. Meth. Eng.
29 (1990) 15951638.
[54] T.J.R. Hughes, Generalization of selective integration procedures to anisotropic and nonlinear media, Int. J. Num. Meth. Eng. 15
(1980) 14131418.
330 M. Jirasek / Comput. Methods Appl. Mech. Engrg. 188 (2000) 307330

You might also like