Vous êtes sur la page 1sur 12

SEECCM 2009

2
nd
South-East European Conference on Computational Mechanics
An IACM-ECCOMAS Special Interest Conference
M. Papadrakakis, M. Kojic, V. Papadopoulos (eds.)
Rhodes, Greece, 2224 June 2009
IMPLICIT NUMERICAL INTEGRATION OF THE MOHR-COULOMB
SURFACE IN PRINCIPAL STRESS SPACE
Fotios E. Karaoulanis
1
and Theodoros Chatzigogos
2

1
Aristotle University of Thessaloniki
Thessaloniki, GR-54124, Greece
e-mail: fkar@civil.auth.gr
2
Aristotle University of Thessaloniki
Thessaloniki, GR-54124, Greece
e-mail: thechatz@civil.auth.gr
Keywords: Nonsmooth yield surfaces, return mapping, spectral representation, Mohr-
Coulomb.
Abstract. The Mohr-Coulomb yield criterion is acknowledged as one of the first and most
important criteria, widely used to describe the yield behavior of a wide range of materials.
However singularities rising due to the nonsmoothness of the Mohr-Coulomb yield surface
introduce severe complexities that lead to different numerical approaches, proposed algo-
rithms and computer implementations. In this article an extremely simple, robust, efficient
and general return mapping method for implicit integration of nonsmooth yield criteria is
utilized. The algorithm is based on a spectral representation of stresses and strains and a re-
turn mapping scheme in principal stress directions. It is shown that in the Mohr-Coulomb
case the return mapping reduces to a one step closest point projection, minimizing the compu-
tational cost and increasing the accuracy when compared to a typical return mapping scheme.
The validation, verification and performance of the present return mapping algorithm is dem-
onstrated by numerical examples.

Fotios E. Karaoulanis and Theodoros Chatzigogos
2
1 INTRODUCTION
The Mohr-Coulomb yield criterion is acknowledged as one of the first and most important
criteria, widely used to describe the yield behavior of a wide range of materials. Named in
honor of Charles-Augustin de Coulomb and Christian Otto Mohr, uses the Coulomb's friction
hypothesis to determine the combination of shear and normal stress that will cause a fracture
of a material and Mohr's circle to determine which principal stresses will produce this combi-
nation of shear and normal stress as well as the angle of the plane in which this will occur. In
three dimensions, the Mohr-Coulomb failure surface is a nonsmooth yield surface, rendered
as a cone with a hexagonal cross section in the deviatoric stress space.
From a mathematical standpoint, the extension of classical plasticity models to accommo-
date nonsmooth yield surface goes back to the fundamental work of Koiter [1]. Later formula-
tions of plasticity employing convex analysis as in Moreau [2] encompass these classical
treatments as a particular case. Modern formulations are usually based on the work of Simo
[3], where the standard Kuhn-Tucker complementarity conditions are used to provide the
characterization of plastic loading/unloading.
However singularities rising due to the nonsmoothness of the Mohr-Coulomb yield surface
introduce severe complexities that lead to different numerical approaches, proposed algo-
rithms and computer implementations. Hence, from a computational standpoint, effort has
been initially directed towards replacing the yield surface in near singular areas (see e.g. [4]
where the authors proposed the replacement of Mohr-Coulomb and Tresca surfaces by
Drucker-Pragers and Von-Mises ones respectively, leading inevitably to a gradient jump) or
smoothing the yield surface in those areas (e.g. by modifying the yield surface in the vicinity
of singularities as in [5]). In early 1990s and within the context of the return mapping algo-
rithm [6], effort has been focused on the proper implementation of the plastic corrector near
corner regions following specific surface dependent implementations.
Formulation of the return mapping algorithm in the principal stress space is not new.
Pankaj and Biani [7] elaborate on the detection of the proper stress return in principal stress
space for the Mohr-Coulomb surface, Peri and Neto [8] for the Tresca one and Borja et al. [9]
for three invariant elastoplastic models. The main benefit of dealing with stress return in prin-
cipal stress space is that differentiation of yield surfaces w.r.t. stresses (namely
1
,
2
and
3
)
in this space is usually much simpler, since almost all yield surfaces of interest are defined in
this space, leading to more robust and efficient algorithms. Recall that within the context of
the return mapping algorithm, the yield function must be at least twice differentiable w.r.t.
stresses. However the above mentioned approaches, have been either focused on specific
yield surfaces ([7]-[8]), or avoided nonsmoothness of the yield surface [9].
In this work, the treatment of multisurface plasticity as proposed by Simo [3] is reformu-
lated and implemented in principal stress space. Contrary to previous mentioned works, a
general algorithm is proposed for the case of nonsmooth multisurface plasticity in this space,
which is applied to the Mohr-Coulomb yield surface.


2 THEORY
In what follows a review of basic notation and theory is provided, the case of multisurface
plasticity is defined and the return mapping in principal stress space is addressed.

Fotios E. Karaoulanis and Theodoros Chatzigogos

3
2.1 Summary of governing equations

Let
n
denote the total strain at a fixed point XB of a solid, X
ndim
, where
n
dim
{1,2,3}; typically X refers to a quadrature point in finite element discretization of the
equations governing its mechanical equilibrium. In the infinitesimal case, the strains are
simply identified as the symmetric part of the gradient of the displacement vector and are as-
sumed additively decomposed, as

p e
+ = , (1)
where
e
and
p
are referred to as the elastic and plastic strain parts.
Then let
n
and
n
denote the stress and a set of n

strain-like internal variables


characterizing the hardening/softening response of the material, respectively. Standard ther-
modynamic arguments identify the following constitutive relation,
,
) , (
e
e
W

= (2)
for the stored energy function W in terms of the elastic strains
e
and a general set of strain-
like variables .
The definition of the elastoplastic problem is completed with the introduction of the evolu-
tion equations for the plastic internal variables, namely the plastic strains
p
and hardening
variables , called flow rule and hardening/softening law, respectively:
). , ( ,
) , (


&
&
&
& =

=
g
p
(3)
The parameter 0
&
is a nonnegative scalar, called the consistency parameter, which is as-
sumed to obey the following Kuhn-Tucker complimentarity conditions,
0 ) , ( and 0 ) , ( , 0 = f f
& &
(4)
and the consistency requirement,
. 0 ) , ( = f
&
(5)
In classical literature, conditions (4) and (5) go by the names loading/unloading and con-
sistency conditions. The functions g(,) and ) , ( are prescribed functions that define the
direction of the flow rule and the type of hardening.
Finally the function f(,):
n

n
is usually known as the yield function and defines
the so called elastic domain, i.e. the following convex set,
} 0 ) , ( | ) , {( = f
n n

(6)
in which the admissible stresses are constrained to lie. Of special significance is the case
where:
f g, (7)
which is called associative flow rule.

Fotios E. Karaoulanis and Theodoros Chatzigogos
4
2.2 Nonsmooth multisurface plasticity

The extension of the above elastoplastic problem to nonsmooth multisurface plasticity is
rather straight forward. The essential feature is the characterization of the elastic domain E

,
which is still a convex subset of
n

n
, however now is defined as
]}, ,... 2 , 1 [ all for , 0 ) , ( | ) , {( n f
a
n n
=

(8)
where f

(,) are n1 functions intersecting nonsmoothly. With this definition of the elastic
domain, the evolution of plastic strain is governed by the following flow rule [1] often re-
ferred to as Koiters rule:
.
) , (
1

=
n
a
a a p
g



&
& (9)
Similarly the complimentarity conditions (4) and the consistency condition (5) are now re-
defined as:
. 0 ) , ( and 0 ) , ( , 0 ) , ( , 0 = =
a a a a a a
f f f
& & &
(10)

2.3 Return mapping in principal stress space

Recall that the spectral decomposition is given as:
, ,
) ( ) ( ) (
3
1
) (
) 1 ( 1
A A A A
n n
n n m m = =

=
+ +
(11)
where
A
and n
(A)
are the principal Cauchy stresses and principal directions, respectively,
denotes a juxtaposition, e.g., (ab)
ij
=a
i
b
j
, and subscript n stands for a converged step in the
incremental solution scheme. The gradients of any scalar functions of stress invariants, such
as the yield and plastic potential functions f and g, can now be evaluated as
, ,
3
1
) (
3
1
) (

= =

A
A
A
A
g g f f
m

m

(12)
with m
(A)
as defined above.
If isotropy in the elastic response is assumed, then it can be easily shown that,
,
3
1
) (
1
=
+
= + =
A e
A n n
m (13)
with m
(A)
being a spectral direction of as well, i.e. the strain spectral directions coincide with
stress spectral directions and therefore the return mapping algorithm may be conveniently
formulated in principal stress space.
Hence, starting from the flow rule (the eigenbases m
(A)
are dropped for claritys shake),

=
m
p
g
1

&
& , (14)
Fotios E. Karaoulanis and Theodoros Chatzigogos

5
where m is the number of active yield surfaces and applying a backward Euler return, yields

1
1
1
+
=
+

+ =
n
m
a
p
n
p
n
g

. (15)
Considering that
n+1
=
n+1

n
and
n+1
= c(
n+1

p
n+1
), the above equation may be written
as
0 :
1
1
1
1
1
=


+
=
+

+
n
m
a
n n
g

, (16)
or in residual form

1
1
1
1
1
:
+
=
+

=
n
m
a
n n n
g

c r


. (17)
The tensor c
-1
is the elastic compliance matrix in principal stress/strain space, which in the
case of isotropic elasticity is given as:

(
(
(

1
1
1
1
1



E
c (18)
and depends on the Youngs modulus E and Poissons ratio .
Linearization of the above equation yields:

) (
) (
1
1
) (
) (
1
1
) (
) (
) (
1
1
2
2
) (
1 ) (
1
:
k
k
n
m
a
k
k
n
m
a
k
k
k
n
m
a
k
k
n
g g g

c r


+
=
+
=
+
=

+
|
|
|

\
|

+ = (19)

Similarly linearization of the hardening rule is given as (I denoting the identity matrix):

n n n
k
n
k
k
n
k
k
n
k
n
k
r I

r + =

=
+ +
+
+
+ 1 1
) ( ) (
) (
1
) (
) (
1
) (
1
) , ( where ,

. (20)
Finally linearization of the yield function 0 =
a
f yields
. where ,
1
) (
) (
1
) (
) (
1
) (
1
a
n
fa k
k
n
a
k
k
n
a
k
n
fa
f
f f
=

=
+
+
+
+
r

r (21)

Combining the above linearized equations (19)-(21) the following system is defined:
Fotios E. Karaoulanis and Theodoros Chatzigogos
6

) (
1
1
1
) (
) (
3
1
1
1
1
1
1
) (
1
1
1
1
1

k
n
n
fa
n
k
a
k
n m
n
n
n
m
n
a
T
n
a
n
m
i
k
n
m
a
n
m
a
a
a
f f
g g g
(
(
(
(

=
(
(
(

(
(
(
(
(
(
(
(
(
(

(
(

+
+
+
+
+ +
+
+
+
+
+
=
+
=
+
=


r
r
r


(22)

for all active (m1) yield surfaces.
The stress tensor is then updated in terms of principal stresses as

+ =
+
) ( ) (
1
k
n
k
n
, (23)

the hardening parameter as

+ =
+
) ( ) (
1
k
n
k
n
, (24)

and the parameters

as

=
k
k a ) (

. (25)
The active set of surfaces, defined as
act
J :{ {1,2,...,m}| f

trial,n+1
>0} is updated during
the iteration procedure, so as the admissibility constraint

remains nonnegative for all

act
J . I.e.


is updated and if negative, the surface is removed from the active set of
surfaces.

3 IMPLEMENTATION
The above described return mapping algorithm is applied in this section to the Mohr-
Coulomb perfectly-plastic yield criterion, where the algorithms simplicity, efficiency and
robustness is demonstrated.
3.1 The Mohr-Coulomb yield criterion
The Mohr-Coulomb yield function is usually defined in terms of the difference between
the maximum and minimum principal stresses, as:
( ) ( ) ) cos( 2 ) sin(
min max min max
c f + + = , (26)
where
max
and
min
are the maximum and the minimum principal stress, is the friction angel
and c is the cohesion. It is subdifferentiable in stress space and for a fixed
y
defines an elas-
tic region given as:
{ } 0 | < = f E

, (27)
Fotios E. Karaoulanis and Theodoros Chatzigogos

7
which is plotted in Fig.1.

Figure 1: The Mohr-Coulomb criterion in principal stress space.

The above yield surface may be defined in terms of principal stresses using the following
six linear equations:


( ) ( )
( ) ( )
( ) ( )
( ) ( )
( ) ( )
( ) ( ) ) cos( 2 ) sin(
) cos( 2 ) sin(
) cos( 2 ) sin(
) cos( 2 ) sin(
) cos( 2 ) sin(
) cos( 2 ) sin(
1 2 1 2 6
2 3 2 3 5
1 3 1 3 4
2 1 2 1 3
3 2 3 2 2
3 1 3 1 1






c f
c f
c f
c f
c f
c f
+ + =
+ + =
+ + =
+ + =
+ + =
+ + =
. (28)

3.2 Return mapping scheme for the Mohr-Coulomb yield criterion

Without any loss of generality, one may assume that the principal stresses are ordered in
descending order, i.e.
1

3
and hence only three of the above constraints can be used,
namely f
1
, f
2
and f
3
defining the sextant shown in Fig.2.
The normal vectors w.r.t. to the principal stress tensor can now be easily calculated as:


(
(
(

+
+ +
=
(
(
(

+
+ + =
(
(
(

+
+ +
=
0
) sin( 1
) sin( 1
,
) sin( 1
) sin( 1
0
,
) sin( 1
0
) sin( 1
3 2 1

d
df
d
df
d
df
. (29)
while the second derivatives are all zero vectors, i.e. 0

=
2
2
d
f d
i
, and the above described algo-
rithm is applied with no modifications.
Fotios E. Karaoulanis and Theodoros Chatzigogos
8

Figure 2. The Mohr-Coulomb criterion in -plane.

It should be clear that due to the linearity of the gradients the return map degenerates to one
step closest point projection, dramatically increasing the accuracy and robustness of the algo-
rithm, while minimizing the computation cost, when compared with typical implementations
of the return mapping schemes, found in classical textbooks of finite element literature (see
e.g. [10]).

4 NUMERICAL EXAMPLES
The Mohr-Coulomb criterion is extensively used in Soil Mechanics, hence typical exam-
ples referring to soil plasticity are presented in this section. In the beginning, a series of soil
tests is presented and then the problem of vertically unsupported excavation is stated and
solved. All examples have been solved using nemesis [11], which is an experimental finite
element code, implemented by the first author.

4.1 Triaxial test
The triaxial test [12] is the most widely laboratory test used to evaluate the mechanical
properties of soils. It is performed in cylindrical soil specimens immersed in a water cell,
which provides the radial pressure
3
, while vertical pressure
1
is applied mechanically on
the bottom and top sides. Assuming a consolidated drained compression test, the loading is
applied in two stages:

1. The soil is consolidating isotropically to a desired effective stress level by pressur-
izing the water in the cell.
2. The water pressure in the cell is kept constant and additional axial displacements in
a very slow rate are added until the soil fails.
Fotios E. Karaoulanis and Theodoros Chatzigogos

9

Figure 3: Idealized triaxial test and corresponding finite element model.
The finite element model consists of one quadrilateral, displacement based, axisymmetric
element as shown in Fig.3. Two load cases are defined:

1. P
1
= P
3
= 1 using load control and thus simulating the isotropic consolidation.
2. P
3
= 1, P
1
= , where is found using displacement control such as an imposed dis-
placement of 0.005 inwards (compression, see Fig.4(a)) or outwards (tension, see
Fig.4(b)).


(a)

(b)
Figure 4: Numerical and analytical results for the triaxial test, in case of compression (a) and tension (b).
In both cases the results are identical to the analytical solution, namely P
1
=44.884 for com-
pression and P
1
=20.516 for tension respectively.

4.2 Direct shear test
The direct shear test [12] is used to find the shear strength parameters of soil. A soil speci-
men is sheared under a constant vertical force and fails across a predefined zone, constrained
by the test apparatus. From the recorded applied forces and the resulting displacements the
Fotios E. Karaoulanis and Theodoros Chatzigogos
10
peak and residual shear strength, the friction and the dilation angle and as well as the cohesion
can be determined.

Figure 5: Idealized shear test and corresponding finite element model.
In Fig.5 an idealized shear test is shown, which will be used to verify the accuracy of the
implemented model, concerning a non-associative flow rule, which in the case of the Mohr-
Coulomb yield criterion is defined using the dilation angle instead of the friction angle ,
when enforcing the normality hypothesis (see Eq. 9 and 28). The finite element model con-
sists of one quadrilateral, displacement based, plane strain element as shown in Fig.5. Assum-
ing for the material E=1000., =0.2, c=15, =20
0
and =10
0
, two load cases are defined:

1. P
1
= 1 using load control.
2. P
1
= 1, P
2
= , where is found using displacement control such as an imposed dis-
placement of 0.01 is applied.

The results are plotted in Fig.6, where the horizontal vs. the vertical displacements are plotted;
it can be verified that the resulted dilation angle, i.e. a =
|
|

\
|


xy
y

arctan =
0
10 arctan =
|
|

\
|


x
y
u
u
, equals to the given one.

Figure 6: Horizontal vs. vertical displacements in the numerical shear test. The dilation angle is also shown.

Fotios E. Karaoulanis and Theodoros Chatzigogos

11
4.3 Unsupported excavation
Coulomb (see e.g [14]) proposed that a condition of limit equilibrium exists through which
a soil mass behind a vertical retaining wall will slip along a plane inclined an angle
4 2

+ = to the horizontal (Fig.7(a)). Therefore, the critical height H
cr
can be found, defined
as the maximum height for an unsupported excavation (P
a
=0) and equals to:

1
4 2
tan 4 |

\
|
+ = c H
cr
(30)
A finite element analysis was performed for a soil described by E=40000kPa, =0.2,
=20kN/m3, c=13.23771kPa and =23
0
which yield according to (34) a H
cr
=4.0m. Two load
cases are defined:

1. An initial stress field is applied for K
0
=1 sin() .
2. The excavation load case which is performed in one stage.

In Fig.7(b) the excavation, the slip line and the plastification zone are shown when a fail-
ure mechanism is deployed and the soil body collapses for an excavation depth of 4.0m, in
perfectly agreement with the analytical solution.


(a)

(b)
Figure 7: Coulomb failure wedge (a) and numerical results for an unsupported excavation (b).

5 CONCLUSIONS
The Mohr-Coulomb yield criterion is acknowledged as one of the first and most important
criteria, widely used to describe the yield behavior of a wide range of materials. However sin-
gularities rising due to the nonsmoothness of the Mohr-Coulomb yield surface introduce se-
vere complexities that lead to different numerical approaches, proposed algorithms and
computer implementations. In this article, a robust and efficient return mapping method for
the implicit integration of nonsmooth yield criteria, based on a spectral representation of
stresses and strains and a return mapping scheme in principal stress directions is presented
Fotios E. Karaoulanis and Theodoros Chatzigogos
12
and applied to the Mohr-Coulomb criterion. It is shown that in this case the return mapping
reduces to a one step closest point projection, minimizing the computational cost and increas-
ing the accuracy when compared with a typical return mapping scheme. Test cases of standard
soil tests and an example of unsupported excavation failure are provided, with the results ob-
tained to be in excellent agreement with the corresponding analytical solutions.
AKNOWLEDGEMENTS
The author gratefully acknowledge financial support from the Greek State Institute of
Scholarships (I.K.Y.); contract/grant number: 4506/05.
REFERENCES
[1] W.T. Koiter, General Theorems for Elastic-plastic Solids, in Progress in Solid Mechan-
ics 6, 167-221, eds. I.N. Sneddon and R. Hill, North-Holland Publishing Company,
Amsterdam, 1960,
[2] J.J. Moreau, Application of Convex Analysis to the Treatment of Elastoplastic Systems,
in Applications of Methods of Functional Analysis to Problems of Mechanics, eds. P.
Germain and B. Nayroles, Springer-Verlag, Berlin, 1976.
[3] J.C. Simo and T.J.R. Hughes, Computational Inelasticity. Springer-Verlag, 1998.
[4] D. R. J. Owen, E. Hinton, Finite Elements in Plasticity: Theory and Practice. Pineridge
Press, Swansea, U.K., 1980.
[5] S.W. Sloan, J.R. Booker, Removal of singularities in Tresca and Mohr-Coulomb yield
functions. Communications in Applied Numerical Methods, 2, 173-179, 1986.
[6] J.C. Simo, R.L. Taylor, Return Mapping Algorithm for Plane Stress Elastoplasticity. In-
ternational Journal for Numerical Methods in Engineering, 22, 649-670, 1986.
[7] Pankaj, N. Biani, Detection of multiple active yield conditions for MohrCoulomb
elastoplasticity. Computers and Structures, 62(1), 51-61, 1997.
[8] D. Peri, E.A. de Souza Neto, A new computational model for Tresca plasticity at finite
strains with an optimal parametrization in the principal space. Computer Methods in
Applied Engineering, 71, 463-489, 1999.
[9] R.I.Borja, K.M. Sama, P.F. Sanz, On the numerical integration of three-invariant elas-
toplastic constitutive models. Computer Methods in Applied Mechanics and Engineer-
ing. 192, 1227-1258, 2003.
[10] M.A. Crisfield, Nonlinear Finite Element Analysis of Solids and Structures, Vols. 1 &
2, John Wiley & Sons, 1991.
[11] nemesis, an experimental finite element code, http://www.nemesis-project.org. Re-
trieved on February 2009.
[12] B. Muni, Soil Mechanics & Foundations, John Wiley & Sons, 2000.

Vous aimerez peut-être aussi