Sunteți pe pagina 1din 19

Article

A nonlinear programming approach


to limit analysis of non-associated
plastic flow materials

Mathematics and Mechanics of Solids


18(5): 524542
The Author(s) 2012
Reprints and permissions:
sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/1081286512444749
mms.sagepub.com

Hua-Xiang Li
Department of Civil Engineering, University of Nottingham, Nottingham, UK
Received 17 February 2012; accepted 18 March 2012
Abstract
A nonlinear programming approach, along with finite element implementation, has been developed to perform plastic
limit analysis for materials under non-associated plastic flow, so the plastic stability condition of non-standard materials
can be directly calculated. In the framework of Radenkovics theorems, a decoupled material model with non-associated
plastic flow was introduced into kinematic limit analysis so that the classic limit theorems were extended for non-standard
plastic flow materials, such as cohesive-frictional materials. A non-associated plastic dissipation power is derived, and
a purely kinematic formulation is obtained for limit analysis. Based on the mathematical programming theory and the
finite element method, the numerical implementation of kinematic limit analysis is formulated as a nonlinear programming
problem subject only to one equality constraint. An extended direct iterative algorithm is proposed to solve the resulting
programming problem. The developed method has a wide applicability for limit analysis. The effectiveness and efficiency
of the proposed method are validated through numerical examples and the influence of non-associated plastic flow on
stability conditions of structures is numerically investigated.

Keywords
Nonlinear programming, limit analysis, non-associated plastic flow, finite element method, cohesive-frictional material

1. Introduction
Limit analysis is a direct method for stability analysis of structures subject to quasi-static loads. It can be
used to calculate the plastic limit loads of structures in a direct way and provides a theoretical foundation for
engineering design and integrity assessment of structures. Compared with the traditional incremental elasticplastic method, which involves the calculation of the full history of load-deformation response to find the
limit load of a structure, limit analysis can obtain the collapse load by searching the critical failure point
and requires much less computational effort. Therefore, limit analysis provides a very rigorous method for
the stability analysis of structures. The limit analysis is based on two dual theorems, the static (lower-bound)
limit theorem [1] and the kinematic (upper-bound) limit theorem [24]. Due to the complexity of engineering
problems, numerical methods are often necessary to implement limit analysis. In the last several decades, many
studies have been devoted to the numerical implementation of limit analysis, especially when the mathematical
programming technique was introduced and the finite element method was applied. However, most of these
works are focused on standard materials (rigid, perfectly plastic body with associated plastic flow).
A linear programming (LP) approach was first developed [59]. One advantage of this approach is that it
enables solving large scale problems. The fundamental technique in this approach is to discretize the nonlinear
Corresponding author:
Hua-Xiang Li, Department of Civil Engineering, University of Nottingham, Nottingham NG7 2RD, UK.
Email: huaxiang.li@hotmail.com

Li

525

yield surface and then use the linearized yield condition in limit analysis. Therefore, some additional constraints
and computational errors will be introduced. In the latter, a nonlinear programming (NLP) approach has been
proposed and widely used in limit analysis [10-15], where the nonlinear yield surface was not discretized and
directly introduced into limit analysis.
Based on the elastic compensation method [16, 17], Ponter et al. [18, 19] developed a linear numerical
method (the linear matching method) to implement kinematic limit analysis. The linear solutions are defined
with a spatially varying shear modulus, which provides a sequence of upper bounds to the limit load. A similar approach, the modified elastic compensation method, was presented by Chen et al. [20] to overcome the
numerical singularity and computational errors in the limit analysis of a structure containing flaws.
By introducing the interior point optimization method [21] into limit analysis, Pastor et al. [22] developed
a decomposition approach to implement the kinematic limit analysis in the plane strain condition. A mixed
kinematic formulation was finally obtained to calculate the dissipation power on discontinuous surfaces in
terms of velocity and stress.
It should be mentioned that almost all of the above works are based on limit analysis for isotropic, pressureindependent materials that obey the von Mises and Tresca yield criteria. To implement limit analysis for soil
materials that obey a pressure-dependent yield rule (e.g. the MohrCoulomb or DruckerPrager yield criterion),
Li and Yu [23] developed a nonlinear programming approach to implement the kinematic limit analysis for these
materials, and they found that an upper bound to limit load can be obtained. Recently, Makrodimopoulos and
Martin [24, 25] developed a second-order cone programming approach to implement both upper- and lowerbound limit analyses for soil materials. The MohrCoulomb yield criterion under the plane-strain condition
was assumed.
There are only a few results about anisotropic limit analysis because it is much more difficult to introduce an
anisotropic yield criterion into limit analysis. An early work was from Kao et al. [26], who theoretically studied
the static and kinematic stability conditions of an orthotropic plate. Hills yield criterion was used and the
plane-stress model was assumed. Zhang and Lu [27] presented a nonlinear programming approach to solve the
upper-bound limit problem of axisymmetric shells, where a two-dimensional Hills yield condition was used.
Recently, a more general nonlinear programming approach has been developed for Hills materials [2831] and
for the TsaiWu type of materials [23, 32, 33]. The three-dimensional model for kinematic limit analysis was
fully implemented.
As composite and inhomogeneous materials are increasingly used in engineering, limit analysis of
microstructures has drawn much attention in recent years. Francescato and Pastor [34] first presented a linear programming approach with the homogenization technique to apply limit analysis to composites. A convex
optimization [3538] was recently proposed to apply limit analysis to porous materials using the Gurson model,
where the homogenization technique for heterogeneous materials was applied. Based on the homogenization
theory, Li et al. [15, 29] and Li and Yu [31] developed a nonlinear programming method to apply the kinematic
limit analysis for heterogeneous materials obeying the von Mises and ellipsoid yield criteria, where nonlinear
yield surface was directly introduced into limit analysis without discretization. Dallot and Sab [39, 40] also
used the homogenization technique to study the limit status of multilayered plates.
Another important application of limit analysis to engineering is to get specific solutions for some engineering structures so that a complex engineering problem can be simplified and some special structural features can
be fully considered, e.g. for arches [41], for frames [4248], for masonry [4956], for friction contacts [57],
and for cylinders [5861].
It should be noted that most of current numerical implementations of limit analysis are carried out by
applying the finite element method. However, Zhang et al. [62, 63] presented a symmetric Galerkin boundary
method to numerically implement the lower-bound limit analysis. Chen et al. [64] and Le et al. [65] used the
element-free Galerkin (EFG) method to numerically perform limit analysis.
The type of the plastic flow is a key condition for limit analysis. Most of the current works on limit analysis
are presented for standard materials (rigid, perfectly plastic model with associated plastic flow). However, it
is well known that most soil materials are non-associated with plastic flow. Moreover, due to the diversity of
engineering materials, more and more materials are being found to have non-associated plastic flow. However,
up to now, there are only a few preliminary studies on limit analysis for non-associated plastic flow materials.

526

Mathematics and Mechanics of Solids 18(5)

The early theoretical study of limit analysis for non-associated materials was first discussed by Drucker [66]
and then later formulated by Radenkovic [67]. The further applications and modifications were presented by
Josselin de Jong [68], Palmer [69], Sacchi and Save [70], Collins [71], and Salencon [72]. Recently, De Saxce
and Bousshine [73] studied the stability condition of soils and rocks under non-associated plastic flow using
the bipotential approach. A coupled upper-lower bound formulation was obtained to calculate the limit state of
implicit standard materials.
The purpose of this paper is to develop a general nonlinear programming approach for limit analysis under
non-associated plastic flow. First, the plastic dissipation power is explicitly expressed in terms of kinematic
admissible velocity by introducing a convex non-associated plastic rule into a general yield criterion. Based
on the nonlinear programming technique, the kinematic limit analysis is formulated as a purely kinematic
formulation and the corresponding finite element model is expressed as a nonlinear programming problem
subject only to one equality constraint. The objective function corresponding to plastic dissipation power is to be
minimized, which can be solved by an extended direct iterative algorithm. Finally, some numerical examples are
given to illustrate the validity, and the efficiency of the proposed approach and the influence of non-associated
plastic flow on limit loads are investigated.

2. A non-associated plastic-flow material


Classic limit analysis is based on the extreme theorem that is proposed for the rigid, perfectly plastic material
model with the associated plastic-flow rule. To capture more plastic behaviours of materials, such as cohesivefrictional materials, a non-associated plastic-flow rule is used in this paper. First, the yield criterion of materials
is mathematically expressed by a second-order polynomial, which can be used for most of current yield criteria, e.g., isotropic material (von Misess criterion), orthotropic material (Hills criterion), generally anisotropic
material (TsaiWu criterion), and pressure-dependent material (MohrCoulomb or DruckerPrager criterion).
Then, a decoupled plastic-flow law is assumed to capture both associated and non-associated plastic flow
of materials. The general expression of plastic-flow rule will be directly introduced into the kinematic limit
theorem so that the effect of non-associated plastic flow on the stability condition of materials can be revealed.
For the simplicity of expression when the finite element method is used, the responses of the body to the load
(true stress , true strain , and
in the three-dimensional
displacement

u) are represented as column


vectors,
i.e.
model, = [11 , 22 , 33 , 212 , 223 , 213 ]T , = [11 , 22 , 33 , 212 , 223 , 213 ]T , and u= [u1 , u2 ,
u3 ]T .

2.1. A general yield criterion


Most yield criteria (isotropic, anisotropic, or pressure-dependent) of materials can be expressed in a general
form as follows:
F( ) = T P + T Q 1 = 0
(1)
where F( ) defines a yield function in terms of strength parameters. P and Q are coefficient matrices and are
related to the strength properties of the material.
Generally speaking, the physical laws of materials must be objective, i.e. frame indifference. Therefore, in
plasticity theory, the general yield condition should depend only on the invariants of a stress tensor. However,
for a specific material on a determined topic, an explicit expression in terms of stress tensor is mostly used.
Moreover, for finite element implementation, all discretized variables must be at the bottom level. So it is convenient to use equation (1) as the yield expression from the start of the analysis. Then, the material coefficient
matrices P and Q in equation (1) will be determined from the physical law, which is originally expressed in
terms of the invariants of the stress tensor for any specific material.
For example, the MohrCoulomb and DruckerPrager yield criteria are often used for cohesive-frictional
material, which is normally accompanied by a non-associated plastic-flow rule. And these criteria are mostly
expressed as follows:

(2)
F(I1 , J2 ) = 0 I1 + J2 c0 = 0

Li

527

where I1 is the first invariant of the stress tensor, J2 is the second invariant of the deviatory stress tensor, 0 and
c0 are the strength parameters of the material.
The general yield equation (1), can be used for the above material, equation (2), when the coefficient
matrices P and Q are chosen as

1302
1+602
1+602
6c2
6c2
0
0
0

3c20
0
0

1+602
1302
1+602
2
6c2
0
0
0

6c0
3c20
0

1+6 2
2
2
1+6
13


0
0
0

0
0
0

2
2
2
(3)
P=
6c0
6c0
3c0

0
0
0
0
0

2c20

0
0
0
0 2c2 0

0
0
0
0
0
0 2c12
0

Q=

20
,
c0

20
,
c0

20
,
c0

T
0, 0, 0

(4)

In a similar way, it is not difficult to determine the material coefficient matrices P and Q, when the general yield
equation (1), is used for any other material, e.g. the von Mises- or Hill-type material.
It should be noted that when Q = 0, the general yield equation (1) is reduced for a pressure-independent
material. And for a generally anisotropic material, such as TsaiWu-type material, the coefficient matrices P
and Q can be directly determined. For example, the quadratic TsaiWu criterion is normally expressed as the
following form
(5)
F( ) = Aij ij + Bijkl ij kl 1 = 0
where Aij and Bijkl (i, j, k, l = 1, 2, 3) are the material-strength parameters. The material-strength parameter Aij
can be experimentally determined by a pure tension/compression or shear test, and the material parameter Bijkl
reflects the interaction effects of stress components.
When the stresses are expressed in the column vector, the TsaiWu yield criterion, equation (5), has the
exact same expression of equation (1), and it is not difficult to find the relation between the coefficients P and
Q and the material parameters Aij and Bijkl .

2.2. Non-associated plastic-flow rule


When the external force is beyond the yield stress of a material, plastic flow and plastic deformation will occur.
Accordingly, a plastic-flow rule is required to determine the pattern of deformation. Although for metal materials it is normally assumed that the plastic flow is associated, this assumption may conflict with the experimental
results for some other materials, e.g. soil or granular materials. In practice, for soil or granular materials, a
non-associated plastic flow is often adopted by using the dilation angle in the plastic-flow rule instead of the
friction angle used in the yield criterion.
The plastic-flow rule determines both the direction and magnitude of plastic strain-rate and is normally
expressed by a plastic potential ( ), as in
p =

( )

(6)

where ( ) denotes a plastic potential function that resembles the yield function F( ), and is a non-negative
plastic proportionality factor. If the associated plastic flow is assumed for a material, ( ) will be equivalent to
F( ). However, in this paper, the non-associated plastic flow is adopted, i.e. ( ) = F( ), so that the developed
method has a general applicability. A similar second-order polynomial is used in this paper to express the plastic
potential function
(7)
( ) = T Pf + T Qf 1 = 0

528

Mathematics and Mechanics of Solids 18(5)

where Pf and Qf are the coefficient matrices of plastic flow, and Pf is symmetric.
For a specific material, there may be an indirect way to determine the plastic potential function. For example,
for a material with MohrCoulomb or DruckerPrager yield criterion, the plastic flow can be determined by
decomposing the plastic potential into two partsthe deviatory plastic flow and the bulk plastic flow as

D
(S) B (m ) B
p
+
(8)
I
=
S
m
where D (S) and B (m ) denote the deviatory- and bulk-plastic potential functions, respectively, S and m are
the deviatory and mean stresses, respectively, and the coefficient matrix I B = [1, 1, 1, 0, 0, 0]T .
Equation (8) is frequently used to express the non-associated plastic flow of cohesive-frictional materials,
e.g. soils. Accordingly, the plastic strain-rate based on the non-associated plastic flow for the type of Mohr
Coulomb or DruckerPrager material can be calculated as


S
18(tan)2 m B
p
(9)
+
I
=
c2
c2
Then, based on the equations (6) and (7), the coefficient matrices Pf and Qf in the plastic-flow rule can be
determined. In this way, the matrix Pf can be ensured to be a positive definite or semi-definite.

2.3. Plastic dissipation power


According to the plastic-flow potential, equation (7), and plastic-flow rule, equation (6), the plastic strain-rate
can be expressed as
f + Q
f
(10)
p = 2P
Then, the stress vector at the yield surface can be expressed in terms of the strain-rate as
=

1 f 1 p 1 f 1 f
(P ) (P ) Q
2
2

(11)

Since Pf is a positive definite or semi-definite matrix, (Pf )1 can be uniquely determined when Pf is nonsingular. However, the matrix Pf may be semi-singular when Pf is a positive semi-definite. For these materials,

1
(Pf )1 can be approximately determined as Pf + I , where I is the identity matrix of the same order with
the matrix Pf and is a very small real number. Theoretically, the smaller is, the more accurate the numerical
calculation will be. However, due to the storage limitation of a computer, in case of data overflow in a numerical
computation, there is a limit for the choice of the small value . In practice, the value , which is between
106 pmin and 1012 pmin , can guarantee both numerical precision and safe data storage in a computer program,
where pmin is the smallest diagonal component of the matrix Pf .
Since the stress at the yield surface must be satisfied with the yield criterion, as in equation (1), then it is
not difficult to find the non-negative plastic-flow factor by introducing equation (11) into the yield criterion
in equation (1), and finally the non-negative plastic-flow factor can be obtained as
= 

( p )T (Pf )1 P(Pf )1 p
( p )T (Pf )1 (4P + QQT )(Pf )1 p + ((Qf )T (Pf )1 P(Pf )1 QT (Pf )1 ) p

(12)

Then, the plastic-dissipation power for a material obeying the yield criterion in equation (1) and the plasticflow rule in equation (7) can be expressed as follows
D( p ) = T p

T
= 21 (Pf )1 p 12 (Pf )1 Qf p
=

1
( p )T (Pf )1 p
2

1
(Qf )T (Pf )1 p
2

(13)

Li

529

Finally, the plastic-dissipation power for a decoupled yieldplastic flow material model is expressed in terms
of a pure strain-rate field without any stress fields. Once the kinematically admissible velocity is obtained, the
plastic-dissipation power can then be determined.

3. Kinematic limit analysis based on non-associated plastic flow


The classic kinematic limit theorem for standard materials was first proposed by Hill [2] and Drucker [3, 4] to
define the stability condition of materials by which the plastic limit load of materials can be directly calculated.
Later, the theorem was extended for non-associated plastic-flow materials by Radenkovic [67].

3.1. Classical kinematic limit theorem


The collapse load of a structure under quasi-static loading can be found by applying the kinematic theorem
of limit analysis [24], which states that among all kinematically admissible velocities, the real one yields the
lowest rate of plastic-dissipation power. In terms of the kinematically admissible velocities, the kinematic limit
theorem can be expressed as follows



T
T
t u d +
f u dv
D( p )dv
(14)

t
V
V
where is the limit load multiplier, t is the reference load of external tractions, f is the reference load of body
forces, u is the displacement velocity, p is the plastic strain rate, D( p ) represents a function for the rate of the
plastic-dissipation power in terms of the admissible strain rate p , the superscript * stands for a parameter
corresponding to the kinematically admissible velocity, t denotes the traction boundary, and V represents the
space domain of the structure.
If the body force is omitted and the geometry-compatible conditions are introduced, the kinematic limit
theorem represented in equation (14) can be re-expressed as the following


t tT u d V D( p )dv
(15)
s.t.
p =
(u)
in V

u = 0
on u
where s.t. is the abbreviation of subject to, u denotes the displacement boundary, and
(u) is a linear
differential operator that defines the geometry-compatibility condition of deformation. The purpose of the
kinematic limit analysis is to find the minimum, optimised limit multiplier with t being the limit load of the
structure.

3.2. Radenkovics theorems


The classic limit theorems of limit analysis are proposed based on the assumption of the rigid, perfectly plastic
model with associated plastic flow. To apply them to non-standard materials, some special constraints must
be imposed on the material model so that the bases of the limit theorems are still valid. It is not possible
now to theoretically prove a sufficient and necessary condition on limit analysis for a general non-standard
material model. However, it is feasible and important to find a sufficient condition for a special kind of nonstandard material. In order to extend limit analysis to non-standard materials such as non-associated plastic
flow materials, Radenkovic proposed two theorems [67, 74] for further application of classic limit analysis to
soil materials.
Radenkovics first theorem states that the limit loading for a body made of a non-standard material is
bounded from above by the limit loading for the standard material obeying the same yield criterion.
Radenkovics second theorem states that the limit loading for a body made of a non-standard material is
bounded from below by the limit loading for the standard material obeying the yield criterion g( ) = 0, which
has the following properties:

530

1.
2.
3.

Mathematics and Mechanics of Solids 18(5)

g( ) is a convex function.
The surface g( ) = 0 lies entirely within the yield surface F( ) = 0.
To any with F( ) = 0, there corresponds a  such that p is normal to the surface g( ) = 0 at  and

( )T p 0.

Radenkovics first theorem defines an upper bound to the plastic limit load of a non-standard material, where
the upper bound is the plastic limit load of a standard material that has the same yield condition as the nonstandard material. Radenkovics second theorem defines a lower bound to the plastic limit load of a non-standard
material, where the lower bound is the plastic limit load of a standard material that has the same plastic flow
potential as the non-standard material. It is well known that limit analysis of standard materials provides a strict
plastic limit load. Therefore, for a non-associated plastic-flow material, if its plastic-flow potential ( ) = 0
is satisfied with the above three conditions, the limit load obtained from using both the non-associated plastic
flow ( ) = 0 and the yield function F( ) = 0 should lie between the upper and lower bounds of Radenkovics
prediction on the two corresponding standard materials.
In practice, Radenkovics theorems are often used as a criterion to create a non-associated plastic-flow
surface for a non-standard material so that the sufficient condition of limit theorems for this kind of material
is satisfied. For examples, for MohrCoulomb or the DruckerPrager type of material, which is often used for
non-associated plastic-flow materials in civil engineering, the dilation angle in the plastic-flow rule must be
less than the frictional angle in the yield condition. The non-associated plastic-flow rule, which is defined in the
above section and frequently used for cohesive-frictional materials, is based on this strict requirement so that the
plastic-flow function is satisfied with the full conditions of g( ) in Radenkovics second theorem. Consequently,
the kinematic limit analysis as given in equation (14) with the plastic dissipation power in equation (13) can
obtain a kinematic limit solution for the non-associated plastic-flow material.

3.3. Nonlinear programming of kinematic limit analysis


The kinematic limit equation (15) provides a direct calculation of the upper bound ( t) to the limit load of
a structure. Based on the solution of the plastic-dissipation power as given in equation (13) for a non-standard
material, the kinematic limit analysis for a material with non-associated plastic flow can then be expressed as
follows


 

t tT u d V 21 ( p )T (Pf )1 p 12 (Qf )T (Pf )1 p dv


(16)
p =
(u)
in V
s.t.

u = 0
on u
The kinematic limit analysis as given in equation (16) is actually an optimization problem subject to equality
constraints. Based on the mathematical programming theory, the kinematic limit analysis, equation (16), can
be formulated as the following nonlinear programming equation:

  1 p T f 1 p 1 f T f 1 p 

( ) (P ) 2 (Q ) (P ) dv

=
min

V 2 T
s. t.
(17)
t t u d = 1

(
u
)
in
V

u = 0
on u
When the plastic flow factor is found by equation (12), the kinematic limit analysis, as in equation (17), can
then be expressed in terms of purely kinematic variables as




( ( p )T Rq p +T f p )( p )T (Pf )1 p

f T f 1 p
1

dv
2 (Q ) (P )

V
= min
2( p )T Rf p
p


T
s. t.
(18)
t t u d = 1

(
u
)
in
V

u = 0
on u

Li

531

where the coefficient matrices Rq , Rf and T f are defined as


Rf = (Pf )1 P(Pf )1

(19)

Rq = (Pf )1 (4P + QQT )(Pf )1

(20)

T f = (Qf )T Rf QT (Pf )1

(21)

Finally, the kinematic limit analysis under non-associated plastic flow is formulated as a minimum optimization
problem subject to equality constraints. The objective function is nonlinear. For the purpose of a potential
iteration algorithm, a ratio factor f ( p ) is defined as follows
f =

( p )T (Pf )1 p

(22)

( p )T Rf p

Then, a reduced form for the kinematic limit analysis, equation (18), can be obtained as





T q p
f T f 1 p
f
1
f p
p

dv
( ( ) R + T ) 2 (Q ) (P )
= min

V
2


T
s. t.
t t u d = 1

=
(u)
in V

u = 0
on u

(23)

The ratio-factor f ( p ) is a solution/material-dependent variable that is related to the plastic-flow type of materials. For a standard (associated plastic flow) material, due to P = Pf and Rf = (Pf )1 , the ratio-factor f ( p )
will become 1.0 ( f 1.0) for any deformation conditions.

4. Finite element modelling


The traditional displacement-based finite element method is used in this paper to implement the numerical
calculation for the kinematic limit analysis as given
in equation (23). Based on the finite element technique, a
structure is discretized by finite elements Ve (V = Ne=1 Ve ), and then the displacement velocity and strain rate
can be interpolated in terms of an unknown nodal displacement velocity vector:
u e (x) = Ne (x)e

(24)

e (x) = Be (x)e

(25)

where, with reference to the eth finite element, e is the nodal displacement column vector, Ne is the
interpolation function, and Be is the strain function.
By means of the Gaussian integration technique, the objective function in equation (23) can be discretized
and expressed in terms of the nodal displacement velocity as follows



T
T
1
f
f
1
f
f
p
p
( ( p ) Rq p + T ) 2 (Q ) (P ) dv
V
2



N 

T






T
1
f
f
1
f
f
q
=
Be e R Be e + T Be e 2 (Q ) (P )
Be e dv
Ve
2
e=1
(26)
 f  


N 
IG

eT BTe i Rq (Be )i e + T f (Be )i e 12 (Qf )T (Pf )1 (Be )i e
(e )i |J |i 2
=
e=1 i=1



N 
IG

f
1
T
=
(e )i |J |i
e (Ke )i e + (He )i e (Ge )i e
e=1 i=1

where (e )i is the Gaussian integral weight at the ith Gaussian integral point in the element e, |J |i is the determinant of the Jacobian matrix at the ith Gaussian integral point, IG is the number of Gaussian integral points
in the finite element e, and
(27)
Ke = BTe Rq Be

532

Mathematics and Mechanics of Solids 18(5)

He = T f Be
f 1

Ge = (Q ) (P ) Be
f T

(28)
(29)

By introducing the transformation matrix of each element Ce , the nodal displacement velocity vector e for
each element can be expressed by the global nodal displacement velocity vector for the structure as
e = Ce
Then, equation (26) can be expressed in terms of the global displacement velocity as follows



T q p
f T f 1 p
f
1
f p
p
( ( ) R + T ) 2 (Q ) (P ) dv
V
2



N 
IG

f
1
T

=
e (Ke )i e + (He )i e 2 (Ge )i e
(e )i |J |i 2
e=1 i=1






f
T Ki + Hi 12 Gi
= i |J |i 2

(30)

(31)

iI

where I denotes the set of all Gaussian integral points of the FE-discretized structure, and
Ki = CeT (Ke )i Ce

(32)

Hi = (He )i Ce

(33)

Gi = (Ge )i Ce

(34)

Accordingly, the normalization condition in the kinematic limit analysis as defined in equation (23) can be
discretized as
F T = 1
(35)
where F is the column vector of the equivalent reference nodal loads.
Finally, the finite element modelling of the kinematic limit analysis can be expressed as the following
discretized, nonlinear programming problem:

 


= min  |J | f
T Ki + Hi 1 Gi

i
i
2
2
p iI
(36a, b)
s. t.
T
F =1
After the displacement boundary condition is imposed by utilizing the conventional finite element technique,
a minimum optimized upper bound to the plastic limit load multiplier of the structure can be obtained by
solving the above mathematical programming equation. The plastic limit load of the structure is given by F.

5. Numerical algorithm
The kinematic limit analysis as given in equation (36) is a nonlinear, minimum optimization problem subject to
only one equality constraint. It can be solved by a direct iterative algorithm that was developed [15, 23, 29, 31]
for limit analysis of an associated plastic flow material. To extend it for non-associated plastic-flow materials,
an additional iterative control factor needs to be introduced. The basic strategy of the numerical algorithm is
given below. First, based on the mathematical programming theory, the equality constraint of the normalization
condition in equation (36b) is introduced into the optimization problem by means of the Lagrangian method
[75, 76]. As a result, an unconstrained minimum optimization problem is obtained as follows


f 

1

F
T

L )=
Ki + Hi Gi + LF (1 F T )
(37)
i |J |i
(,
2
2
iI
where LF are the Lagrangian multiplier.

Li

533


Then, according to the KuhnTucker stationary condition [75, 77], once applying
= 0, L
F = 0 to

equation (37), the following set of equations is obtained to solve the kinematic limit analysis equation (36):





Ki
f
1 T
T

i |J |i 2
+ Hi 2 Gi LF F = 0
(38a, b)
T Ki
iI
T
F =1

It is quite difficult to directly solve the set of equations (38a,b) because it is nonlinear and non-differentiable
as well. An iteration technique is needed to linearize the equation set (38a,b) and to distinguish the nondifferentiable areas. In order to implement an iteration technique, the set of equations (38a,b) is re-expressed
as follows



 |J | f ICP K + H T  1 GT LF F = 0
i
i 2
i
i
2 i
(39a, b)
iI
F T = 1
where ICP is the coefficient parameter, which is defined by

ICP


1
T

=
Ki

(40)

and the superscript ICP indicates that ICP is an iteration control parameter.
By solving the linearized equation (39a,b), the variable can be determined. Then the limit load multiplier
can be calculated as
f 




1
T

=
i |J |i
Ki + Hi Gi
(41)
2
2
iI
In order to trigger an iteration to linearize equation (38a,b), two iteration seeds are defined as
ICP 
= 1.0

0
f
0 = 1.0

(42)
(43)

where the subscript 0 denotes that the variables are determined at step 0. During the iterative calculation, these
two iteration factors need to be updated at each iterative step.
Once the iteration starts, it will be repeated until the following convergence criteria are satisfied


h+1 h 
|h+1 h |

 2
(44a, b)
1 ;
h+1 
|h+1 |
where h is the number of the iterative step, and 1 and 2 are the computational error tolerances.
Once the convergence condition in equation (44) is attained, the limit load multiplier at the current iterative
step is the optimized solution to the real limit load. More details about the direct iterative algorithm can be
found in the work of Li et al. [15, 23, 29, 31].
Mathematically speaking, equation (18) of kinematic limit analysis is a fractional programming subjected to
equality constraints, and the finite element approximation will lead to the so-called sum-of-ratios problems,
which have been studied extensively in global optimization for several decades [84]. The uniqueness of the
global optimal solution is governed by the convexity of the problem. As it was indicated in [84], the fractional
programming problems are generally non-convex and may have multiple local optima. Therefore, such problems are usually NP-hard. In order to solve such problems, a canonical duality theory has been proposed in
Fang et al. [84]. The direct iterative algorithm proposed in this paper is based on the strict requirements on
the material laws and purely kinematic analysis. The initial ratio factor f and the initial parameter ICP are
triggered based on the upper-bound kinematic assumption. If a new triggering mechanism is adopted, these two
iterative factors must be carefully chosen. A general theory for these kinds of nonlinear programming problems

534

Mathematics and Mechanics of Solids 18(5)

can be found in Fang et al. and Gao [84, 85]. Moreover, for a similar problem to equation (23), Gao [86] proposed a pan-penalty method to solve the difficulty of the flat objective function. Fundamentally, the pan-penalty
method is a mixed static-kinematic dual approach to plastic limit analysis. The motivation of this paper is to
develop a purely kinematic approach for plastic limit analysis, and therefore the corresponding direct iterative
algorithm can be regarded as a purely kinematic solution. If the static theorem of limit analysis is introduced
into the analysis, a similar solution to Gaos will be obtained.

6. Applications
Soil is a typical kind of non-associated plastic-flow material. Due to the theoretical complexity and numerical
difficulties in non-associated flow analysis, most of limit analyses of soil materials are still based on associated
plastic flow. The purpose of this section will give some preliminary investigation about how non-associated
plastic flow affects the plastic limit load of a structure.

6.1. Stability of a wedge


The first numerical application of the developed method is for the stability analysis of a simple wedge under
uniform pressure, as shown in Figure 1. The plane-strain condition is assumed. The weightless soil material
obeys the MohrCoulomb yield criterion. When the wedge angle = 0 , the solution becomes the stability
analysis of a half space under uniform pressure. First, the numerical simulation is implemented for = 0
and the associated plastic flow ( = ) using the developed approach. The numerical results for this case are
presented in Figure 2, where ps denotes the limit pressure. The numerical results agree well with Prandtls exact
solution [78]. In order to investigate how non-associated plastic flow affects bearing capacity of a structure,
more numerical simulations are implemented for different dilation angles ( = /2 and = 0 ), and the
numerical results are presented in Figure 3. It can be seen in Figure 3 that non-associated plastic flow brings
a decreased collapse load. The plastic flow behaviour has a significant effect on the bearing capacity of a
structure, and the classic associated plastic flow may overestimate the limit load.
More numerical simulations are implemented for a wedge with the angle = 60 . Due to the symmetry, it
is only necessary to simulate half of the wedge, and the finite element mesh is shown in Figure 1. Numerical
results for both associated and non-associated plastic flow are presented in Figure 4, which further shows that
the bearing capacity of a structure drops when the degree of non-associated plastic flow increases.

6.2. A layered thick-walled cylinder under internal pressure


Components in engineering structures, such as pipes or tunnels, can be simplified as thick-walled cylinders.
In this section, the limit load of a two-layered, thick-walled cylinder under internal pressure is calculated by
applying the proposed approach. The size of the simulated cylinder here is defined as: the inner radius of the
cylinder R0 = 2, the outer radius of the first layer R1 = 4, and the outer radius of the second layer R2 = 5 in
an arbitrary unit. The outer layer of the two-layered cylinder is made of an anisotropic material that is assumed
to obey the Hill yield criterion with associated plastic flow. The inner layer of the two-layered cylinder is made
of a pressure-dependent material that here is assumed to obey the DruckerPrager yield criterion with nonassociated plastic flow. The plane-strain model is adopted for the numerical simulation. In the plane-strain
model, the strength parameters 0 and c0 in the DruckerPrager criterion can be expressed by the friction angle
and cohesion c in the MohrCoulomb criterion as follows
0 = 

tg
9 + 12tg2

3c
c0 = 
9 + 12tg2
The materials of the two layers of the cylinder are chosen for the numerical simulation as follows:

(45)

(46)

Li

535

Figure 1. The finite element mesh of a half wedge under uniform pressure

Figure 2. The limit loads of the wedge under associated plastic flow ( = 0 )

1)
2)

the inner, the DruckerPrager material: the cohesion c with various friction angles ( = 0 45 )
the outer, Hills strength material in terms of the cohesion c of the inner material: rr =4.0c, =3.2c,
zz =5.0c, and r =2.5c.

Due to the symmetry of the structure and force, only one quarter of the structure is to be numerically analyzed.
The finite element mesh of one-quarter of the cylinder is plotted, as shown in Figure 5. The convergence
tolerances in the numerical simulation are chosen as 1 = 2 = 103 . Using the developed approach, the

536

Mathematics and Mechanics of Solids 18(5)

Figure 3. The limit loads of the wedge under non-associated flow ( = 0 )

Figure 4. The limit loads of the wedge ( = 60 )

limit load of this cylinder subjected to internal pressure is numerically calculated, and the numerical results are
shown in Figure 6. The numerical results further prove that the type of plastic flow has a dramatic effect on the
bearing capacity, and non-associated plastic flow may have a smaller limit load than associated plastic flow.
Figures 7 and 8 further show how the non-associated plastic flow affects the bearing capacity and limit
analysis, where the degree of non-associated flow is defined by d=(-)/. As the degree of non-associated
plastic flow increases, the bearing capacity of structures dramatically decreases.

Li

537

Figure 5. The finite element mesh of one-quarter of the cylinder

Figure 6. The limit loads of the cylinder under internal pressure

7. Conclusion
A general nonlinear approach is proposed to implement kinematic limit analysis for materials obeying nonassociated plastic flow. The dissipation power based on a general yield criterion with non-associated plastic

538

Mathematics and Mechanics of Solids 18(5)

Figure 7. Effect of non-associated plastic flow on limit loads

Figure 8. Effect of non-associated plastic flow on limit loads

flow is explicitly expressed in terms of the kinematic admissible velocity, and the nonlinear yield surface is
directly introduced into the kinematic limit theorem. Based on the mathematical programming theory, the finite
element model of the kinematic limit analysis is finally formulated as a nonlinear programming problem subject
to only one equality constraint. Therefore, the computational effort of the proposed approach is very modest.
The extended direct iterative algorithm can solve the resulting nonlinear programming problem. The numerical
results reveal that the associated plastic-flow rule may overestimate the bearing capacity of structures and a
non-associated plastic-flow rule is necessary to calculate the plastic limit load of materials more accurately.
The proposed approach provides a general methodology for the application of the classic limit theorem to a
material with different plastic behaviours.

Li

539

It should be emphasized that for the MohrCoulomb or DruckerPrager types of soil materials, the nonassociated plastic-flow function should have the same form as the yield function but with a dilation angle in
the plastic-flow rule less than the friction angle in the yield criterion. The full conditions of the non-associated
plastic flow, as stated in the Radenkovic theorems, must be satisfied, and then the limit load of the developed
approach in this paper can be regarded as a kinematic solution to the collapse load of structures.
However, it should be also emphasized that non-associated plastic flow violates the normality condition,
and this may bring some theoretical difficulties in the implementation and application of limit analysis for these
kinds of materials. Due to the lack of extreme theorems for limit analysis of any non-standard materials, it is
not possible to theoretically prove that the kinematic limit analysis can still provide a strict upper bound to the
true plastic limit load for non-associated plastic-flow materials. Nevertheless, based on the sufficient bounding
conditions, the kinematic solution of limit analysis presented in this paper provides a quasi-upper bound to the
true collapse load of structures.
Uniqueness of solution for boundary-value problems is a fundamental issue in plasticity analysis for nonassociated plastic-flow materials. A pioneering work was presented by Hill [79], who first formulated a rigorous
theory of the uniqueness and stability condition for standard materials. Later, Maier [80] proposed a weak,
yet sufficient, condition for the uniqueness of non-associated flow and working-hardening/softening materials,
which involved Greens function of the infinitesimal elasticity theory. Raniecki and Bruhns [81] defined a oneparameter family of linear comparison solids and proposed a sufficient uniqueness condition for non-associated
plastic-flow materials. All of these fundamental works provide insight into uniqueness and the stability condition of different plastic-flow materials. However, it is still hard to theoretically and rigorously prove the general
condition of uniqueness for non-associated plastic-flow materials. This paper is based on the strict sufficient
condition of uniqueness for non-associated plastic flow as defined by the Radenkovic theorems. Palmer [69]
also proposed a similar approach to construct a non-associated plastic potential so that a sufficient condition of
uniqueness can be obtained.
The non-associated plastic-flow law used in this paper is actually a three-parameter (cohesion, friction angle,
and dilation angle) material model, which is frequently used for soil materials. The yield function and the
plastic-potential function are separately expressed. However, according to thermodynamic principles, some
constraints are required to be imposed on the intrinsic relation between these two functions. More details
about this analysis can be found in the works of Palmer, Maier, Raniecki and Bruhns, and Mroz [69, 8082].
These necessary conditions also construct the basis of the Radenkovic theorems. It should be noted that there
exist other kinds of non-standard plastic-flow laws, e.g. the fissure model proposed by Rudnicki and Rice [83]
for dilatant materials. This is actually a non-coaxial plastic-flow theory and can be regarded as one kind of
generalized non-associated plastic-flow rule. In this fissure model, two modules were adopted, one of which
was the traditional plastic-hardening modulus governing straight ahead stressing and another of which was a
hardening modulus governing the tangential response to the stress increment. Then, the full plastic flow can be
decomposed into two partsthe coaxial and non-coaxial plastic flow. The methodology developed in this paper
for implementing kinematic limit analysis for non-associated flow materials can be extended for Rudnicki and
Rice non-coaxial plastic-flow materials, and this will be presented in future work.
Funding
This research received no specific grant from any funding agency in the public, commercial, or not-for-profit sectors.

Conflict of interest
None declared.

References
[1]
[2]

Hill, R. A variational principle of maximum plastic work in classical plasticity. Q J Mech Appl Math 1948; 1: 1828.
Hill, R. On the state of stresses in a plastic-rigid body at the yield point. Phil Mag 1951; 42: 868875.

540
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]

Mathematics and Mechanics of Solids 18(5)


Drucker, DC, Prager, W, and Greenberg, HJ. Extended limit design theorems for continuous media. Q Appl Math 1952; 9:
381389.
Drucker, DC. Limit analysis of two- and three-dimensional soil mechanics problems. J Mech Phys Solids 1953; 1: 217226.
Koopman, DCA, and Lance, RH. On linear programming and plastic limit analysis. J Mech Phys Solids 1965; 13: 7787.
Maier, G, Giacomini, S, and Paterlini, F. Combined elastoplastic and limit analysis via restricted basis linear programming.
Comput Meth Appl M 1979; 19: 2148.
Cristiansen, E. Computation of limit loads. Int J Numer Meth Eng 1981; 17: 15471570.
Pastor, J, and Turgeman, S. Limit analysis in axisymmetrical problems: numerical determination of complete statistical solutions.
Int J Mech Sci 1982; 24: 95117.
Pellegrino, S. Efficient upper-bound limit analysis. Int J Mech Sci 1988; 30: 455474.
Gao, Y. On the complementary bounding theorems for limit analysis. Int J Solid Struct 1988, 24: 545556.
Zouain, N, Herskovits, J, Borges, LA, et al. An iterative algorithm for limit analysis with nonlinear yield functions. Int J Solid
Struct 1993; 30: 13971417.
Zhang, YG, and Lu, MW. An algorithm for plastic limit analysis. Comput Meth Appl M 1995; 126: 333341.
Liu, YH, Cen, ZZ, and Xu, BY. A numerical method for plastic limit analysis of 3-D structures. Int J Solid Struct 1995, 32:
16451658.
Capsoni, A, and Corradi, L. A finite element formulation of the rigid-plastic limit analysis problem. Int J Numer Meth Eng 1997,
40: 20632086.
Li, HX, Liu, YH, Feng, XQ, et al. Limit analysis of ductile composites based on homogenization theory. Proc R Soc Lond A
2003, 459: 659675.
Mackenzie, D, Shi, J, and Boyle, JT. Finite element modelling for limit analysis by the elastic compensation method. Comput
Struct 1994; 51: 403410.
Mackenzie, D, Boyle, JT, and Hamilton, R. The elastic compensation method of limit and shakedown analysis: a review. J Anal
Eng Des 2000; 35: 171188.
Ponter, ARS, and Carter, KF. Limit state solutions based upon linear elastic solutions with a spatially varying elastic modulus.
Comput Meth Appl M 1997; 140: 237258.
Ponter, ARS, Fuschi, P, and Engelhardt, M. Limit analysis for a general class of yield conditions. Eur J Mech A Solids 2000; 19:
401421.
Chen, L, Liu, Y, Yang, P, et al. Limit analysis of structures containing flaws based on a modified elastic compensation method.
Eur J Mech A Solids 2008; 27: 195209.
Pastor, J, Thai, T, and Francescato, P. Interior point optimization and limit analysis: an application. Commun Num Meth Eng
2003; 19: 779785.
Pastor, F, Loute, E, and Pastor, J. Limit analysis and convex programming: a decomposition approach of the kinematic mixed
method. Int J Numer Meth Eng 2009; 78: 254274.
Li, HX, and Yu, HS. Kinematic limit analysis of frictional materials using nonlinear programming. Int J Solid Struct 2005; 42:
40584076.
Makrodimopoulos, A, and Martin, CM. Lower bound limit analysis of cohesive-frictional materials using second-order cone
programming. Int J Numer Meth Eng 2006; 66: 604634.
Makrodimopoulos, A, and Martin, CM. Upper bound limit analysis using simplex strain elements and second-order cone
programming. Int J Numer Anal Methods Geomech 2007; 31: 835865.
Kao, JS, Mura, T, and Lee, SL. Limit analysis of orthotropic plates. J Mech Phys Solids 1963; 11: 429436.
Zhang, YG, and Lu, MW. Computational limit analysis of anisotropic axisymmetric shells. Int. J. Pressure Vessels Piping 1994;
58: 283287.
Capsoni, A, Corradi, L, and Vena, P. Limit analysis of orthotropic structures based on Hills yield condition. Int J Solid Struct
2001; 38: 39453963.
Li, HX, Liu, YH, Feng, XQ, et al. Micro/macromechanical plastic limit analyses of composite materials and structures. Acta
Mech Solida Sin 2001; 14: 323333.
Corradi, L, and Vena, P. Limit analysis of orthotropic plates. Int J Plast 2003; 19: 15431566.
Li, HX, and Yu, HS. Limit analysis of composite materials based on an ellipsoid yield criterion. Int J Plast 2006; 22: 19621987.
Capsoni, A, Corradi, L, and Vena, P. Limit analysis of anisotropic structures based on the kinematic theorem. Int J Plast 2001;
17: 15311549.
Corradi, L, Luzzi, L, and Vena, P. Finite element limit analysis of anisotropic structures. Comput Meth Appl M 2006; 195:
54225436.
Francescato, P, and Pastor, J. Lower and upper numerical bounds to the off-axis strength of unidirectional fiber-reinforced
composite by limit analysis methods. Eur J Mech A Solids 1997; 16: 213234.
Thai, T, Francescato, P, and Pastor, J. Limit analysis of unidirectional porous media. Mech Res Commun 1998; 25: 535542.
Trillat, M, and Pastor, J. Limit analysis and Gursons Model. Eur J Mech A Solids 2005; 24: 800819.
Pastor, F, Thor, P, Loute, E, et al. Convex optimization and limit analysis: application to Gurson and porous DruckerPrager
materials. Eng Fract Mech. 2008; 75: 13671383.

Li
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]
[65]
[66]
[67]
[68]
[69]
[70]
[71]

541
Pastor, F, Loute, E, Pastor, J, et al. Mixed method and convex optimization for limit analysis of homogeneous Gurson materials:
a kinematical approach. Eur J Mech A Solids 2009; 28: 2535.
Dallot, J, and Sab, K. Limit analysis of multi-layered plates, part I: the homogenized LoveKirchhoff model. J Mech Phys Solids
2008; 56: 561580.
Dallot, J, and Sab, K. Limit analysis of multi-layered plates, part II: shear effects. J Mech Phys Solids 2008; 56: 581612.
Onat, ET, and Prager, W. Limit analysis of arches. J Mech Phys Solids 1953; 1: 7789.
Cocchetti, G, and Maier, G. Elastic-plastic and limit-state analyses of frames with softening plastic-hinge models by mathematical
programming. Int J Solids Struct 2003; 40: 72197244.
Tin-Loi, F. A GAMS model for the plastic limit analysis of plane frames. Appl Math Modelling 1993; 17: 595602.
Tin-Loi, F, Tangaramvong, S, and Xia, SH. Limit analysis of frames involving unilateral supports with frictional contact. Int J
Mech Sci 2007; 49: 454465.
Tin-Loi, F. Plastic limit analysis of plane frames and grids using GAMS. Comput Struct 1995; 54: 1525.
Tangaramvong, S, and Tin-Loi, F. Limit analysis of elastoplastic frames considering 2nd-order geometric nonlinearity and
displacement constrains. Int J Mech Sci 2009; 51: 179191.
Shi, J, Boyle, JT, Mackenzie, D, et al. Approximate limit design of frames using elastic analysis. Comput Struct 1996; 61:
495501.
Shen, WQ. Limit analyses of plane frames with a penalty linear programming method. Comput Struct 1995, 56: 687695.
Milani, G, Milani, E, and Tralli, A. Upper-bound limit analysis model for FRP-reinforced masonry curved structures, part I:
unreinforced masonry failure surfaces. Comput Struct 2009; 87: 15161533.
Milani, G, Milani, E, and Tralli, A. Upper-bound limit analysis model for FRP-reinforced masonry curved structures, part II:
structural analyses. Comput Struct 2009; 87: 15341558.
Cavicchi, A, and Gambarotta, L. Two-dimensional finite element upper-bound limit analysis of masonry bridges. Comput Struct
2006; 84: 23162328.
Block, P, Ciblac, T, and Ochsendorf, J. Real-time limit analysis of vaulted masonry buildings. Comput Struct 2006; 84: 1841
1852.
Gilbert, M, Casapulla, C, and Ahmed, HM. Limit analysis of masonry block structures with non-associative frictional joints using
linear programming. Comp Struct 2006; 84: 873887.
Milani, G, Loureno, PB, and Tralli, A. Homogenised limit analysis of masonry walls, part I: failure surfaces. Comput Struct
2006; 84: 166180.
Milani, G, Loureno, PB, and Tralli, A. Homogenised limit analysis of masonry walls, part II: structural examples. Comput Struct
2006; 84: 181195.
Milani, G, and Loureno, PB. A simplified homogenized limit analysis model for randomly assembled blocks out-of-plane loaded.
Comput Struct 2010; 88: 690717.
Mihai, LA. A fixed-point approach to the limit load analysis of multibody structures with Coulomb friction. Comput Struct 2010;
88: 859869.
Leu, SY. Convergence analysis and validation of sequential limit analysis of plane-strain problems of the von Mises model with
nonlinear isotropic hardening. Int J Numer Meth Eng 2005, 64: 322334.
Leu, SY. Analytical and numerical investigation of strain-hardening viscoplastic thick-walled cylinders under internal pressure
by using sequential limit analysis. Comput Meth Appl M 2007; 196: 27132722.
Leu, SY. Limit analysis of strain-hardening viscoplastic cylinders under internal pressure by using the velocity control: analytical
and numerical investigation. Int J Mech Sci 2008; 50: 15781585.
Leu, SY. Investigation of rotating hollow cylinders of strain-hardening viscoplastic materials by sequential limit analysis. Comput
Meth Appl M 2008; 197: 48584865.
Zhang, X, Liu, Y, Zhao, Y, et al. Lower bound limit analysis by the symmetric Galerkin boundary element method and the
complex method. Comput Meth Appl M 2002; 191: 19671982.
Zhang, X, Liu, Y, and Cen, Z. Boundary element methods for lower bound limit and shakedown analysis. Eng Anal Boundary
Elements 2004; 28: 905917.
Chen, S, Liu, Y, and Cen, Z. Lower-bound limit analysis by using the EFG method and nonlinear programming. Int J Numer
Meth Eng 2008; 74: 391415.
Le, C, Gilbert, M, and Askes, H. Limit analysis of plates using the EFG method and second-order cone programming. Int J
Numer Meth Eng 2009; 78: 15321552.
Drucker, DC. Coulomb friction, plasticity, and limit loads. J Appl Mech 1954; 21: 7174.
Radenkovic, D. Thormes limites pour un materiau de Coulomb dilatation non standardise. C R Acad Sci Paris 1961; 252:
41034104.
Josselin de Jong, G. Lower bound collapse theorem and lack of normality of strain rate to yield surface for soils. In: Proceedings
of IUTAM, Symposium: Rheology and Soil Mechanics, Grenoble, France, 1964, pp. 6978.
Palmer, AC. A limit theorem for materials with non-associated flow laws. J Mec 1966; 5: 217222.
Sacchi, G, and Save, MA. A note on the limit loads of nonstandard materials. Meccanica 1968; 3: 4345.
Collins, IF. The upper bound theorem for rigid plastic solids generalised to include Coulomb friction. J Mech Phys Solids 1969;
17: 323338.

542

Mathematics and Mechanics of Solids 18(5)

[72]
[73]

Salencon, J. Applications of the theory of plasticity in soil mechanics. Chichester: John Wiley & Sons, 1977.
De Saxce, G, and Bousshine, L. Limit analysis theorems for implicit standard materials: application to the unilateral contact with
dry friction and the non-associated flow rules in soils and rocks. Int J Mech Sci 1998; 40: 387398.
Lubliner, J. Plasticity theory. New York: Macmillan, 1990.
Bazaraa, MS, Sherali, HD, and Shetty, CM. Nonlinear programming: theory and algorithm. Hoboken, NJ: John Wiley & Sons,
2006.
Nocedal, J, and Wright, SJ. Numerical optimization. New York: Springer, 2006.
Himmelblau, DM. Applied nonlinear programming. New York: McGraw-Hill, 1972.
Prandtl, L. ber die Hrte plastischer Krper. Nachrichten von der Gesellschaft der Wissenschaften zu Gttingen. MathematischPhysikalische Klasse 1920; 12: 7485.
Hill, R. A general theory of uniqueness and stability in elastic-plastic solids. J Mech Phys Solid 1958; 6: 236250.
Maier, G. A minimum principle for incremental elastoplasticity with non-associated flow laws. J Mech Phys Solid 1970; 18:
319330.
Raniecki, B, and Bruhns, OT. Bounds to bifurcation stresses in solids with non-associated plastic flow law at finite strain. J Mech
Phys Solid 1981; 29: 153172.
Mroz, Z. Non-associated flow laws in plasticity. J Mec 1963; 2: 2142.
Rudnicki, JW, and Rice, JR. Conditions for the localization of deformation in pressure-sensitive dilatant materials. J Mech Phys
Solid 1975; 23: 371394.
Fang, SC, Gao, DY, Sheu, RL, et al. Global optimization for a class of fractional programming problems. J Global Optimization
2009; 45: 337353.
Gao, DY. Duality principles in nonconvex systems: theory, methods and applications. Dordrecht, The Netherlands: Kluwer
Academic Publishers, 2000.
Gao, Y. Panpenalty finite element programming for plastic limit analysis. Comput Struct 1998; 28: 749755.

[74]
[75]
[76]
[77]
[78]
[79]
[80]
[81]
[82]
[83]
[84]
[85]
[86]

S-ar putea să vă placă și