02.04.2014 Views

An improved reproducing kernel particle method for nearly ...

An improved reproducing kernel particle method for nearly ...

An improved reproducing kernel particle method for nearly ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

www.elsevier.com/locate/cma<br />

<strong>An</strong> <strong>improved</strong> <strong>reproducing</strong> <strong>kernel</strong> <strong>particle</strong> <strong>method</strong> <strong>for</strong> <strong>nearly</strong><br />

incompressible ®nite elasticity<br />

Jiun-Shyan Chen a, *<br />

, Sangpil Yoon b , Hui-Ping Wang b , Wing Kam Liu c<br />

a Department of Mechanical Engineering & Center <strong>for</strong> Computer-Aided Design, The University of Iowa, 2137 Engineering Building,<br />

Iowa City, IA 52242-1527, USA<br />

b Department of Mechanical Engineering & Center <strong>for</strong> Computer-Aided Design, The University of Iowa, Iowa, USA<br />

c Department of Mechanical Engineering, Northwestern University, 2145 Sheridon Road, Evanston, IL 60208-3111, USA<br />

Received 8 August 1997; received in revised <strong>for</strong>m 8 August 1998<br />

Abstract<br />

The previously developed <strong>reproducing</strong> <strong>kernel</strong> <strong>particle</strong> <strong>method</strong> (RKPM) employs a high-order quadrature rule <strong>for</strong> desired domain<br />

integration accuracy. This leads to an over-constrained condition in the limit of incompressibility, and volumetric locking and pressure<br />

oscillation were encountered. The employment of a large support size in the <strong>reproducing</strong> <strong>kernel</strong> shape function increases the dependency<br />

in the discrete constraint equations at quadrature points and thereby relieves locking. However, this approach consumes high<br />

CPU and it cannot e€ectively resolve pressure oscillation diculty. In this paper, a pressure projection <strong>method</strong> is introduced by locally<br />

projecting the pressure onto a lower-order space to reduce the number of independent discrete constraint equations. This approach<br />

relieves the over-constrained condition and thus eliminates volumetric locking and pressure oscillation without the expense of employing<br />

large support size in RKPM. The <strong>method</strong> is developed in a general framework of <strong>nearly</strong> incompressible ®nite elasticity and<br />

there<strong>for</strong>e linear problems are also applicable. To further reduce the computational cost, a stabilized reduced integration <strong>method</strong> based<br />

on an assumed strain approach on the gradient matrix associated with the de<strong>for</strong>mation gradient is also introduced. The resulting<br />

sti€ness matrix and <strong>for</strong>ce vector of RKPM are obtained explicitly without numerical integration. Ó 2000 Elsevier Science S.A. All<br />

rights reserved.<br />

1. Introduction<br />

The regularity requirement of the interpolation functions and meshes in the ®nite element <strong>method</strong>s<br />

obscures their applications to problems involving large material distortions, crack propagation, and high<br />

gradients, among others. In recent years, a number of meshless <strong>method</strong>s have been proposed to address<br />

these diculties. Among them are the smooth <strong>particle</strong> hydrodynamics (SPH) [1±3], the di€use element<br />

<strong>method</strong> (DEM) [4], the element free Galerkin (EFG) <strong>method</strong> [5±9], the <strong>reproducing</strong> <strong>kernel</strong> <strong>particle</strong> <strong>method</strong><br />

(RKPM) [10±17], the HP cloud <strong>method</strong> [18,19], and the partition of unity <strong>method</strong> (PUM) [20]. In these<br />

<strong>method</strong>s, the approximation of ®eld variables is constructed in terms of nodes. The actual implementation<br />

of some of the meshless <strong>method</strong>s, in fact, requires the partition of the domain through the use of a<br />

``background grid'' <strong>for</strong> domain integration, and this leads to some speculation on whether the <strong>method</strong>s are<br />

truly ``meshless.'' Nevertheless, due to the ¯exibility in constructing the con<strong>for</strong>ming shape functions to meet<br />

speci®c needs <strong>for</strong> di€erent applications, it has been reported [5±20] that the meshless <strong>method</strong>s are particularly<br />

suitable <strong>for</strong> crack propagation, hp-adaptivity, multiple-resolution, and large de<strong>for</strong>mation problems.<br />

* Corresponding author.<br />

0045-7825/00/$ - see front matter Ó 2000 Elsevier Science S.A. All rights reserved.<br />

PII: S 0 0 4 5 - 7 8 2 5 ( 9 9 ) 0 0 067-5


118 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

One of the major disadvantages of the meshless <strong>method</strong>s is its relatively high computational cost.<br />

The supports of the meshless shape functions usually cover more surrounding points than those in the<br />

®nite element <strong>method</strong>s. This requirement increases the number of numerical operations in the sti€ness<br />

matrix <strong>for</strong>mation and assembly, and the resulting global sti€ness matrix has large bandwidth. The<br />

situation is more signi®cant when dealing with incompressible problems, in which suciently large<br />

support sizes need to be used in the meshless shape functions to avoid incompressible locking<br />

[5,6,15,16]. In meshless <strong>for</strong>mulation, higher-order weight functions and <strong>kernel</strong> functions such as the<br />

exponential function, Gauss function, and the cubic B-spline function are usually used, and there<strong>for</strong>e<br />

higher-order<br />

p p<br />

Gauss<br />

<br />

integration is required. Belytschko et al. [5,6] suggested an integration order of<br />

… m ‡ 2† … m ‡ 2†, with m the number of points within the integration zone <strong>for</strong> two-dimensional<br />

problems using an exponential typed weight function. For a two-dimensional integration zone containing<br />

four points, <strong>for</strong> example, meshless <strong>method</strong>s require 4 4 quadrature points, whereas only 2 2<br />

quadrature points are needed in 4-node ®nite elements. Other numerical operations that consume additional<br />

CPU time in meshless computation are the imposition of the essential boundary conditions by<br />

the Lagrangian multiplier <strong>method</strong> [5] or the direct trans<strong>for</strong>mation <strong>method</strong> [15,16], and the construction<br />

of meshless shape functions.<br />

The incompressible locking in ®nite elements has been studied extensively and various <strong>method</strong>s have<br />

been proposed to resolve this diculty resulting from the over-constrained nature in the displacementbased<br />

®nite element <strong>method</strong>s. Among them are the Lagrange-multiplier-type mixed <strong>method</strong>s [21±28], the<br />

selective reduced integration <strong>method</strong> [29,30], the hourglass control <strong>method</strong> on under-integrated elements<br />

[31±33], the Taylor expansion <strong>method</strong> [34±36], the volumetric strain projection <strong>method</strong> [37], the pressure<br />

projection <strong>method</strong> [38±40], and the rank-one ®ltering <strong>method</strong> [41]. In the pressure projection <strong>method</strong><br />

[38,39], the displacement-calculated pressure is projected onto a lower-order pressure space by a leastsquares<br />

projection at the element level. The resulting equilibrium equation is equivalent to the perturbed<br />

Lagrangian <strong>for</strong>mulation [27,28]. With an appropriate decomposition of the strain energy density function,<br />

and a selection of particular pressure interpolation functions, a fairly simple pressure projection <strong>for</strong>mulation<br />

was obtained and it can be easily implemented into displacement-based ®nite element code. The<br />

degeneration of the pressure projection <strong>method</strong> to linear elasticity leads to the well known mean dilatation<br />

[42] and selective reduced integration [30] <strong>method</strong>s. The Taylor expansion approach is based on the assumed<br />

strain <strong>method</strong> [34±36], in which the assumed strain ®eld is obtained by the Taylor expansion of the<br />

displacement gradient matrix. The resulting sti€ness matrix and <strong>for</strong>ce vector can be expressed explicitly by<br />

one-point quadrature terms and their stabilization.<br />

In this paper, it is shown that the fully integrated RKPM is severely over-constrained in <strong>nearly</strong> incompressible<br />

problems, and a pressure projection <strong>method</strong> is developed to reduce the independent discrete<br />

constrained equations. A stabilized reduced integration <strong>method</strong> is also introduced to further reduce the<br />

computational e€ort in RKPM.<br />

The meshless shape functions developed from RKPM are reviewed in Section 2. Section 3 discusses the<br />

pressure projection <strong>method</strong> <strong>for</strong> RKPM ®nite elastic <strong>for</strong>mulation. In Section 4, the RKPM <strong>for</strong>mulation is<br />

further simpli®ed by introducing a Taylor series expansion of the gradient matrix on the selected terms of<br />

sti€ness matrix and <strong>for</strong>ce vector. Several linear and nonlinear problems are analyzed to verify the per<strong>for</strong>mance<br />

of the proposed <strong>method</strong> in Section 5, and conclusions are given in Section 6.<br />

2. Overview of <strong>reproducing</strong> <strong>kernel</strong> <strong>particle</strong> <strong>method</strong><br />

2.1. Reproducing <strong>kernel</strong> approximation<br />

With no loss of generality, in this section we shall use the following notations: x ‰x 1 ; x 2 ; x 3 Š,<br />

y ‰y 1 ; y 2 ; y 3 Š, and dy dy 1 dy 2 dy 3 . The <strong>kernel</strong> estimate was ®rst introduced to the SPH [1±3], in which the<br />

<strong>kernel</strong> estimate of a function u…x† is<br />

Z<br />

u K …x† ˆ U a …x y†u…y† dy;<br />

…1†<br />

X


where u K …x† is the <strong>kernel</strong> estimate of u…x†, and U a …x y† the <strong>kernel</strong> function with support measure of a.<br />

The <strong>kernel</strong> function U a …x y† is a positive function with the following properties:<br />

Z<br />

U a …x y† dy ˆ 1;<br />

…2†<br />

X<br />

u K …x† ! u…x† as a ! 0: …3†<br />

The imposition of Eq. (2) is in fact the zeroth order consistency condition. Eq. (2), however, does not assure<br />

the consistency condition in the discrete <strong>for</strong>m. To further restore the higher-order consistency conditions of<br />

the <strong>kernel</strong> estimate, Liu et al. [10] proposed a <strong>reproducing</strong> <strong>kernel</strong> approximation by introducing a correction<br />

function to the <strong>kernel</strong> estimate<br />

Z<br />

u R …x† ˆ C…x; x y†U a …x y†u…y† dy;<br />

…4†<br />

X<br />

J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 119<br />

where u R …x† is the ``reproduced'' function of u…x†, and Eq. (4) is called the <strong>reproducing</strong> equation. The<br />

function C…x; x y† is the correction function de®ned by<br />

C…x; x y† ˆ b T …x†H…x y†;<br />

…5†<br />

where H…x y† is the vector of monomial basis functions,<br />

H T …x y† ˆ ‰1; x 1 y 1 ; x 2 y 2 ; x 3 y 3 ; …x 1 y 1 † 2 ; . . . ; …x 3 y 3 † N Š;<br />

b T …x† ˆ ‰b 0 …x†; b 1 …x†; . . .Š<br />

…6†<br />

…7†<br />

and b i …x†Õs are determined by imposing the following Nth order completeness requirement, i.e., if u…x† is a<br />

Nth order monomial, then we require<br />

Z<br />

u R …x† ˆ u…x† ˆ C…x; x y†U a …x y†u…y† dy:<br />

…8†<br />

X<br />

If u…x† is an Nth order monomial, one can express u…y† in terms of an Nth order Taylor expansion of x in<br />

Eq. (8) to yield<br />

u…y† ˆ u…x† ‡ ou…x†<br />

ox 1<br />

where<br />

<br />

w…x† T ˆ<br />

u…x†; ou…x†<br />

ox 1<br />

…y 1 x 1 † ‡ ou…x† …y 2 x 2 † ‡ ou…x† …y 3 x 3 † ‡ ˆ w…x† T H…x y†; …9†<br />

ox 2<br />

ox 3<br />

; ou…x†<br />

ox 2<br />

; ou…x† ; <br />

ox 3<br />

<br />

…10†<br />

and one can also express u…x† by<br />

u…x† ˆ w T …x†H…0†:<br />

…11†<br />

Substituting Eqs. (9) and (11) into Eq. (8) yields<br />

w…x† T ‰H…0† M…x†b…x†Š ˆ 0;<br />

where<br />

Z<br />

M…x† ˆ H…x y†H T …x y†U a …x y† dy:<br />

X<br />

Since u…x† is an arbitrary Nth order monomial, Eq. (12) implies<br />

b…x† ˆ M…x† 1 H…0†<br />

…12†<br />

…13†<br />

…14†


120 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

and<br />

C…x; x y† ˆ H T …0†M 1 …x†H…x y†:<br />

…15†<br />

Introducing Eq. (15) into Eq. (4) leads to the following <strong>reproducing</strong> <strong>kernel</strong> approximation<br />

Z<br />

Z<br />

u R …x† ˆ C…x; x y†U a …x; y†u…y† dy ˆ H T …0†M 1 …x† H…x y†U a …x y†u…y† dy:<br />

X<br />

Eq. (16) can be recast into the following <strong>for</strong>m<br />

Z<br />

u R …x† ˆ U a …x; x y†u…y† dy;<br />

X<br />

where U a …x; x y† ˆ C…x; x y†U a …x y† is called the reproduced <strong>kernel</strong>. Since Eq. (17) exactly reproduces<br />

Nth order mononomials, the <strong>method</strong> ful®lls the Nth order consistency conditions, i.e.,<br />

Z<br />

U a …x; x y†y l 1 ym 2 yn 3 dy ˆ xl 1 xm 2 xn 3<br />

<strong>for</strong> l ‡ m ‡ n ˆ 0; . . . ; N: …18†<br />

X<br />

X<br />

…16†<br />

…17†<br />

2.2. Discretization of <strong>reproducing</strong> <strong>kernel</strong> approximation<br />

Suppose that the domain X is discretized by a set of nodes fx 1 ; . . . ; x NP g, where x I is the location of node<br />

I, and NP the total number of points (nodes). By the use of a simple trapezoidal rule, Eq. (17) is discretized<br />

into<br />

u h …x† ˆ XNP<br />

C…x; x x I †U a …x x I † d I DV I ;<br />

Iˆ1<br />

…19†<br />

where U a …x x I † is the multi-dimensional <strong>kernel</strong> function that can be <strong>for</strong>med by taking the product of the<br />

one-dimensional <strong>kernel</strong> functions to yield<br />

U a …x x I † ˆ 1 <br />

/<br />

x <br />

1 x 1I<br />

/ x <br />

2 x 2I<br />

/ x <br />

3 x 3I<br />

; …20†<br />

a 1 a 2 a 3 a 1 a 2 a 3<br />

where a 1 , a 2 , and a 3 are measures of support in x 1 , x 2 , and x 3 directions, respectively, and this <strong>kernel</strong><br />

function has rectangular (or brick) support. <strong>An</strong> alternative multi-dimensional <strong>kernel</strong> function with circular<br />

(or spherical) support is constructed by<br />

U a …x x I † ˆ 1<br />

a /<br />

<br />

kx x I k<br />

a<br />

<br />

: …21†<br />

<strong>An</strong> example of the one-dimensional <strong>kernel</strong> function is a cubic B-spline function:<br />

8<br />

2<br />

4jz<br />

3 j2 ‡ 4jzj 3 <strong>for</strong> 0 6 jzj 6 ><<br />

1 2<br />

/…z† ˆ 4<br />

4jzj<br />

‡ 4jz<br />

3 j2 4 jz<br />

3 j3 1<br />

<strong>for</strong> < jzj<br />

6 1 : …22†<br />

2<br />

>:<br />

0 otherwise<br />

The support of U a …x x I †, x I a 6 x 6 x I ‡ a, is centered at the point x I , and the size of the support is<br />

measured by the parameter a.<br />

In a multi-dimensional case, the determination of a volume associates with each node I, DV I , is not a<br />

straight<strong>for</strong>ward task. Since the main purpose of this discretization is only to develop shape functions, we<br />

can simply set DV I to be unity<br />

u h …x† ˆ XNP<br />

C…x; x x I †U a …x x I † d I :<br />

Iˆ1<br />

…23†


J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 121<br />

Note that the correction function C…x; x x I † was determined previously from the completeness requirement<br />

of a continuous <strong>reproducing</strong> equation of Eq. (8). With the discretization of the <strong>reproducing</strong> equation,<br />

we shall re-impose the completeness requirement on Eq. (23) following the same procedures as were previously<br />

discussed to obtain:<br />

C…x; x x I † ˆ H T …0†M 1 …x†H…x x I †;<br />

…24†<br />

M…x† ˆ XNP<br />

H…x x I †H T …x x I †U a …x x I †:<br />

Iˆ1<br />

Eq. (23) can be expressed in the following <strong>for</strong>m:<br />

u h …x† ˆ XNP<br />

W I …x† d I ;<br />

Iˆ1<br />

W I …x† ˆ C…x; x x I †U a …x x I †:<br />

…25†<br />

…26†<br />

…27†<br />

The function W I …x† is interpreted as the <strong>particle</strong> or meshless shape function of node I, and d I is the<br />

associated coef®cient of approximation. Since W I …x† is constructed based on the monomial basis functions<br />

and the <strong>kernel</strong> function U a …x x I †, the support of W I …x† is the same as the support of U a …x x I †. Further,<br />

the order of differentiability of W I …x† is also identical to that of U a …x x I † as discussed in [15]. Since the<br />

<strong>reproducing</strong> <strong>kernel</strong> shape functions satisfy the discrete completeness requirement, they meet the following<br />

consistency conditions<br />

X NP<br />

Iˆ1<br />

W I …x†x l 1I xm 2I xn 3I ˆ xl 1 xm 2 xn 3<br />

; l ‡ m ‡ n ˆ 0; . . . ; N: …28†<br />

It should be noted that, in general, the shape functions do not bear the Kronecker delta properties, i.e.,<br />

W I …x J † 6ˆ d IJ . There<strong>for</strong>e, <strong>for</strong> general function u…x† that is not a monomial, d I in Eq. (26) is not the nodal<br />

value of u…x†. This leads to some complication in imposing the essential boundary conditions in the<br />

meshless <strong>method</strong>s. Additional development such as the Lagrange multiplier <strong>method</strong> [5], the trans<strong>for</strong>mation<br />

<strong>method</strong> [15,16], or the singular <strong>kernel</strong> <strong>method</strong> [9] is required to impose the essential boundary conditions.<br />

3. Pressure projection<br />

3.1. Fundamental diculties in RKPM discretization of <strong>nearly</strong> incompressible ®nite elasticity<br />

For elastic materials that are <strong>nearly</strong> incompressible, the strain energy density W which can be expressed<br />

by the three invariants of strain, is penalized in the following <strong>for</strong>m [25,27,28]<br />

W …I 1 ; I 2 ; J† ˆ W …I 1 J 2=3 ; I 2 J 4=3 † ‡ k ~w…J†<br />

and the corresponding potential energy is<br />

Z<br />

P ˆ ‰W …I 1 J 2=3 ; I 2 J 4=3 † ‡ k ~w…J†Š dX W ext ;<br />

X X<br />

…29†<br />

…30†<br />

where W is the distortional strain energy density, k ~w…J† the dilatational strain energy density, k is a penalty<br />

to impose incompressibility, J ˆ I 1=2<br />

3<br />

the volume ratio, I 1 ; I 2 ; I 3 are the three invariants of the Green de<strong>for</strong>mation<br />

tensor, X X the domain of the original con®guration of the material, and W ext is the external<br />

energy. The stress-strain relation of the material is<br />

S ij ˆ oW ;<br />

…31†<br />

oE ij


122 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

where S is the second Piola±Kirchho€ stress, and E is the Green Lagrangian strain. By using the relation<br />

between the second Piola±Kirchho€ stress and the Cauchy stress, the pressure (trace of Cauchy stress) is<br />

obtained by [38]<br />

P ˆ k ~w 0 …J†:<br />

…32†<br />

A commonly used ~w function is [38]:<br />

~w…J† ˆ 1<br />

2 …J 1†2 …33†<br />

and the corresponding pressure equation is<br />

P ˆ k…J 1†:<br />

…34†<br />

By introducing the Galerkin approximation of the variation of the potential energy in Eq. (30), and<br />

approximating displacements by the <strong>reproducing</strong> <strong>kernel</strong> shape functions, the RKPM discrete equations can<br />

be obtained by [16]. As reported in [16], however, a volumetric locking exists unless a relatively large<br />

support size of the <strong>reproducing</strong> <strong>kernel</strong> shape function is used in the RKPM discretization.<br />

We introduce the <strong>method</strong> of constraint count [43] to study the ability of RKPM to per<strong>for</strong>m well in <strong>nearly</strong><br />

incompressible media. A discrete constraint ratio r c is de®ned by the ratio between the number of discrete<br />

equilibrium equations N eq and the number of independent constraint equations N c . We want the discrete<br />

constraint ratio r c to go toward the number of equilibrium equations divided by the number of incompressibility<br />

conditions in the continuum system when the number of discrete nodes approaches in®nity.<br />

There<strong>for</strong>e the desired discrete constraint ratio in two-dimension is two. Let us introduce a rectangular<br />

domain partitioned into n integration zones<br />

p<br />

per<br />

<br />

side, and<br />

p <br />

each integration zone contains m nodes as shown<br />

in Fig. 1. Using an integration order of … m ‡ 2† … m ‡ 2† [5], the pressure equation obtained from<br />

Eq. (34) is computed at each integration point x l<br />

p <br />

P…x l † ˆ k…J…x l † 1† <strong>for</strong> l ˆ 1; . . . ; … m ‡ 2† 2 : …35†<br />

As the material behavior goes toward incompressible, <strong>for</strong> a bounded P the following incompressibility<br />

constraints are imposed at the integration points<br />

p <br />

2<br />

J…x l † 1 0 <strong>for</strong> l ˆ 1; . . . ; … m ‡ 2† …36†<br />

Fig. 1. Standard RKPM discretization.


or the linearized <strong>for</strong>m<br />

p <br />

J…x l †Du i;i …x l † 0 ! Du i;i …x l † 0 <strong>for</strong> l ˆ 1; . . . ; … m ‡ 2† 2 : …37†<br />

p <br />

Note that the … m ‡ 2† 2 constraint equations in each integration zone are not necessarily li<strong>nearly</strong> independent.<br />

If dependency exists, <strong>for</strong> a two-dimensional RKPM discretization the discrete constraint ratio is<br />

p<br />

2<br />

N eq 2‰n… m 1† ‡ 1Š<br />

r c ˆ lim ˆ lim<br />

n!1 N c n!1<br />

<br />

n 2 ‰…<br />

p m ‡ 2† 2 N d Š ‡ nb ˆ 2… p 2<br />

m 1†<br />

p ; m P 3; …38†<br />

… m ‡ 2† 2 N d<br />

where N d is the number of dependent incompressibility constraints in one integration zone, and this dependency<br />

is in¯uenced by the order and the support size of <strong>reproducing</strong> <strong>kernel</strong> shape functions. For the<br />

integration zones that are close to the boundary, the number of independent constraints is di€erent from<br />

those of the interior zones, and the di€erence is lumped into nb in Eq. (38). Since nb is only li<strong>nearly</strong><br />

proportional to n, it has no in¯uence on r c . If shape functions are monomials, it is easier to estimate N c<br />

directly. In a four-node ®nite element discretization <strong>for</strong> example, each integration point is covered by only<br />

four bilinear shape functions. Since 1; x; y; xy are independent monomials, the derivatives of displacement<br />

are represented by linear functions, and hence the total of independent constraints is N c ˆ 3n 2 , and this leads<br />

to a discrete constraint ratio of<br />

r c ˆ 2 < 2: …39†<br />

3<br />

This constraint ratio <strong>for</strong> a bilinear ®nite element is solely related to the order of bilinear ®nite element<br />

shape functions unless a one-point quadrature rule is used, which increases the constraint ratio to r c ˆ 2<br />

since N c ˆ n 2 .<br />

In RKPM discretization, the shape functions are generally not monomials. The number of independent<br />

constraints is then determined by the order of quadrature rule and the number of shape functions that cover<br />

each integration point. The number of shape functions that cover each integration point is directly related<br />

to the support size of shape functions. In a typical domain partitioning using a 4-node integration zone with<br />

4 4 integration order, the constraint ratio is:<br />

2<br />

r c ˆ ; …40†<br />

16 N d<br />

where N d is related to the support size. One can see that if no dependency in the incompressibility constraints<br />

exists, N d ˆ 0 results in a very low discrete constraint ratio<br />

r c ˆ 1 2: …41†<br />

8<br />

This causes a severe locking and a pressure oscillation, and there<strong>for</strong>e an increase of dependency in the<br />

constraint equations is necessary in RKPM discretization. One way to increase the dependency in the<br />

constraint equations is to enlarge the support size of <strong>reproducing</strong> <strong>kernel</strong> shape functions. As the <strong>kernel</strong><br />

support size is enlarged, the number of <strong>reproducing</strong> <strong>kernel</strong> shape functions that cover each integration zone<br />

increases proportionally. This results in an increase of N d , and consequently r c increases and locking is<br />

reduced. This approach, however, signi®cantly raises the computational e€ort. One other approach is to let<br />

m ! 1 to yield lim m!1 r c ˆ 2 according to Eq. (38), but the <strong>method</strong> is practically useless.<br />

3.2. Pressure projection<br />

J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 123<br />

In this work, we reduce the independent constraint equations in each integration zone by projecting<br />

displacement calculated pressure (Eq. (34)) onto a lower-order space using a pressure projection <strong>method</strong><br />

proposed by Chen et al. [38]. The lower-order approximation of pressure is per<strong>for</strong>med locally by introducing<br />

a least-squares based projection on the displacement calculated pressure onto a pressure space<br />

spanned by a set of functions Q ˆ fQ 1 …x†; Q 2 …x†; . . . ; Q n …x†g within each zone of integration X s X<br />

. That is,<br />

determine p s ˆ ‰p1 s; ps 2 ; . . . ; ps n ŠT to minimize


124 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

w…p s † ˆ kk…J<br />

1† Qp s k 2 L 2 …X s X<br />

†; …42†<br />

where kk 2 L 2 …X s X † is the L 2 norm in the integration zone X s X<br />

. Note that the selection of pressure basis functions<br />

p<br />

Q should be based on Eq. (38) such that N c ! n 2 … m 1† 2 . The minimization of w…p s † leads to<br />

M s p s ˆ Z s ;<br />

where<br />

Z<br />

M s ˆ Q T Q dX;<br />

X s X<br />

Z<br />

Z s ˆ Q T k…J 1† dX<br />

X s X<br />

…43†<br />

…44†<br />

…45†<br />

and the projected hydrostatic pressure P s in X s X is<br />

P s Qp s ˆ QM s1 Z s :<br />

…46†<br />

In Eq. (46), the displacement calculated pressure k…J 1† is projected onto a space spanfQ i …x†g n iˆ1 .<br />

3.3. Incremental equation<br />

The variation of the potential energy in ®nite elasticity according to Eq. (30) is<br />

Z<br />

Z<br />

dP ˆ dF ij r ij dX ‡ dF ij ~r ij …P† dX dW ext<br />

X X X X<br />

…47†<br />

and the linearization is<br />

Z<br />

DdP ˆ dF ij ‰ T ijkl ‡ T ~ ijkl …P†ŠDF kl dX ‡<br />

X X<br />

Z<br />

Z<br />

dF ij ‰ D ijkl ‡ ~D 1 ijkl …P†ŠDF kl dX<br />

X X<br />

‡ dF ij F 1<br />

ji<br />

JDP dX DdW ext ;<br />

X X<br />

…48†<br />

where<br />

r ij ˆ o W<br />

oF ij<br />

ˆ F mi<br />

S mj ;<br />

S ij ˆ o W<br />

oE ij<br />

;<br />

…49†<br />

~r ij ˆ k o ~w<br />

oF ij<br />

ˆ F mi<br />

~S mj ; ~S ij ˆ k o ~w<br />

oE ij<br />

;<br />

…50†<br />

DP ˆ kJF 1<br />

lk DF kl; …51†<br />

where F is the de<strong>for</strong>mation gradient, r and ~r are the distortional and dilatational parts of ®rst Piola±<br />

Kirchho€ stress, respectively, and S and S ~ are the distortional and dilatational parts of the second Piola±<br />

Kirchho€ stress, respectively. The explicit expressions of S ij ; ~S ij ; D ijkl ; ~D 1 ijkl ; ~D 2 ijkl ; T ijkl ; ~T ijkl are listed in<br />

Appendix A.<br />

For the Galerkin approximation of Eq. (48), we further consider a local least-squares based projection of<br />

the pressure increment DP ˆ kJFkl<br />

1 DF kl in each integration zone X s X , by determining ^ps ˆ ‰^p<br />

1 s; ^ps 2 ; . . . ; ^ps n ŠT to<br />

minimize<br />

<br />

^w…^p s † ˆ kJF 1 Q^p s <br />

2<br />

…52†<br />

lk<br />

DF kl<br />

L 2 …X s X † :


The minimization of ^W…^p s † leads to<br />

J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 125<br />

^p s ˆ kM s1 L s Dd s ;<br />

where<br />

Z<br />

L s ˆ Q T gB dX;<br />

X s X<br />

…53†<br />

…54†<br />

g is the row vector <strong>for</strong>m of JFij<br />

1<br />

increment, DP s , is obtained by<br />

and B the gradient matrix of de<strong>for</strong>mation gradient. The projected pressure<br />

DP s Q^p s ˆ QM s1 L s Dd s :<br />

…55†<br />

By substituting Eqs. (46) and (55) into Eqs. (47) and (48) and introducing the RKPM displacement shape<br />

functions and domain discretization, the following incremental equation is obtained<br />

… K ‡ ~ K ‡ ~ K † m‡1<br />

n‡1 Dd ˆ … f ext ‡ ~ f ext † n‡1<br />

…f int † m n‡1 ; …56†<br />

where n is the load step counter and m the iteration counter. The stiffness term K IJ in Eq. (56) is related to<br />

the distortional energy that is independent to the pressure projection,<br />

Z<br />

K s IJ ˆ B T I … T ‡ D†B J dX;<br />

…57†<br />

X s X<br />

where the notation …† s IJ<br />

denotes the sti€ness matrices associated with <strong>particle</strong> I; J integrate over the integration<br />

zone X s X . The sti€ness matrix K ~ in Eq. (56) is related to the dilatational energy with pressure<br />

projected onto spanfQ i …x†g n iˆ1<br />

Z<br />

… K ~ † s IJ ˆ B T I ‰~ T…P s † ‡ D ~ 1 …P s †ŠB J dX:<br />

…58†<br />

X s X<br />

This sti€ness matrix contains the projected pressure. The third sti€ness term in Eq. (56) is the dilatational<br />

sti€ness associated with the projection of pressure increment:<br />

… ~ K † s IJ ˆ kLsT I<br />

M s1 L s J ;<br />

Z<br />

L s I ˆ<br />

X s X<br />

Q T gB I dX:<br />

…59†<br />

…60†<br />

Similarly, the internal <strong>for</strong>ce vector is also separated into f int and f ~ int . The internal <strong>for</strong>ce vector f int is related<br />

to the distortional energy that is independent of the pressure<br />

Z<br />

… f int † s I ˆ r dX;<br />

…61†<br />

X s X<br />

B T I<br />

where …† s I<br />

denotes the <strong>for</strong>ce vector associated with <strong>particle</strong> I integrate over the integration zone X s X . The<br />

second <strong>for</strong>ce vector in Eq. (56) is related to the dilatational energy<br />

Z<br />

… f ~ int † s I ˆ B T I ~r…P s † dX:<br />

…62†<br />

X s X<br />

This internal <strong>for</strong>ce vector contains the projected pressure. The matrices T, T, ~ D, D ~ 1 and Bdd are the matrix<br />

<strong>for</strong>ms of d ik<br />

S jl , d ik<br />

~S jl , D ijkl , ~D 1 ijkl and dF ij, respectively, r and ~r are the vector <strong>for</strong>m of r ij and ~r ij . The<br />

functions ~r ij , ~T ijkl and ~D 1 ijkl<br />

contain pressure that is calculated using the pressure projection equations. Note<br />

that one can also consider an incremental pressure equation by the linearization of Eq. (43). In this case, a<br />

pressure residual needs to be added consistently in Eqs. (53), (55) and (56).


126 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

3.4. Pressure projection <strong>for</strong> the most severe over-constrained condition<br />

Eq. (38) indicates that the most severe over-constrained condition in RKPM discretization is when only<br />

three or four nodes are included in each integration zone and when small <strong>kernel</strong> support size is used<br />

(N d 0). Under this condition, any pressure ®eld higher than a constant ®eld results in a very low discrete<br />

constraint ratio. For instance, a linear pressure ®eld leads to r c ˆ 1=3 and 2=3 <strong>for</strong> m ˆ 3 and 4, respectively.<br />

A basic 3-node integration unit X s X with centroid X s is shown in Fig. 2. Within each basic integration<br />

unit, the pressure is projected onto a constant ®eld, and the pressure projection equations reduce to<br />

P s ˆ k R …J 1† dX<br />

X s X<br />

; …63†<br />

A s<br />

DP s ˆ k R gB dX<br />

X s X<br />

Dd; …64†<br />

A s<br />

where A s is the area of X s X<br />

. Note that with constant pressure projection, the constraint count <strong>for</strong> a 3-node<br />

integration zone increases from 1/3 to 1, and <strong>for</strong> a 4-node integration zone it increases from 2/3 to 2. Using<br />

Eq. (64), the dilatational sti€ness matrix K ~ becomes<br />

Z ! Z !<br />

… K ~ † s IJ ˆ B T kAs1<br />

I gT dX gB J dX : …65†<br />

X s X<br />

X s X<br />

If one further employs one-point integration to <strong>for</strong>m Z s and L s , then the pressure projection equations<br />

reduce to following <strong>for</strong>m:<br />

P s ˆ k‰J…X s † 1Š;<br />

DP s ˆ k…gB†j …X s ;Y s † Dd:<br />

…66†<br />

…67†<br />

Note that all the <strong>reproducing</strong> <strong>kernel</strong> shape functions that cover X s ˆ …X s ; Y s †, both inside and outside of<br />

X s X , contribute to the calculation of projected pressure P s . The dilatational sti€ness K also reduces to the<br />

following <strong>for</strong>m<br />

Fig. 2. Pressure projection location in a triangular integration zone.


J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 127<br />

… ~ K † s IJ ˆ kAs …B T I gT gB J † <br />

…X s ;Y s † :<br />

…68†<br />

3.5. Degeneration to linear elasticity<br />

The pressure projection <strong>for</strong>mulation can be degenerated to linear elasticity by setting F ij ! d ij to result<br />

in:<br />

T ijkl ˆ ~T ijkl ˆ ~D 1 ijkl ˆ 0;<br />

…69†<br />

D ijkl ˆ C ijkl ˆ C 1 ijkl ‡ C 2 ijkl ˆ 2<br />

3 …A 10 ‡ A 01 †‰3…d ik d jl ‡ d il d jk † 2d ij d kl Š: …70†<br />

In linear elasticity, the shear modulus can be related to A 10 and A 01 by<br />

<br />

l ˆ 2<br />

oW<br />

oI 1<br />

‡ oW <br />

ˆ 2…A 10 ‡ A 01 †<br />

oI 2<br />

I1ˆI 2ˆ3<br />

and thus<br />

<br />

C ijkl ˆ l …d ik d jl ‡ d il d jk † 2 3 d ijd kl<br />

: …72†<br />

…71†<br />

The pressure projection sti€ness matrix degenerates to<br />

Z<br />

K s IJ ˆ B T CB dX ‡ kA s …B T I gT gB J † : …X s ;Y s †<br />

X s X<br />

…73†<br />

Further reduction of g to g ˆ ‰1; 1; 1; 0; 0; 0Š, and taking the bulk modulus k ˆ k ‡ 2=3l, Eq. (73) can be<br />

rearranged as<br />

Z<br />

<br />

K s IJ ˆ B T I<br />

CB J dX ‡ k ‡ 2 …A<br />

3 l s b T I b J† ; …74†<br />

…X s ;Y s †<br />

X s X<br />

where<br />

b I ˆ ‰W I;X ; W I;Y Š:<br />

…75†<br />

4. Stabilized reduced integration<br />

4.1. Finite elasticity<br />

The computational eciency of RKPM can be further <strong>improved</strong> by introducing a stabilized reduced<br />

integration <strong>method</strong> <strong>for</strong> the construction of the sti€ness matrix and <strong>for</strong>ce vector. It has been discussed by<br />

Simo and Hughes [44], that if a certain orthogonality condition exists between the stress ®eld, strain ®eld,<br />

and displacement gradient in the Hu-Washizu variational principle, a single-®eld variational principle with<br />

strain as the independent variable can be obtained. Liu et al. [34] then introduced an assumed strain in the<br />

following <strong>for</strong>m<br />

e ˆ Bd;<br />

where the assumed strain matrix B is the Taylor expansion of the displacement gradient matrix <strong>for</strong>mulated<br />

in the natural coordinate. In meshless <strong>for</strong>mulation, since the <strong>reproducing</strong> <strong>kernel</strong> shape functions are<br />

constructed at the global Cartesian coordinate, we introduce the following assumed de<strong>for</strong>mation gradient<br />

®eld in the integration zone X s X :<br />

…76†


128 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

dF ˆ Bdd; DF ˆ BDd; …77†<br />

B I …X ; Y † ˆ B I j …X s ;Y s † ‡ …X X s †B I;X j ‡ …Y Y s …X s ;Y s †<br />

†B I;X j …X s ;Y s †<br />

<strong>for</strong> …X ; Y † 2 X s X ; …78†<br />

where …X s ; Y s † is the centroid of X s X as shown in Fig. 1. By replacing B I by B I of Eq. (78) in Eqs. (57) and<br />

(58), and further evaluating D , T, D, ~ and T ~ at …X s ; Y s † <strong>for</strong> …X ; Y † 2 X s X<br />

, one can obtain the following<br />

sti€ness matrices associated with the integration zone X s X :<br />

… K† s IJ ˆAs ‰B T I … T ‡ D†B J Š ‡ m s<br />

…X s ;Y s XX ‰BT I;<br />

† X<br />

… T ‡ D†B J;X Š ‡ m s<br />

…X s ;Y s YY ‰BT I;<br />

† Y<br />

… T ‡ D†B J;Y Š …X s ;Y s †<br />

‡ m s XY ‰BT I; X<br />

… T ‡ D†B J;Y ‡ B T I; Y<br />

… T ‡ D†B J;X Š ‡ K s<br />

…X s ;Y s IJ ; …79†<br />

†<br />

… K† ~ s IJ ˆ … K ~ † s IJ ‡ … K ~ † s IJ<br />

ˆ A s ‰B T I …~ T ‡ D ~ 1 †B J Š ‡ m s<br />

…X s ;Y s XX ‰BT I;<br />

† X<br />

… T ~ ‡ D ~ 1 †B J;X Š ‡ m s<br />

…X s ;Y s YY ‰BT I;<br />

† Y<br />

… T ~ ‡ D ~ 1 †B J;Y Š …X s ;Y s †<br />

‡ m s XY ‰BT I; X<br />

… T ~ ‡ D ~ 1 †B J;Y ‡ B T I; Y<br />

… T ~ ‡ D ~ 1 †B J;X Š ‡ kA s …B T<br />

…X s ;Y s I gT gB J † ‡<br />

† …X<br />

K ~ s s ;Y s † IJ : …80a†<br />

The ®rst term and the second last term in Eq. (80a) can be combined to yield<br />

… K† ~ s IJ ˆ As ‰B T I …~ T ‡ D ~ 1 ‡ D ~ 2 †B J Š ‡ m s<br />

…X s ;Y s XX ‰BT I;<br />

† X<br />

… T ~ ‡ D ~ 1 †B J;X Š ‡ m s<br />

…X s ;Y s YY ‰BT I;<br />

† Y<br />

… T ~ ‡ D ~ 1 †B J;Y Š …X s ;Y s †<br />

‡ m s XY ‰BT I; X<br />

… T ~ ‡ D ~ 1 †B J;Y ‡ B T I; Y<br />

… T ~ ‡ D ~ 1 †B J;X Š ‡ K ~ s<br />

…X s ;Y s IJ ; …80b†<br />

†<br />

where A s is the area of X s X , and ms XX , ms YY and ms XY are the second area moments of Xs X<br />

that can be directly<br />

related to or approximated by the moment of inertia IXX s , I YY s , and I XY s of Xs X<br />

as shown in Table 1. Note<br />

that the moment of inertia can be obtained analytically without numerical integration. The matrices K s IJ<br />

and K ~ s IJ<br />

that contain the ®rst moments vanish in plane-strain case, and are nonzero <strong>for</strong> an axisymmetric<br />

problem:<br />

K s IJ ˆ ms X ‰BT I; X<br />

… T ‡ D†B J ‡ B T I … T ‡ D†B J;X Š ‡ m s<br />

…X s ;Y s Y ‰BT I;<br />

† Y<br />

… T ‡ D†B J ‡ B T I … T ‡ D†B J;Y Š …81†<br />

…X s ;Y †; s<br />

~K s IJ ˆ ms X ‰BT I; X<br />

… T ~ ‡ D ~ 1 †B J ‡ B T I …~ T ‡ D ~ 1 †B J;X Š ‡ m s<br />

…X s ;Y s Y ‰BT I;<br />

† Y<br />

… T ~ ‡ D ~ 1 †B J ‡ B T I …~ T ‡ D ~ 1 †B J;Y Š …X s ;Y †:<br />

s<br />

By combining Eqs. (79) and (80b), we obtain the following explicit tangential sti€ness matrix <strong>for</strong> ®nite<br />

elasticity<br />

…82†<br />

…K† s IJ ˆ … K† s IJ ‡ … ~K † s IJ ‡ … ~K † s IJ ˆ …K 0† s IJ ‡ …K stab† s IJ ;<br />

…83†<br />

where<br />

…K 0 † s IJ ˆ Ak ‰B T I …D ‡ T†B JŠ <br />

…X s ;Y s † ;<br />

…84†<br />

…K stab † s IJ ˆ ms XX ‰BT I; X<br />

…D ‡ T†B J;X Š ‡ m s<br />

…X s ;Y s YY ‰BT I;<br />

† Y<br />

…D ‡ T†B J;Y Š …X s ;Y s †<br />

‡ m s XY ‰BT I; X<br />

…D ‡ T†B J;Y ‡ B T I; Y<br />

…D ‡ T†B J;X Š ‡ K s<br />

…X s ;Y s IJ ; …85†<br />


J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 129<br />

Table 1<br />

Area moments <strong>for</strong> plane-strain and axisymmetric cases<br />

Plane-strain<br />

Axisymmetric<br />

m s X<br />

0 Z<br />

m s Y<br />

0 Z<br />

m s XX<br />

m s YY<br />

m s XY<br />

Z<br />

Z<br />

Z<br />

X s X<br />

X s X<br />

X s X<br />

…X X s † 2 dX ˆ I s XX<br />

…Y Y s † 2 dX ˆ I s YY<br />

…X X s †…Y Y s † dX ˆ I s XY<br />

Z<br />

Z<br />

Z<br />

Z<br />

X s X<br />

X K X<br />

X S X<br />

X s X<br />

X s X<br />

X s X<br />

Z<br />

…X X s †X dX ˆ …X X s † 2 dX ˆ I s XX<br />

X s X<br />

Z<br />

…Y Y S †X dX ˆ …X X S †…Y Y S †dX ˆ I S XY<br />

X K X<br />

Z<br />

…Y Y S †X dX ˆ …X X S †…Y Y S † dX ˆ I S XY<br />

X S X<br />

…X X s † 2 X dX X s Z<br />

…Y Y s † 2 X dX X s Z<br />

X s X<br />

X s X<br />

…X X s †…Y Y s †X dX X s I s XY<br />

…X X s † 2 dX ˆ X s I s XX<br />

…Y Y s † 2 dX ˆ X s I s YY<br />

K s IJ ˆ ms X ‰BT I; X<br />

…D ‡ T†B J ‡ B T I …D ‡ T†B J;X Š ‡ m s<br />

…X s ;Y s Y ‰BT I;<br />

† Y<br />

…D ‡ T†B J ‡ B T I …D ‡ T†B J;Y Š …X s ;Y †;<br />

s …86†<br />

D ˆ D ‡ ~ D 1 ‡ ~ D 2 ;<br />

D ˆ D ‡ ~ D 1 ˆ D ~ D 2 ;<br />

T ˆ T ‡ ~ T<br />

…87†<br />

…88†<br />

…89†<br />

and …K 0 † s IJ and …K stab† s IJ<br />

are the one-point quadrature and the stabilization sti€ness matrices, respectively.<br />

To <strong>for</strong>mulate the internal <strong>for</strong>ce vector, the ®rst Piola±Kirchho€ stress is calculated by<br />

r ij ˆ F im S mj ‰F im j …X s ;Y s † ‡ F im; X<br />

j …X X s …X s ;Y s †<br />

† ‡ F im;Y j …Y Y s<br />

…X s ;Y s †<br />

†Š…S …X †<br />

mj s ;Y s †<br />

ˆ …F im S mj † <br />

…X s ;Y s † ‡ …F im; X<br />

S mj † <br />

…X s ;Y s † …X X s † ‡ …F im;Y S mj † <br />

…X s ;Y s † …Y Y s †;<br />

…90†<br />

where r ij ˆ r ij ‡ ~r ij , S ij ˆ S ij ‡ ~S ij , and the resulting internal <strong>for</strong>ce vector is<br />

…f int † s I ˆ … f int † s I ‡ …~ f int † s I<br />

int ˆ …f<br />

0 †s int<br />

I<br />

‡ …f<br />

stab †s I ;<br />

…91†<br />

where<br />

…f int<br />

0 †s I ˆ As …B T I r† <br />

…X s ;Y s † ; …92†<br />

stab †s I ˆ ms XX …BT I; X<br />

r X † ‡ m s …X s ;Y s YY …BT I;<br />

† Y<br />

r Y † ‡ m s<br />

…X s ;Y s XY …BT I;<br />

† X<br />

r Y ‡ B T I; Y<br />

r X † ‡ a s<br />

…X s ;Y s I ; …93†<br />

†<br />

…f int<br />

r ˆ ^FS; r X ˆ ^F ;X S; r Y ˆ ^F ;Y S; …94†


130 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

2<br />

S ˆ<br />

6<br />

4<br />

S 11<br />

S 12<br />

S 21<br />

S 22<br />

3<br />

7<br />

; a ˆ 5<br />

1 <strong>for</strong> axisymmetric<br />

0 <strong>for</strong> plane-strain;<br />

…95†<br />

2<br />

^F ˆ<br />

6<br />

4<br />

aS 33<br />

F 11 0 F 12 0 0<br />

0 F 11 0 F 12 0<br />

F 21 0 F 22 0 0<br />

0 F 21 0 F 22 0<br />

0 0 0 0 aF 33<br />

3<br />

7<br />

: …96†<br />

5<br />

The vector a s I ˆ 0 <strong>for</strong> plane-strain case, and in axisymmetric problems<br />

a s I ˆ ms X ‰BT I; X<br />

r ‡ B T I r X Š ‡ m s<br />

…X s ;Y s Y ‰BT I;<br />

† Y<br />

r ‡ B T I r Y Š …97†<br />

…X s ;Y †: s<br />

4.2. Degeneration to linear elasticity<br />

The linear dilatational sti€ness matrix given in Eq. (74) is ®rst trans<strong>for</strong>med into the following <strong>for</strong>m<br />

<br />

k ‡ 2 A<br />

3 l s b T I b J ˆ A s …B T ~ I<br />

CB J †; …98†<br />

where<br />

<br />

~C ijkl ˆ k ‡ 2 d<br />

3 l ij d kl; : …99†<br />

Taking the similar degeneration procedures as discussed in Section 3.3, and employing Eq. (99), the sti€ness<br />

matrix <strong>for</strong> linear elasticity is obtained,<br />

K s IJ ˆ …K 0† s IJ ‡ …K stab† s IJ ;<br />

…100†<br />

…K 0 † s IJ ˆ As B T I CB J;<br />

…101†<br />

…K stab † s IJ ˆ ms XX ‰BT I; CBJ;X X<br />

Š ‡ m s<br />

…X s ;Y s YY ‰BT I; CBJ;Y<br />

† Y<br />

Š ‡ m s<br />

…X s ;Y s XY ‰BT I; CBJ;Y<br />

<br />

† X<br />

‡ B T I; CBJ;X Y<br />

Š ‡ K s<br />

…X s ;Y s IJ ;<br />

†<br />

K s IJ ˆ ms X ‰BT I; CBJ <br />

X<br />

‡ B T I<br />

CB J;X Š ‡ m s<br />

…X s ;Y s Y ‰BT I; CBJ <br />

† Y<br />

‡ B T <br />

I<br />

CB J;Y Š<br />

<br />

…X s ;Y s †<br />

…axisymmetric†;<br />

…102†<br />

…103†<br />

where K s IJ ˆ 0 <strong>for</strong> plane-strain case, C, ~ C, and C are the matrix <strong>for</strong>ms of C ijkl (Eq. (72), ~C ijkl (Eq. (99)), and<br />

C ijkl , where<br />

C ijkl ˆ ~C ijkl ‡ C ijkl ˆ kd ij d kl ‡ l…d ik d jl ‡ d il d jk †:<br />

Note that the stabilization matrix is only associated with distortional energy.<br />

…104†<br />

5. Numerical examples<br />

In this section, cubic B-spline <strong>kernel</strong> function with linear monomial basis functions is employed. The<br />

abbreviations de®ned in Table 2 denote di€erent numerical <strong>method</strong>s used in the numerical examples. For


J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 131<br />

Table 2<br />

Abbreviations used in the numerical examples<br />

Abbreviation<br />

[RKPM:GI]<br />

[RKPM:PP-GI]<br />

[RKPM:PP-SRIM]<br />

[T]<br />

[Q]<br />

[R]<br />

Employed numerical <strong>method</strong>s<br />

The original RKPM <strong>for</strong>mulation using Gauss integration <strong>method</strong>.<br />

RKPM with pressure projection <strong>method</strong>. Gauss integration <strong>method</strong> is used to per<strong>for</strong>m<br />

domain integration.<br />

RKPM with pressure projection <strong>method</strong>. Stabilized reduced integration <strong>method</strong> is used to<br />

per<strong>for</strong>m domain integration.<br />

Triangular integration zone de®ned by 3 points.<br />

Quadrilateral integration zone de®ned by 4 points.<br />

Normalized dilation parameter, R ˆ a , where a is the dilation parameter of the <strong>kernel</strong><br />

h<br />

function, and h is the averaged nodal distance.<br />

p <br />

Gauss integration (GI), the integration order of …m ‡ 2† … m ‡ 2† is used, where m is the number of<br />

nodes inside the integration zone. For simplicity, if neither [T] nor [Q] is speci®ed in the numerical examples,<br />

the quadrilateral integration zone is employed.<br />

5.1. Linear patch test<br />

The patch test is often used to verify the consistency of a numerical <strong>method</strong>. A patch test model used in<br />

this example is shown in Fig. 3. Linear boundary displacements that satisfy the incompressible condition<br />

are prescribed on the boundary nodes,<br />

u x ˆ ax; u y ˆ ay …a 6ˆ 0; a 2 R†: …105†<br />

Several skewness of the center node as shown in Table 3 are investigated, and the triangular integration<br />

zones are used to handle the severe skewness. The center node displacements are solved using RKPM:PP-<br />

SRIM-T <strong>method</strong>. This proposed <strong>method</strong> passes all the patch tests regardless of the center node location and<br />

the choice of <strong>kernel</strong> support size as long as it satis®es the minimum support size requirement.<br />

5.2. Linear elastic beam subjected to transverse shear load and tip bending moment<br />

A plane-strain beam with E ˆ 3:0 10 7 , m ˆ 0:4999, L ˆ 4 and D ˆ 1 is subjected to a shear <strong>for</strong>ce P ˆ 1<br />

as shown in Fig. 4. The problem is modeled with regularly spaced <strong>particle</strong>s as shown in Fig. 4. Both<br />

Fig. 3. Patch test Model.


132 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

Table 3<br />

Center node locations in the patch test<br />

x-coordinate<br />

y-coordinate<br />

0.5 0.5<br />

0.55 0.55<br />

0.6 0.55<br />

0.05 0.9<br />

Fig. 4. Incompressible beam subjected to a transverse shear load: problem description.<br />

Table 4<br />

Solution accuracy and eciency comparison of an incompressible beam subjected to transverse load<br />

Support size (R) Normalized time Normalized displacement<br />

RKPM:GI RKPM:PP-SRIM RKPM:GI RKPM:PP-SRIM<br />

1.001 1.0 0.702 0.066 0.865<br />

2.0 1.932 0.765 0.315 0.999<br />

3.0 3.951 1.052 0.922 0.999<br />

triangular and quadrilateral integration zones are considered. The analysis is per<strong>for</strong>med using RKPM:GI<br />

(the original RKPM) and the proposed RKPM:PP-SRIM, and the results are compared with linear elasticity<br />

solution [45].<br />

The predicted normalized vertical displacements at the middle point of the loading surface and the<br />

normalized CPU time are summarized in Table 4. In this table, the CPU time is normalized by that in<br />

the case of RKPM:GI (the original RKPM) with R ˆ 1:001, and the displacements are normalized by the<br />

analytical solution. The results indicate that RKPM:PP-SRIM provides a better solution accuracy than<br />

RKPM:GI regardless of the support size. RKPM:PP-SRIM is particularly advantageous over RKPM:GI<br />

when the support size is small, in which case RKPM:GI is severely locked. In this problem, the proposed<br />

RKPM:PP-SRIM using triangular and quadrilateral integration zones have similar per<strong>for</strong>mance. The shear<br />

stress distributions <strong>for</strong> normalized support sizes of R ˆ 1:001; 2:0; 3:0 are compared in Fig. 5a±c. Note that<br />

R > 1:0 is required to meet <strong>kernel</strong> stability. For the RKPM:GI <strong>method</strong>, the shear stress distributions are


J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 133<br />

Fig. 5. (a) Incompressible beam subjected to a transverse shear load: comparison of shear stress distribution <strong>for</strong> R ˆ 1.001. (b) Incompressible<br />

beam subjected to a transverse shear load comparison of shear stress distribution <strong>for</strong> R ˆ 2.0. (c) Incompressible beam<br />

subjected to a transverse shear load comparison of shear stress distribution <strong>for</strong> R ˆ 3.0.<br />

completely o€, especially when R ˆ 1:001. The proposed RKPM:PP-SRIM <strong>method</strong>, on the other hand,<br />

predicts a much accurate shear stress distribution.<br />

The second analysis deals with the same plane-strain beam subjected to a tip bending moment as shown<br />

in Fig. 6. The same analysis procedures as those used in the shear problem were per<strong>for</strong>med using<br />

RKPM:PP-SRIM and RKPM:GI. The predicted normal stress distribution plotted in Fig. 7a±c also favors<br />

RKPM:PP-SRIM.<br />

5.3. Plane-strain tube subjected to internal pressure<br />

<strong>An</strong> in®nitely long tube, with an inner radius of 6 cm and an outer radius of 8 cm, is subjected to<br />

an internal pressure as shown in Fig. 8. This problem is analyzed by an axisymmetric RKPM <strong>for</strong>mulation<br />

with constraints introduced in the axial direction to impose plane-strain condition as described<br />

in Fig. 8.<br />

A Rivlin type hyperelastic material model is used in this example in which the distortional strain energy<br />

density is expressed by<br />

W …I 1 ; I 2 † ˆ X1<br />

A mn …I 1 3† m …I 2 3† n ; I 1 ˆ I 1 I 1=3<br />

3<br />

; I 2 ˆ I 2 I 2=3<br />

3<br />

: …106†<br />

m‡nˆ1


134 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

Fig. 6. Incompressible beam subjected to a tip bending moment: problem description.<br />

The material properties are A 10 ˆ 0:373 MPa, A 20 ˆ 0:031 MPa, A 30 ˆ 0:005 MPa , and the penalty<br />

associated with the dilatational strain energy density of Eq. (29) is k ˆ 10 5 MPa. A total of 18 <strong>particle</strong>s,<br />

arranged regularly and irregularly, are used to discretize the thickness of the tube. Both triangular and<br />

quadrilateral integration zones are used <strong>for</strong> RKPM:PP-SRIM analysis, whereas <strong>for</strong> the RKPM:GI (the<br />

original RKPM) <strong>method</strong>, only the quadrilateral integration zone is considered. The analytical solution can<br />

be found in Refs. [46,39].<br />

The pressure-displacement curves predicted by the RKPM:GI are shown in Fig. 9a and b <strong>for</strong> regular<br />

and irregular models, respectively. The results show that large support size must be used in the RKPM:GI<br />

(the original RKPM) computations to avoid incompressible locking. With a pressure projection <strong>method</strong>,<br />

on the other hand, very accurate nonlinear load-displacement response is captured even with the minimum<br />

support size. The solution of RKPM:PP-SRIM, using either regular or irregular models with<br />

rectangular or triangular integration zones, follows the analytical solution almost perfectly, as shown in<br />

Fig. 9c.<br />

The predicted pressure and circumferential stress are also compared. Knowing the large support size is<br />

required <strong>for</strong> RKPM:GI (the original RKPM), the pressure distributions computed by RKPM:GI with large<br />

support size R ˆ 3:0 using regularly spaced and irregularly spaced models are plotted in Figs. 10 and 11,<br />

respectively. Signi®cant pressure oscillation is observed <strong>for</strong> both cases. With the proposed RKPM:PP-<br />

SRIM, accurate pressure solutions <strong>for</strong> both regular and irregular models are obtained. A similar trend is<br />

also found in the circumferential stress distribution as shown in Figs. 12 and 13. Moreover, RKPM:PP-<br />

SRIM requires relatively less computation time compared to RKPM:GI (the original RKPM) as listed in<br />

Table 5.<br />

5.4. Radial compression of a rubber bushing<br />

<strong>An</strong> annular bushing, composed of an inner metal sleeve, an outer metal sleeve, and a rubber insert, is<br />

subjected to a radial prescribed displacement as shown in Fig. 14. Due to the relatively higher sti€ness of<br />

the metal sleeves, only the rubber insert is modeled with the outer surface completely ®xed and the inner<br />

surface moved as a rigid surface in the vertical direction. The rubber properties of Eq. (106) are<br />

A 10 ˆ 0:373 MPa, A 20 ˆ 0:031 MPa, A 30 ˆ 0:005 MPa, and k ˆ 1000 MPa. Two analysis models as<br />

shown in Fig. 15 are considered. The coarse model consists of 153 nodes and the ®ne model consists of 297<br />

nodes. A total of 8 16 and 8 32 quadrilateral integration zones are used in the coarse and ®ne models,<br />

respectively.


J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 135<br />

Fig. 7. (a) Incompressible beam subjected to a tip bending moment: comparison of axial stress distribution <strong>for</strong> R ˆ 1.001. (b) Incompressible<br />

beam subjected to a tip bending moment: comparison of axial stress distribution <strong>for</strong> R ˆ 2.0. (c) Incompressible beam<br />

subjected to a tip bending moment: comparison of axial stress distribution <strong>for</strong> R ˆ 3.0.<br />

Several RKPM:GI (the original RKPM) plane-strain analyses were ®rst per<strong>for</strong>med using coarse and ®ne<br />

models with various dilation parameters to study the solution accuracy. Since it is easier to identify locking<br />

at a linear range, a linear solution by Stevenson [47] is also compared <strong>for</strong> the veri®cation of locking<br />

phenomenon in Fig. 16. When the coarse model is used with a small <strong>kernel</strong> support, locking behavior is<br />

observed in the RKPM:GI (the original RKPM) solution. This locking can be alleviated either by using a<br />

larger <strong>kernel</strong> support or by re®ning the model as shown in Fig. 16. However, these two alternatives<br />

consume considerably more computation time. Employing an RKPM:PP-GI <strong>method</strong>, on the other hand,


136 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

Fig. 8. Plain-strain tube subjected to an internal pressure: problem description.<br />

does not encounter locking even when a 153-node coarse model with a small <strong>kernel</strong> support is used <strong>for</strong><br />

analysis.<br />

Fig. 17 demonstrates the bushing full cross-sectional de<strong>for</strong>mations predicted by the RKPM:PP-GI<br />

<strong>method</strong>. The pressure solutions obtained by RKPM:GI (the original RKPM) and RKPM:PP-GI <strong>method</strong>s<br />

are compared in Fig. 18. The RKPM:GI analysis encounters a severe pressure oscillation in the 153-node<br />

model with small <strong>kernel</strong> support size. As the <strong>kernel</strong> support size or the number of nodes is increased, the<br />

pressure oscillation decreases only marginally. Conversely, the RKPM:PP-GI <strong>method</strong> predicts a very<br />

smooth pressure distribution as shown in Fig. 18d. The comparison of the linear solution accuracy (normalized<br />

by the Stevenson's linear solution) and the CPU time (normalized by RKPM:GI <strong>method</strong> using the<br />

153-node model with dilation parameter R ˆ 1.001) listed in Table 6 demonstrates the advantages of using<br />

the pressure projection <strong>method</strong> in RKPM computation. Note that in this problem the dilation parameters<br />

are normalized with respect to the maximum nodal distance.<br />

In this problem, however, the RKPM:PP-SRIM <strong>method</strong> requires many more iterations to converge<br />

during the nonlinear iteration compared to RKPM:PP-GI; there<strong>for</strong>e, no additional eciency is gained<br />

using RKPM:PP-SRIM.<br />

6. Conclusion<br />

Numerical <strong>method</strong>s to improve the computational eciency and accuracy of the RKPM <strong>for</strong> <strong>nearly</strong><br />

incompressible ®nite elasticity are introduced. Since the <strong>method</strong> is a general approach <strong>for</strong> ®nite elasticity,


J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 137<br />

Fig. 9. (a) Plain-strain tube subjected to an internal pressure: pressure-displacement curves from RKPM:GI using a regular model. (b)<br />

Plain-strain tube subjected to an internal pressure: pressure-displacement curves from RKPM:GI using an irregular model. (c) Plainstrain<br />

tube subjected to an internal pressure: pressure-displacement curves from RKPM:PP-SRIM and RKPM:PP-GI using regular<br />

and irregular models.<br />

it can also be applied to <strong>nearly</strong> incompressible linear elasticity. The study of constraint count suggests<br />

that the original RKPM using Gauss integration (RKPM:GI) is over-constrained and it causes locking<br />

and pressure oscillation. To reduce the independent discrete constraint equations imposed at the quadrature<br />

points, the pressure is projected onto a lower-order space by a least-squares projection<br />

(RKPM:PP-GI). In the most severe over-constrained situation where each integration zone contains only<br />

3 or 4 nodes, the pressure is projected onto a constant ®eld within the triangular or quadrilateral basic


138 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

Fig. 10. Plain-strain tube subjected to an internal pressure: comparison of pressure distribution from RKPM:GI, RKPM:PP-SRIM<br />

and RKPM:PP-GI using a regular model.<br />

Fig. 11. Plain-strain tube subjected to an internal pressure: comparison of pressure distribution from RKPM:GI, RKPM:PP-SRIM<br />

and RKPM:PP-GI using an irregular model.<br />

integration units. The resulting constraint count increases from 1/3 to 1 <strong>for</strong> a triangular integration zone,<br />

and from 2/3 to 2 <strong>for</strong> a quadrilateral integration zone. As a result, locking and pressure oscillation are<br />

eliminated, and it allows the use of <strong>kernel</strong> functions with small support size to signi®cantly reduce<br />

computation time.<br />

In addition to pressure projection, a stabilized reduced integration <strong>method</strong> is also introduced to further<br />

reduce the computational cost (RKPM: PP-SRIM). The stabilization terms are obtained by taking an<br />

assumed strain approach, in which the assumed strain ®eld associated with the de<strong>for</strong>mation gradient is<br />

approximated by a Taylor series expansion of the gradient matrix along the centroid of the integration<br />

zone. The resulting sti€ness matrix and <strong>for</strong>ce vector are expressed in the <strong>for</strong>m of one-point integration and


J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 139<br />

Fig. 12. Plain-strain tube subjected to an internal pressure: comparison of circumferential stress distribution from RKPM:GI,<br />

RKPM:PP-SRIM and RKPM:PP-GI using a regular model.<br />

Fig. 13. Plain strain tube subjected to an internal pressure: comparison of circumferential stress distribution from RKPM:GI,<br />

RKPM:PP-SRIM and RKPM:PP-GI using an irregular model.<br />

Table 5<br />

Normalized CPU time comparison in tube in¯ation problem<br />

R RKPM:GI RKPM:PP-SRIM-Q RKPM:PP-SRIM-T<br />

1.0 1.00 0.32 0.56<br />

2.0 2.39 0.49 1.12<br />

3.0 3.66 0.67 1.58


140 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

Fig. 14. Rubber bushing under radial compression : problem description.<br />

Fig. 15. Bushing analysis models.<br />

stabilization terms. The stabilization sti€ness matrix and <strong>for</strong>ce vector are derived explicitly without the need<br />

of numerical integration.<br />

The accuracy and eciency of RKPM: PP-GI is demonstrated in the numerical examples. The numerical<br />

results show that the <strong>method</strong> e€ectively resolves the volumetric locking and stress oscillations<br />

and the CPU time is signi®cantly reduced. It is also observed that RKPM: PP-SRIM is more e€ective<br />

<strong>for</strong> linear problems. For ®nite elasticity, RKPM: PP-GI leads to a faster convergence during Newton<br />

iteration.


J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 141<br />

Fig. 16. Predicted radial load-displacement response of rubber bushing.<br />

Table 6<br />

Solution accuracy and eciency comparison of a bushing compression problem<br />

<strong>An</strong>alysis model Normalized time Normalized linear solution accuracy<br />

RKPM:GI (153N, R ˆ 1.001) 1.00 1.64<br />

RKPM:GI (153N, R ˆ 1.8) 3.66 0.99<br />

RKPM:GI (297N, R ˆ 2.0) 6.79 0.99<br />

RKPM:PP-GI (153N, R ˆ 1.001) 1.07 0.99<br />

Acknowledgements<br />

The support of this research by Army TARDEC to the University of Iowa, and National Science<br />

Foundation and Oce of Naval Research to Northwestern University is greatly acknowledged. This work<br />

is also sponsored in part by the Army High Per<strong>for</strong>mance Computing Research Center under the auspices of<br />

the Department of the Army, Army Research Laboratory cooperative agreement number DAAH04-95-2-<br />

003/contract number DAAH04-95-C-0008. The content of which does not necessarily re¯ect the position or<br />

the policy of the government, and no ocial endorsement should be inferred.<br />

Appendix A. Explicit expressions of S ij ; ~ S ij ; D ijkl ; ~ D 1 ijkl ; ~ D 2 ijkl ; T ijkl ; ~ T ijkl<br />

<br />

S ij ˆ 2 K 1 I 1=3<br />

3<br />

d ij 1 <br />

3 I 1G 1<br />

ij<br />

‡ K 2 I 2=3<br />

3<br />

I 1 d ij G ij 2 <br />

3 I 2G 1<br />

ij<br />

; …A:1†<br />

~S ij ˆ PJG 1<br />

ij ;<br />

…A:2†


142 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

Fig. 17. Progressive de<strong>for</strong>mations of rubber bushing.<br />

T ijkl ˆ d ik<br />

S jl ;<br />

…A:3†<br />

~T ijkl ˆ d ik<br />

~S jl ; …A:4†<br />

D ijkl ˆ F im F kn<br />

~C 1 mjnl ;<br />

~C 1 ijkl ˆ JP…G1 ij G1 kl<br />

G 1<br />

ik G1 jl<br />

G 1<br />

il G1 jk †;<br />

…A:5†<br />

~D ijkl ˆ F im F kn<br />

~C 2 mjnl ; ~C 2 ijkl ˆ JG1 ij<br />

G ij ˆ F ki F kj ;<br />

oP<br />

;<br />

oE kl<br />

…A:6†<br />

…A:7†<br />

H ij ˆ o2 W …I 1 ; I 2 †<br />

oI i oI j<br />

; I 1 ˆ I 1 I 1=3<br />

3<br />

; I 2 ˆ I 2 I 2=3<br />

3<br />

: …A:8†


J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 143<br />

Fig. 18. Comparison of pressure distribution using RKPM:GI and RKPM:PP-GI.<br />

References<br />

[1] L. Lucy, A numerical approach to testing the ®ssion hypothesis, Astron. J. 82 (1977) 1013±1024.<br />

[2] J.J. Monaghan, <strong>An</strong> introduction to SPH, Comput. Phys. Commun. 48 (1988) 89±96.<br />

[3] P.W. Randles, L.D. Libersky, Smoothed <strong>particle</strong> hydrodynamics: some recent improvements and applications, Comput. Meth.<br />

Appl. Mech. Engng. 139 (1996) 375±408.<br />

[4] B. Nayroles, G. Touzot, P. Villon, Generalizing the ®nite element <strong>method</strong>: di€use approximation and di€use elements, Comput.<br />

Mech. 10 (1992) 307±318.<br />

[5] T. Belytschko, Y.Y. Lu, L. Gu, Element free Galerkin <strong>method</strong>s, Int. J. Numer. Meth. Engng. 37 (1994) 229±256.<br />

[6] Y.Y. Lu, T. Belytschko, L. Gu, A new implementation of the element free Galerkin <strong>method</strong>, Comput. Meth. Appl. Mech. Engng.<br />

113 (1994) 397±414.<br />

[7] T. Belytschko, M. Tabarra, Dynamic fracture using Element-Free Galerkin <strong>method</strong>s, Int. J. Numer. Meth. Engng. 39 (1996) 923±<br />

938.


144 J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145<br />

[8] T. Belytschko, Y. Krongauz, D. Organ, M. Fleming, P. Krysl, Meshless <strong>method</strong>s: an overview and recent development, Comput.<br />

Meth. Appl. Mech. Engng. 139 (1996) 3±49.<br />

[9] I. Kaljevic, S. Saigal, <strong>An</strong> <strong>improved</strong> element free Galerkin <strong>for</strong>mulation, Int. J. Numer. Meth. Engng. Accepted.<br />

[10] W.K. Liu, S. Jun, Y.F. Zhang, Reproducing Kernel <strong>particle</strong> <strong>method</strong>s, Int. J. Numer. Meth. Fluids 20 (1995) 1081±1106.<br />

[11] W.K. Liu, S. Jun, S. Li, J. Adee, T. Belytschko, Reproducing Kernel <strong>particle</strong> <strong>method</strong>s <strong>for</strong> structural dynamics, Int. J. Numer.<br />

Meth. Engng. 38 (1995) 1655±1679.<br />

[12] W.K. Liu, Y. Chen, C.T. Chang, T. Belytschko, Advances in multiple scale <strong>kernel</strong> <strong>particle</strong> <strong>method</strong>s, Comput. Mech. 18 (2) (1996)<br />

31±111.<br />

[13] W.K. Liu, Y. Chen, R.A. Uras, C.T. Chang, Generalized multiple scale <strong>reproducing</strong> Kernel <strong>particle</strong> <strong>method</strong>s, Comput. Meth.<br />

Appl. Mech. Engng. 139 (1996) 91±158.<br />

[14] W.K. Liu, S. Li, T. Belytschko, Moving least square <strong>kernel</strong> Galerkin <strong>method</strong> <strong>method</strong>ology and convergenceTrans I, Comput.<br />

Meth. Appl. Mech. Engng. 143 (1997) 113±154.<br />

[15] J.S. Chen, C. Pan, C.T. Wu, W.K. Liu, Reproducing <strong>kernel</strong> <strong>particle</strong> <strong>method</strong>s <strong>for</strong> large de<strong>for</strong>mation analysis of nonlinear<br />

structures, Comput. Meth. Appl. Mech. Engng. 139 (1996) 195±227.<br />

[16] J.S. Chen, C. Pan, C.T. Wu, Large de<strong>for</strong>mation analysis of rubber based on a Reproducing Kernel Particle Method, Comput.<br />

Mech. 19 (1997) 211±227.<br />

[17] J.S. Chen, C. Pan, C.T. Wu, Applications of reproduction Kernel <strong>particle</strong> <strong>method</strong> to large de<strong>for</strong>mation and contact analysis of<br />

elastomers, Rubber Chem. Technol. (1997) 191±213.<br />

[18] C.A.M. Duarte, J.T. Oden, HP Clouds ± a Meshless <strong>method</strong> to solve boundary value problems, Technical Report 95±05, TICAM,<br />

The University of Texas, Austin.<br />

[19] C.A.M. Duarte, J.T. Oden, A h-p adaptive <strong>method</strong> using clouds, Comput. Meth. Appl. Mech. Engng. 139 (1996) 237±262.<br />

[20] J.M. Melenk, I. Babuska, The partition of unity ®nite element <strong>method</strong>: basic theory and applications, Comput. Meth. Appl. Mech.<br />

Engng. 139 (1996) 289±314.<br />

[21] I. Babuska, The ®nite element <strong>method</strong> with Lagrangian multipliers, Numer. Math. 20 (1973) 179±192.<br />

[22] F. Brezzi, On the existence uniqueness and approximation of Saddle-point problems arising from Lagrangian multipliers,<br />

R.A.I.R.O., Numer. <strong>An</strong>al. 8 (1974) 129±151.<br />

[23] J.C. Simo, R.L. Taylor, K.S. Pister, Variational and projection <strong>method</strong>s <strong>for</strong> volume constraint in ®nite de<strong>for</strong>mation Elastoplasticity,<br />

Comput. Meth. Appl. Mech. Engng. 51 (1985) 177±208.<br />

[24] T. Belytschko, W.E. Bachrach, Ecient implementation of quadrilaterals with high coarse-mesh accuracy, Comput. Meth. Appl.<br />

Mech. Engng. 54 (1986) 279±301.<br />

[25] T.S. Sussman, K.J. Bathe, A ®nite element <strong>for</strong>mulation <strong>for</strong> incompressible elastic and inelastic analysis, Computers & Structures<br />

26 (1987) 357±409.<br />

[26] W.K. Liu, T. Belytschko, J.S. Chen, Nonlinear versions of ¯exurally superconvergent elements, Comput. Meth. Appl. Mech.<br />

Engng. 71 (1988) 241±256.<br />

[27] T.Y.P. Chang, A.F. Saleeb, G. Li, Large strain analysis of rubber-like materials based on a perturbed Lagrangian variational<br />

principle, Comput. Mech. 8 (1991) 221±233.<br />

[28] J.S. Chen, W. Han, C.T. Wu, W. Duan, On the perturbed Lagrange <strong>for</strong>mulation <strong>for</strong> <strong>nearly</strong> incompressible and incompressible<br />

hyperelasticity, Comput. Meth. Appl. Mech. Engng. 142 (1997) 335±351.<br />

[29] D.S. Malkus, T.J.R. Hughes, Mixed ®nite element <strong>method</strong>s_reduced and selective integration techniques: a uni®cation of<br />

concepts, Comput. Meth. Appl. Mech. Engng. 15 (1978) 63±81.<br />

[30] T.J.R. Hughes, Generalization of selective integration procedures to anisotropic and nonlinear media, Int. J. Numer. Meth.<br />

Engng. 15 (1980) 1413±1418.<br />

[31] D.P. Flanagan, T. Belytschko, A uni<strong>for</strong>m strain hexahedron and quadrilateral with orthogonal hourglass control, Int. J. Numer.<br />

Meth. Engng. 17 (1981) 679±706.<br />

[32] T. Belytschko, C.S. Tsay, W.K. Liu, A stabilization matrix <strong>for</strong> the Bilinear Mindlin plate elements, Comput. Meth. Appl. Mech.<br />

Engng. 29 (1981) 313±327.<br />

[33] T. Belytschko, J.S.-J. Ong, W.K. Liu, J.M. Kennedy, Hourglass control in linear and nonlinear problems, Comput. Meth. Appl.<br />

Mech. Engng. 43 (1984) 251±276.<br />

[34] W.K. Liu, J.S.-J. Ong, R.A. Uras, Finite element stabilization matrices-a uni®cation approach, Comput. Meth. Appl. Mech.<br />

Engng. 53 (1986) 13±46.<br />

[35] W.K. Liu, Y.-K. Hu, T. Belytschko, Multiple quadrature underintegrated ®nite elements, Int. J. Numer. Meth. Engng. 37 (1994)<br />

3263±3289.<br />

[36] W.K. Liu, Y. Guo, S. Tang, T. Belytschko, A multiple-quadrature eight-node hexahedral ®nite element <strong>for</strong> large de<strong>for</strong>mation<br />

elastoplastic analysis, Comput. Meth. Appl. Mech. Engng. 154 (1998) 69±132.<br />

[37] J.S. Chen, K. Satyamurthy, L.R. Hirschfelt, Consistent ®nite element procedures <strong>for</strong> nonlinear rubber elasticity with a higher<br />

order strain energy function, Computers & Structures 50 (1994) 715±727.<br />

[38] J.S. Chen, C. Pan, A pressure projection <strong>method</strong> <strong>for</strong> <strong>nearly</strong> incompressible rubber hyperelasticity Part I: Theory, J. Appl. Mech. 63<br />

(1996) 862±868.<br />

[39] J.S. Chen, C.T. Wu, C. Pan, A pressure projection <strong>method</strong> <strong>for</strong> <strong>nearly</strong> incompressible rubber hyperelasticity Part II: Applications,<br />

J. Appl. Mech. 63 (1996) 869±876.<br />

[40] J.S. Chen, C.T. Wu, On computational issues in large de<strong>for</strong>mation analysis of rubber bushings, Mech. Struct. & Mach. 25 (3)<br />

(1997) 287±309.


J.-S. Chen et al. / Comput. Methods Appl. Mech. Engrg. 181 (2000) 117±145 145<br />

[41] J.S. Chen, C. Pan, T.Y.P. Chang, On the control of pressure oscillation in bilinear-displacement constant-pressure element,<br />

Comput. Meth. Appl. Mech. Engng. 128 (1995) 137±152.<br />

[42] J.C. Nagtegaal, D.M. Parks, J.R. Rice, On numerical accurate ®nite element solution in the fully plastic range, Comput. Meth.<br />

Appl. Mech. Engng. 4 (1974) 153±178.<br />

[43] T.J.R. Hughes, The Finite Element Method, Prentice-Hall, New Jersey, 1987.<br />

[44] J.C. Simo, T.J.R. Hughes, On the variational foundation of assumed strain <strong>method</strong>s, J. Appl. Mech. 53 (1986) 51±54.<br />

[45] S.P. Timoshenko, J.N. Goodier, Theory of Elasticity, 3rd ed., McGraw-Hill, New York, 1970.<br />

[46] R.S. Rivlin, Large elastic de<strong>for</strong>mation of isotropic materials Part VI: further results in the theory of torsion shear and ¯exure,<br />

Philos. Trans. R. Soc. London 242 (1949) 173±195.<br />

[47] A.C. Stevenson, Some boundary problems of two-dimensional elasticity, Philos. Mag. 34 (1943) 766±793.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!