Você está na página 1de 87

INVESTIGATION OF KERNELS FOR THE REPRODUCING KERNEL PARTICLE

METHOD

by
BALA PRIYADARSHINI SHANMUGAM

DAVID L. LITTLEFIELD, COMMITTEE CHAIR


ROY P. KOOMULLIL
BHARAT K. SONI

A THESIS
Submitted to the graduate faculty of The University of Alabama at Birmingham,
in partial fulfillment of the requirements for the degree of
Master of Science Mechanical Engineering
BIRMINGHAM, ALABAMA
2009

INVESTIGATION OF KERNELS FOR THE REPRODUCING KERNEL PARTICLE


METHOD
BALA PRIYADARSHINI SHANMUGAM
MECHANICAL ENGINEERING
ABSTRACT

Research on meshfree methods has shown promise and its applicability to


practical applications has grown significantly. Yet, there are certain aspects of meshfree
methods that limit their potential, which could be better understood by revisiting their
mathematical foundation.
This research was aimed to probe into the mathematical framework of meshfree
methods. The primary focus is on the Reproducing Kernel Particle Method (RKPM)
and its applications to structural analysis. One problem with some meshfree particle
methods, especially RKPM, is the failure of the window function to satisfy the partition
of unity property. In order to come satisfy this property, the effect of different kernels/
window functions on the solution accuracy is studied to help develop a more suitable
kernel. In the recent years, a number of new methods have been proposed to address the
problem, but in this thesis attention is drawn to deriving polynomial kernels that enable
easier implementation of RKPM and reduction of computational expense, without
having to switch to a new method. The enforcement of essential boundary conditions,
which is another difficulty of RKPM, is also addressed. Simulation of numerical
ii

problems for linear elastostatic and elastodynamic structures using the reproducing
kernel particle method is illustrated. Example calculations suggest that this class of
polynomial kernels can overcome some of the current limitations of RKPM.
Keywords: Meshfree methods, Reproducing Kernel Particle Method, Kernels, Essential
Boundary conditions

iii

This thesis is dedicated to my Mom, Dad, Sathish and my dear Dilip.

iv

ACKNOWLEDGEMENTS

I would like to express my heartfelt thanks to my advisor, Dr. David Littlefield, who has
supported and guided me through the course of this research. His ideas and input were
crucial to this work.

My sincere thanks to the committee members, Dr. Bharat Soni and Dr. Roy Koomullil for
the motivation and support they offered.

I would also like to thank my colleagues at the Computational Simulations Lab for all the
insightful discussions we have had, especially Sandeep.

TABLE OF CONTENTS

ABSTRACT ii
ACKNOWLEDGEMENTS v
TABLE OF CONTENTS ...vi
LIST OF ABBREVIATIONS ix
LIST OF FIGURES x
LIST OF TABLES xi
CHAPTER 1 . 1
INTRODUCTION . 1
1.1

Background . 1

1.2

Limitations of Finite Element Method 2

1.3

Meshfree Methods 3

1.4

Advantages and Limitations of Meshfree Method .. 4

CHAPTER 2 ... 8
MATHEMATICAL BACKGROUND .. 8
2.1

Method of Weighted Residuals ... 8

2.2

Basis Functions ... 9

2.3

Boundary Conditions . 10

CHAPTER 3 11
CLASSIFICATION OF MESHFREE METHODS . 11
3.1

Smoothed Particle Hydrodynamics (SPH) 13

3.2

Corrected Smooth Particle Hydrodynamics .. 15

3.3

Moving Least Square Particle Hydrodynamics (MLSPH) 16

3.4

Moving Least Squares Method .. 16

3.5

Element Free Galerkin Method . 17


vi

3.6

Meshless Local Petrov-Galerkin Method .. 18

3.7

Finite Cloud Method . 19

3.8

Partition Of Unity Methods ... 20

3.9

h-p Cloud Method . 21

CHAPTER 4 23
REPRODUCING KERNEL PARTICLE METHOD . 23
4.1

Wavelet Theory . 23

4.2

Definition .. 25

4.3

Formulation 25

4.4

Window Function .. 27

4.4

Boundary Correction Term 29

4.5

Shape Function .. 34

4.6

Choice of Window Function . 36

4.6.1

Gauss, Cubic Spline Kernels .. 36

4.6.2

Lagrange Kernel . 39

4.7

Solution Procedure 41

CHAPTER 5 44
RKPM IN TWO DIMENSIONS . 44
5.1

Two Dimensional Formulation . 44

5.2

Boundary Correction Term in 2D.. 45

5.3

Window Function .. 48

5.4

Solution Procedure . 50

CHAPTER 6 55
NUMERICAL EXAMPLES ... 55
6.1

One Dimensional Examples .. 55

6.1.1

Axial Deformation of a Bar ... 55

6.1.2

Deformation of a Dirichlet Bar .. 59

6.2

Numerical Example Two Dimensional Poissons equation 60

6.3

Numerical Example 2D Laplaces equation ... 63


vii

6.4

Results Discussion 67

CHAPTER 7 69
CONCLUSION ... 69
REFERENCES 71

viii

LIST OF ABBREVIATIONS

CAD

Computer Aided Design

DEM

Diffuse Element Method

EBC

Essential Boundary Condition

EFG

Element Free Galerkin method

FDM

Finite Difference Method

FEM

Finite Element Method

FVM

Finite Volume Method

MLPG

Moving Least squares Petrov-Galerkin method

MLS

Moving Least Squares

MLSPH

Moving Least Squares Particle Hydrodynamics

MLSRK

Moving Least Squares Reproducing Kernel

NBC

Natural Boundary Condition

PDE

Partial Differential Equation

PUFEM

Partition of Unity Finite Element Method

PUM

Partition of Unity Method

RKEM

Reproducing Kernel Element Method

RKPM

Reproducing Kernel Particle Method

RPIM

Radial Point Interpolation Method

SPH

Smoothed Particle Hydrodynamics

ix

LIST OF FIGURES

Figure 4.1 Illustration of the Gaussian and cubic spline kernel functions........................ 38
Figure 4.2 Shape functions of nodes 1, 10, 21.................................................................. 39
Figure 4.3 Shape function derivatives of nodes 1, 10 and 21 ........................................... 39
Figure 5.1 Shape function of node 25 using Lagrangian kernel ....................................... 49
Figure 5.2 Shape function of node 13 using Lagrangian kernel ....................................... 49
Figure 5.3 Flowchart depicting 2D RKPM solution procedure........................................ 54
Figure 6.11D bar subjected to body force......................................................................... 55
Figure 6.2 Axial deformation of a bar - RKPM (Gauss kernel) vs analytical solution .... 58
Figure 6.3 Axial deformation of a bar - RKPM (cubic spline kernel) vs analytical
solution.............................................................................................................................. 58
Figure 6.4 Axial deformation of a bar - RKPM (Lagrange kernel) vs analytical solution59
Figure 6.5 1D bar subject to body force with Dirichlet boundaries on both ends ............ 59
Figure 6.6 Axial deformation of a 1D Dirichlet bar: RKPM solution vs Analytical
solution.............................................................................................................................. 60
Figure 6.7 Temperature distribution over a 2D plate by Poissons equation, solved using
RKPM with 25 nodes........................................................................................................ 62
Figure 6.8 Temperature distribution over a 2D plate by Poissons equation, solved using
RKPM with 36 nodes........................................................................................................ 63
Figure 6.9 RKPM vs Analytical solutions for 25 nodes .................................................. 65
Figure 6.10 RKPM vs analytical solution for 100 nodes.................................................. 65
Figure 6.11 RKPM vs analytical solution for 625 nodes.................................................. 66
Figure 6.12 RKPM vs analytical solution for 2500 nodes................................................ 66
Figure 6.13 Logarithmic plot of L2 norm of error vs nodal density for RKPM using the
Lagrange kernel, cubic spline kernel and a corrected collocation method ....................... 67

LIST OF TABLES

Table 6.1 Dilation parameter and radius of support of kernels.57

xi

CHAPTER 1

INTRODUCTION

1.1

Background

A vast majority of engineering problems require the solution to partial differential


equations. In most cases, analytical solutions of these PDEs are not feasible. As such,
here arises a necessity to approximate these complex PDEs with the aid of numerical
methods. The dawn of computational mechanics was triggered by an application in the
field of aeroelasticity by Frazer [1] in the early 1920's. Frazer introduced the concept of
applying matrix methods to engineering problems. Yet, flexibility matrices were
calculated manually until 1957 when a team of scientists from IBM invented the first
high level programming language, FORTRAN I, which led to the era of computational
mechanics.
More modern numerical schemes for solution to PDEs make use of the Method
of Weighted Residuals [2]. Conventional methods implement the principle of weighted
residuals by assigning a set of nodes and establishing a fixed connectivity (mesh)
between them during the stage of problem definition. These mesh-based methods have
been very popular especially for their mathematical simplicity and computational
efficiency. Well known techniques such as the Finite Difference (FDM), Finite Volume
(FVM) and Finite Element Methods (FEM) are ideal examples of mesh-based methods.

Finite difference methods are very popular for their application in simulation of fluids
with regular node distribution, but their major disadvantage is that they are not very
effective for irregular meshes. The Finite Element Method (FEM) (the term introduced
by Clough in 1960), emerged by the end of 1970s and has, by far, been the most
successful of these mesh-based methods. However, it is not without its own
shortcomings.
1.2

Limitations of Finite Element Method

Despite its plethora of applications and significant advantages, the finite element
method has its own disadvantages which limit its efficiency in meeting the growing
demands of the computational world. One of its biggest drawbacks is that it is a meshbased method. Firstly, development of the mesh alone takes considerable amount of time,
is an expensive process and requires more skill for complex geometry. Mesh resolution
needs to be very fine in problems involving large parameters or where localized
characteristics are different. FEM is not suitable for solving many real-life problems with
moving discontinuities such as crack propagation. In order to gain accurate results for
problems involving large parameters, the mesh resolution needs to be high. FEM
generally uses only linear or quadratic shape functions to determine elemental properties.
Other problems encountered are those due to mesh entanglement and mesh distortion,
where the defined representation is not exactly compatible with the physical problem.
FEM works acceptably fine with static problems where the same mesh is used throughout
the solution procedure. Difficulties arise when applied to constantly changing geometry
or moving boundaries, in which case adaptive remeshing is required. A new mesh would
be required for every change encountered, which makes the entire process time-

consuming and computationally expensive. Adaptive remeshing also hosts a number of


challenges. It requires complex data structure and solution algorithms. Numerical errors
are introduced when state variables are mapped from the old mesh to the new mesh. The
procedure cannot be easily implemented for complex jobs such as fluid-structure
interaction, impact, and explosion problems.
These setbacks became evident when the complexity of problems faced by
engineers was increased and better methods were required. The difficulties in applying
FEM to solve problems with moving boundaries, large deformations etc. as well as the
multitude of problems associated with meshing motivated the development of a new class
of numerical schemes called meshfree methods.

1.3

Meshfree Methods

Definition:
As the name suggests, meshfree methods are a class of numerical methods to
solve partial differential equations based on particle distribution over the domain without
the need for a preset grid/ mesh. In meshfree methods, information regarding the relative
position of the particles is collected during runtime thus alleviating the need for the
process of mesh generation. The definition for meshfree methods can also be stated as a
generalization of Odens and Duartes words on the hp Cloud Method [3]: Meshfree
methods aim to construct accurate solutions to problems involving PDEs without the
need of explicitly partitioning the domain of interest in many smaller suitable
subdomains, as is done in traditional numerical schemes like FEM, FDM or FVM.
Methodology:
The basic methodology of most meshfree particle methods is to represent the

domain by a collection of particles or nodes, xj. Weighting functions, that have


compact support are defined and applied to the nodes. The support of weight functions
are the regions or sub-domains over which the weight function value is non-zero; outside
the support the value of the weight function diminishes to zero. Thus, each node or
particle in the domain has its own area of influence within which its weight function is
non-zero. This region is called the window of support or domain of influence of the
node [4]. The support windows could be of various arbitrary shapes and sizes, but
generally the size and shape of the support has effects on the consistency and stability of
the method. Usually, discs or rectangles are the commonly used shapes in two
dimensions. The weighting function is used to develop a partition of unity. This leads to
the formulation of an approximation to the given function, which serves to solve the
given partial differential equation. The shape functions of meshfree methods do not
generally satisfy the Kronecker delta property which is the reason why they are called
approximation functions rather than interpolation functions, as in FEM. In FEM,
interpolation functions are developed based on the mesh which defines the connectivity
between elements or nodes, whereas in most meshless methods, the shape functions/
interpolation functions are developed for the particular point of interest.
The objective of meshless methods is to overcome all of the shortcomings of
conventional grid-based methods and to provide efficient computational methods to meet
the growing demands of challenging simulation problems in scientific computing.
1.4

Advantages and Limitations of Meshfree Method

Meshfree methods host a number of advantages over their mesh-based


counterparts; the most important advantage is attributed to the absence of a mesh. The

entire time and cost of mesh generation is eliminated. Since the domain can be
represented by a collection of particles it becomes easier to provide an accurate
representation of complex geometries. The methodology can be more easily linked to
CAD because there is no element-based mesh [5]. Adaptive remeshing is not required,
which gives meshfree methods an advantage over other mesh-based schemes, especially
in problems involving moving boundaries and growing discontinuities. The fact that
connectivity is calculated at run-time and can be changed any time makes meshfree
methods ideal for computation of large deformations by eliminating any mesh
entanglement problems. Also, the support size of meshfree interpolant is much larger
than FEM interpolant which implies that
(i) for the same support size the meshfree interpolant covers more particles than
the FEM interpolant and
(ii) for the same particle density the meshfree interpolant has a larger support size
than the FEM interpolant.
This makes the meshfree interpolant ideal for large deformation problems [6]. Accuracy
can be increased at desired areas by simply adding more nodes in the region of interest.
Another motivating feature of meshfree methods is the continuum advantage. The
solutions and its derivatives are continuous in meshfree methods, unlike other numerical
schemes. For instance, the Reproducing Kernel Particle Method (RKPM) is capable of
reproducing a polynomial of order n exactly; given a polynomial basis of order n.
Meshfree methods can provide even C interpolations, which makes them desirable for
simulations and problems involving higher derivatives.
Despite all these merits, meshfree methods have challenges that yet have to be

resolved. The mathematical complexity of meshfree methods has not yet been completely
understood as many of the methods are still considered intuitive. Efforts have been
made by Monaghan [7] to provide mathematical basis for the meshfree shape functions;
he has shown that these functions are in fact representations of well-known interpolation
functions. Yet, in the view of many researchers, there the mathematical framework of
meshfree methods is still fraught with uncertainties.
Meshfree methods are considerably time consuming and also require extensive
computational effort. They involve a large number of small matrix-matrix operations and
inversions of huge moment matrices. One other important factor is the difficulty in
employing the essential boundary conditions as the shape functions do not always satisfy
the Kronecker delta property. An important impediment is numerical integration of the
Galerkin weak form.
Several improvisations are being made to existing meshfree procedures to
overcome these challenges. As JS Chen [8] points out in his preface, research advances
have been made in the mathematical framework of meshfree methods, enhanced
treatment of boundary conditions, hierarchical meshfree approximation, stability analysis
of meshfree methods, fast and stable domain integration techniques. The problems with
nodal integration have been addressed and several nodal integration techniques applied to
meshfree methods have been analyzed and discussed in detail in a very recent paper by
Puso et al [9].
Meshfree methods have proven to be the most plausible solution to problems not
solvable by traditional FEM such as problems involving large deformation, moving
discontinuities, evolving material interfaces, and multiple-scale phenomena. The

methodology to apply meshfree methods such as RKPM to computational fluid dynamics


problems has been explored in detail by Liu et al in [10].

CHAPTER 2

MATHEMATICAL BACKGROUND

2.1

Method of Weighted Residuals

The weighted residual methods are a class of variational methods that seek an
approximate solution to differential equations, expressed as a linear combination of
approximation functions and unknown parameters.
Any differential equation can be written in a weighted integral form, wherein all
the terms of the differential equation are moved to one side (such that the other side of
the equation reads zero) called the residual, R, the entire equation is multiplied with a
weight function w, and integrated over the domain.

w( x) R dx = 0

(2.1)

Consider a differential equation given by

D(u ) = f

(2.2)

In the weighted residual method, the function u is approximated by a function uh as


n

u h ( x) = j ( x) d j + 0 ( x),
j =1

where 0 and j are approximation functions


d j is an unknown parameter

(2.3)

The difference that arises between the values obtained by substituting the
approximate solution uh(x) in the differential equation and the value of the known
function of the independent variables, f, is called the residual.
R = D(u h ) f

(2.4)

The unknown parameters dj are computed by requiring the residual, R, to satisfy


Eqn. (2.1). The weight functions, w, are generally not equal to the approximation
functions, .
There are several types of weighted residual methods [11] such as the PetrovGalerkin method, the collocation method, least squares method, sub-domain method,
method of moments, and the Galerkin method. As will be demonstrated later in this
thesis, the Galerkin method is the basis for the Reproducing Kernel Particle Method
formulation. When the weight function is chosen to be equal to the approximation
function, the method is called the Galerkin method, i.e. w(x) = (x). In this case Eqn (2.1)
becomes :

( x) {D( d
i

2.2

+ 0 ) f }dx = 0

(2.5)

Basis Functions:

A basis of a vector space is defined as a subset of linearly independent vectors in the


vector space that span the vector space.
Examples of monomial and quadratic basis functions in one and two dimensions are
given below:

10

Monomial:

Quadratic:

1
P ( x) =
x

1

P ( x) = x1
x
2

(2.6)

1

P T ( x) = x
x 2

1
x
1
x 2
P T ( x) = 2
x1
x1 x 2
2
x 2

(2.7)

2.3

Boundary Conditions

In this context, boundary conditions are additional restraints on the governing differential
equation that have to be satisfied to arrive at the solution to the problem at hand. There
are two types of boundary conditions that are considered in this section. When the
boundary condition specifies the value of the dependent variable at the boundary, it is
called an essential boundary condition. When the value of the derivative is specified at
the boundary, the boundary conditions are called a natural boundary condition. Thus,
essential boundary conditions specify the value for primary variables of the problem,
while natural boundary conditions specify the value for secondary variables. Essential
boundary conditions are also called Dirichlet boundary conditions, where as natural
boundary conditions are referred to as Neumann boundary conditions.
Examples include:
Essential boundary conditions, u = u on
u
Natural boundary conditions, = q on
x

11

CHAPTER 3

CLASSIFICATION OF MESHFREE METHODS

The task of classifying meshfree methods is quite perplexing given the variety of
methods which have similar roots or ideas but are quite different from each other. Fries
[11] has, to the best of the authors knowledge, given a comprehensive classification of
meshfree methods. Fries charts three classifying steps of a meshfree method: the
methodology used to construct the partition of unity, the choice of an intrinsic or extrinsic
basis for the approximation and the choice of test functions. Both mesh-based (FEM) or
meshfree (MLS, RKPM) principles or a combination of both can be used to derive the
partition of unity of nth order. If the partition of unity function is directly taken to be the
shape function for the approximation, an intrinsic basis is used; else if a different
approximation is desired, an extrinsic basis is used.
Another way of classifying meshfree particle methods is based on physical
principles and computational formulation. Liu and Li [5] adopt classification based on
computational formulation by looking at whether the method is based on the strong (eg:
SPH, vortex method, collocation method, generalized finite difference method) or weak
(RKPM, EFG, hp cloud, RKEM, Point Interpolation Method, Finite cloud method, Finite
Point Method, RPIM, MLPG, Partition of Unity method etc.) form of the partial
differential equations. Several methods based on the weak form require a background

12

mesh for integration, which does not make them completely meshfree. Methods like the
finite point method that have been developed from the strong form of the differential
equation based on the Taylor series of the function, thereby not needing a background
mesh. But these methods compromise on stability and accuracy, which is not very
desirable.
The weighted residual method adopted to develop the approximation can also be
used to classify meshfree methods. The Galerkin method is very commonly employed,
eg. in EFGM, RKPM etc. The collocation method, as the name suggests uses the
associated weighted residual method. MLPG is based on the Petrov-Galerkin
formulation. Many other methods can be classified under these categories.
Classification can also be based on the way the approximation function is
represented in the methods. A moving least squares approximation is fairly common and
easy to formulate (DEM, EFGM, MLPG, MLSRK, MLSPH etc.). The point interpolation
scheme proposed by Liu et al [12] is another representation (Point Interpolation Method,
Radial Point Interpolation Method). The reproducing kernel particle method and
smoothed particle hydrodynamics are based on an integral representation of the
approximation function. There are also other representations adopted by methods such as
the partition of unity method, hp cloud method, etc.
A popular, yet simple way of grouping meshfree particle methods is to classify
the procedure as based on a set of disordered points (eg: SPH) or on a Moving Least
Squares Method (MLSM) (eg: DEM, EFGM).
Among the existing meshfree methods, the element free Galerkin method and the
reproducing kernel particle method are consistently viewed as the best technique for

13

structural analysis [13].

3.1

Smoothed Particle Hydrodynamics (SPH)

Astrophysical phenomena such as stellar evolution and dust clouds laid the
foundation for the invention of meshless methods, the most important reason being that
these systems could be more appropriately represented by a collection of particles than
by other means. The harbinger in modeling these cosmological phenomena was the
Smoothed Particle Hydrodynamics (SPH) method framed independently by Lucy and
also by Monaghan et al. [14]. This was very different from the then predominant
Particle-In-Cell method, which is more of a grid-based method [15]. The most attractive
feature of the SPH is the free Lagrange concept. SPH was primarily a collocation
technique and used interpolants that could be directly differentiated, as the interpolation
functions were based on separate particles. The reason that a mesh is not necessary is
that SPH uses information from discrete particles and an interpolation kernel to
approximate a function [16]. Monaghan [17] explains the SPH method from a
physically understandable perspective, explaining the fundamental concept of the
interpolation functions developed from a kernel standpoint and illustrating it by using
this idea to reduce physical governing partial equations, such as the energy and
momentum equations into a set of ordinary differential equations. He even laid down
the golden rules of SPH, stating first that although spline kernels are computationally
more efficient than exponential kernels, the latter relates more to the physical meaning
of the problem. (The concept of kernels will be discussed in detail in later sections).
This is convincing enough, considering the fact that SPH was originally developed to
solve physical problems and this interpretation would help correlate to the solution

14

much more lucidly.


The kernel function used to reproduce the solution over the given domain is
required to satisfy the following conditions [4]
i.

ii.

( x, h ) 0

( y, h) d

(3.1)

=1

(3.2)

Rn

iii.
iv.

( y, h) ( y ), as h 0
( y, h) C P (R n )

(3.3)
(3.4)

As implied by condition (iv), the kernel should be differentiable at least once since its
derivatives should be continuous to avoid large fluctuations in the force applied to the
particle. The continuity requirement and the resultant shape function smoothness is the
reason why the method is named Smoothed Particle Hydrodynamics (SPH) [5].
SPH is based on the Lagrangian particle method. The basic principle of SPH is to
represent the domain by a set of points, with a mass attached to each of them, the points
are referred to as particles. Using the information carried by these particles (for
instance: velocity, energy etc.), the forces acting on an element would be constructed by
considering integral interpolants [17]. This allows smoothed approximations of the
physical properties in the domain.
Material tracking traditionally obtained by Lagrangian methods and the freedom
of flow obtained with only Eulerian methods are both offered by SPH, making it very
desirable. Though SPH was one of the first and very popular meshfree methods, it had

15

some major disadvantages. Since SPH was developed for large unbounded domains
encountered in astrophysical problems, it does not work well when applied to domains
with boundaries, including stationary or moving boundaries. The equations achieve
remarkable accuracy in the interior of the domain, but deteriorate near the boundaries
due to the failure to meet reproducing conditions of even the lowest order at the
boundaries. This failure is introduced by the intrinsically nonlocal SPH interpolants and
by the fact that the shape functions do not form a partition of unity. This leads to a lack
of consistency in the boundary regions and spurious boundary effects [18]. Other
problems associated with SPH are the so-called tension instability, which is a
consequence of using Eulerian kernels in a Lagrangian collocation scheme, [19] and
rank deficiency of the discrete divergence operator [19]. Also, the number of particles
has to be large to obtain accurate results.
Several groups worked on enhancements SPH either by improving its
consistency or its stability. Efforts were taken to overcome problems such as the tensile
instability and spatial instabilities. Libersky and Randles [16] published an extensive
research paper on their work summarizing their efforts to improve the SPH method.

3.2

Corrected Smooth Particle Hydrodynamics

This method proposed to overcome two major disadvantages of SPH by


incorporating certain corrections to its formulation. The accuracy of SPH, especially
near boundaries, is improved by making corrections to the interpolating kernel function.
Also, the tensile or spatial instabilities that arise due to errors resulting from point
integration are addressed by an integral correction [20].

16

3.3

Moving Least Square Particle Hydrodynamics (MLSPH)

Dilts [21] developed the Moving Least Square Particle Hydrodynamics method
inspired by a comment by Liu et al. in [22]. Here, an MLS interpolant is used as the
kernel function in SPH. Traditional SPH employs the kernel approximation and then
the particle approximation to develop an approximation to the equations of motion.
Dilts instead uses the SPH kernel functions as a set of trial functions for the formulation
of a Galerkin approximation to the Lagrangian partial differential equations and then
employing kernel approximation which results in a collocation-Galerkin approximation.

3.4

Moving Least Squares Method

A new form of meshfree methods arose in the early 90s based on the Moving
Least Squares (MLS) method. These were basically Galerkin methods that depended on
the weak form of the PDE. The concept of least squares is well-known and is widely-used
in the area of curve-fitting. Moving least squares is analogous to the motion of moving
average. The method of MLS was first applied to generate two dimensional interpolants
in a lower-order case by Shepard. Hence the lowest order MLS approximations are
called Shepard functions. Lancaster and Salkauskas [23] introduced a general higher
order approach for data smoothing and interpolation. Though Lancaster and Salkauskas
had given a detailed introduction to the theory behind the moving least squares method,
they had not implemented it into practical applications wherein many other complexities
concerning computational efficiency and convergence would have been encountered [24].
The methodology of the MLS approximation makes use of the fact that any function (in

17

our case, the approximation function) can be expressed as the product of basis functions
and a set of coefficients.
A weighted least squares fit is applied to the approximation and its extremum
leads to the evaluation of the unknown coefficients. Finally, the approximation is
represented in terms of its shape functions. The MLS procedure bears similarities to the
reproducing kernel particle method in many aspects. However, both methods make use of
unique ways to arrive at a partition of unity.
The moving least squares procedure was first employed to obtain meshfree
approximations by Nayroles et al. [25], who named the technique Diffuse Element
Method (DEM). The constant interpolation coefficients assumed by DEM reduces the
accuracy of the method. The procedure was then improved by Belytschko et al [27] by
including certain terms in the interpolant derivatives that were omitted in DEM.

3.5

Element Free Galerkin Method

The Element Free Galerkin method (EFG) was developed based on the MLS
approximation by Belytschko et al. [27]. In the EFG method, a set of nodes are
distributed over the domain and boundaries. The moving least squares approximation is
employed to approximate the solution at the nodes. A Galerkin weak form of the
governing partial differential equation is developed. The resultant shape functions from
the MLS approximation do not satisfy the Kronecker delta property, which yields in a
difference between the approximated solution and the original solution on the boundaries.
In order to fulfill the essential boundary conditions, the method of Lagrange multipliers is
incorporated into the development of the weak form of the PDE. EFGM requires a

18

background mesh and an additional grid of Gauss quadrature points. The requirement of a
background mesh limits the truly meshfree characteristics of EFGM.
A very high rate of convergence was achieved in EFG method by adding more
accurate derivatives and by using the Lagrange multipliers to enforce the boundary
conditions [6]. Problems related to volumetric locking are not encountered in EFG. One
drawback of the EFG method is that it is computationally expensive to construct its
complex shape functions, unlike FEM whose polynomial-based shape functions are easy
and economical to construct. Krysl and Belytschko [28] have developed a library called
the ESFLIB to demonstrate research advancements to make EFGM computations
efficient. Their paper gives a straightforward way to program the Element Free
Galerkin shape function construction by providing an abstract form of the EFG shape
function construction algorithm. But, there are certain nodal arrangements that could

break down the algorithm used to construct the MLS functions [29].

3.6

Meshless Local Petrov-Galerkin Method

The Meshless Local Petrov-Galerkin (MLPG) method is also based on the


weighted residual methods. In the Petrov-Galerkin formulation, the weight functions and
trial functions can be chosen differently. Like other particle-based methods, MLPG
makes use of a set of nodes distributed over the domain, but the difference lies in the fact
that the weighted integral form represents a local sub domain. Thus, the weak form of the
differential equation is satisfied node-wise locally, independent of other nodes. As such,
the background mesh is only needed for local quadrature, thereby overcoming the
difficulty of having to construct a background mesh for the entire domain. Thus, MLPG

19

can be considered a truly meshfree method in comparison with EFGM [12].


Yet, MLPG has its disadvantages. The Petrov-Galerkin formulation results in
asymmetry of the system matrices and also makes the local integration of the complex
Petrov-Galerkin integrand very difficult.

3.7

Finite Cloud Method

The finite cloud method, proposed by Aluru et al. [30], can be considered an
improvisation of the reproducing kernel particle method. Hence, the reader is strongly
urged to return to this topic after perusing the section on the methodology of RKPM. The
basic idea of the finite cloud method is to replace the classical reproducing kernel with a
fixed reproducing kernel, centered at a particular point (xk , yk) and hence has a slight
difference in the definition of its correction and kernel functions. The method defines a
cloud associated with each point that is finite in extent and encompasses enough
number of points such that the moment matrix is not singular. The interpolation functions
associated with a kernel centered at a particular point have values at nodal positions
inside the cloud, but are zero at the nodes outside the cloud. The paper illustrates the
multi-valued nature of these interpolation functions at places where two clouds overlap.
The interpolation functions hence have to be restricted to a single point by using a point
collocation approach. The shape functions are made to satisfy the partition of unity by
fixing the kernel at every point in the domain to compute all the required interpolation
functions, whose sum becomes unity.
The paper claims that because the moment matrix turns out to be constant within
each cloud, the computational burden is reduced as compared to traditional reproducing

20

kernel methods. But, the finite cloud method, which is very similar to RKPM, is not yet
as popular as the latter.

3.8

Partition Of Unity Methods

Additionally, there exists another class of meshfree methods that are based on the
partition of unity property. Shape functions that satisfy the partition of unity have nonzero values inside their domain of influence, and are zero outside the domain.

Partition of Unity Finite Element Method:

The Partition of Unity Finite Element Method (PUFEM) was developed by


Babuska and Melenk [31] to solve problems where traditional FEM proved
disappointing, such as problems where the solution is rough (or highly oscillatory),
problems of singularly perturbed type or where the solution exhibits a boundary layer and
problems that require frequent remeshing. Their approach was to proceed with a
formulation that is developed using variational techniques and then design the trial (and
test) spaces according to the particular problem at hand [31].

Intrinsic Partition of Unity Method (PUM):

One of the more recently introduced meshless methods is the intrinsic partition of
unity method proposed by Belytschko and Fries [32]. The method eliminates the need for
extrinsic enrichment functions. Unlike the standard partition of unity methods where
classical finite element shape functions are taken as the PUM shape functions, the
moving least squares technique is employed to obtain special shape functions based on an

21

intrinsic basis vector and weight functions based on FEM shape functions. The
approximation is intrinsically enriched, which explain the methods name. As in
standard MLS procedure, since a weighted least square functional is minimized to obtain
the shape functions, the weight functions have compact supports [32].

3.9

h-p Cloud Method

The hp cloud method developed by Oden and Duarte [3] is a very popular
partition of unity based method. The method hosts some of the desirable features of the
hp finite element method, without the need for a mesh. The problem domain is pictured

to have a set of arbitrarily scattered nodes. A number of balls are constructed, each
centered on a node and an open covering within the domain, consisting of N number of
balls is analysed. From this, a signed partition of unity is developed. This procedure has
been detailed in [3] in an axiomatic approach as well as in a constructive moving least
squares approach. The partition of unity functions thus developed serve as the incredibly
smooth basis functions for the hp cloud method. An important step is the construction of
the family of functions, which is low cost, from the partition of unity functions. The
signed partition of unity when multiplied by polynomial or other appropriate functions
give rise to functions called the hp clouds.
The overall solution procedure of hp clouds can be outlined as follows: The
discretized equations are obtained using the usual Galerkin procedure and the essential
boundary conditions are imposed using Lagrange Multipliers, as in EFGM. A
background cell structure is required to perform the domain integration and then the
family of functions is used to build the discrete approximations [33].

22

Though the hp cloud method proves advantageous in hosting h and p adaptivities,


there are issues relevant to the domain integration that cannot be overlooked.

Adaptive and Multiresolution Analysis: Meshfree methods are ideal for adaptive

procedures and problems involving multiresolution analysis. Several studies have been
undertaken to improve the implementation of adaptive and multi resolution methods.
Alterations of well-established meshfree methods are the most common. Notable
examples are the Adaptive Galerkin Particle Method by Chen et al. [34] and the study
done by Liu et al [10] of multiresolution analysis for Computational Fluid Dynamics
(CFD) problems.

23

CHAPTER 4

REPRODUCING KERNEL PARTICLE METHOD

The reproducing kernel particle method was inspired by the theory of wavelets.
Research on the application of wavelet theory to develop meshfree approximations was
originally pursued by Liu et al [22]. Before examining the formulation of RKPM, it is
instructive to review a brief introduction to wavelets.
4.1

Wavelet Theory

Wavelets are a family of transformation functions similar to the well-known


Fourier transforms. Joseph Fourier, in the early 1800s discovered that he could represent
complex functions by a combination of sine and cosine functions, which led to the
development of Fourier analysis. Mathematical transformations in general translate
functions from one domain to another domain; eg: Fourier transforms translate functions
from the time domain to the frequency domain and the inverse Fourier transforms do the
exact opposite. These transformations can therefore be employed as filters, so that when a
signal is fed to these filters, they separate the low and high frequencies, remove the
unnecessary frequencies (depending on the type of filter) and pass back the remaining
frequencies which are then reverse transformed.
Wavelet functions are also mathematical transformations, but apart from merely

24

translating the functions from one domain to another, they also analyze according to
scale. Thus both temporal and frequency analysis is carried out when employing a
localized function. Wavelets use scale-varying basis functions; that is, the basis function
divides the given signal at varying scales by separating it into the required number of
divisions and these divisions can then analyze the function at different resolutions [35].
A detailed study on the relationship between filters and the reproducing kernel has
been provided by Chen et al. [36]. Filters basically differentiate and separate frequencies
of the input signal. The band of frequencies that are allowed to pass through the filter is
called the pass-band and the band of frequencies that are removed by the filter is called
the stop-band. A low-pass filter allows frequencies below the stop-band to pass through
it, while retaining the remaining high frequencies. A high-pass filter does the opposite.
When wavelet functions are employed as low-pass filters, they allow the smooth data to
pass through, eliminating the high-frequency spikes. Wavelet functions have been
employed in multi-resolution studies where the scaling function is applied to various
scales. Apart from decomposing the signal into various frequency bands, they also can be
sequenced according to the scales of the wavelet function.
The output signal produced by a given unit impulse input is called the impulse
response of a filter or the filter kernel. The output signal can be represented by a
convolution of the input signal and the filter kernel, similar to the convolution in Fourier
transformation, provided both the input and the output signals belong to the same
function space [36]. This idea forms the basis of the reproducing kernel particle method,
as will be explained in later sections.

25

4.2

Definition

Reproducing kernels can be viewed as a class of operators that reproduce


themselves through integration over the domain [37]. Fourier transforms and wavelet
functions are examples of reproducing kernels. Thus, the reproducing kernel particle
method is a meshfree method that uses reproducing kernels to frame a partition of unity;
this which approximates the given function over a domain and thereby enables a solution
to a partial differential equation.

4.3

Formulation

The reproducing kernel particle method can be viewed from two standpoints: as
an extension to improve the SPH method and also as a method originating from wavelet
theory. We proceed from the latter starting point and in the course of the formulation will
explain the extension to SPH perspective. The formulation adopted below uses the
convolution theorem followed in many papers by Li, Liu et al [22,37,38], although there
is also an alternative approach to carry out the derivation based on the moving least
squares approach [26]. The procedure can also be detailed using a Taylor series approach
[18]. It is observed however, that all approaches give the same resulting approximation
for the given function and hence can be considered equivalent.
Considering the reproducing kernel to be a wavelet filter, we proceed with the
following formulation. We are interested in approximating a given function u(x) over a
domain . The approximation uh (output signal) can be represented by a convolution of
the given function u(x) (input signal) and a window function a (the filter kernel) as:

26

u h ( x) = u ( x) * ( x) = a ( x; x x s ) u ( x s ) d s

(4.1)

Thus, the window function can be viewed as a projection operator which reproduces the
original solution u(x) to a resolution defined by a ( x; x xs ) [38].
Since x is a dummy variable in the transform, it could be replaced with y, giving,
u h ( x) =

( x; x y ) u ( y ) d y

(4.2)

In fact, Liu et al [37] state that all methods involving meshfree Lagrangian
particles concerned with the solution to Eqn. (4.2) are referred to as reproducing kernel
particle methods. Eqn. (4.2) is similar to the notion of SPH formulation, but there is a
subtle, yet important difference due to the presence of a correction term, which will be
elaborated shortly. Thus, an alternative approach to formulate RKPM is to start with the
addition of a boundary correction term to SPH formulation. The reproducing kernels have
compact support and give continuous solutions within the support, thereby satisfying the
well-posedness condition that the input function and the output approximation must lie in
the same domain [36].
It is important to note that the window functions are not exactly wavelets, they are
scaling functions used to produce wavelets [22].
The discretized version of Eqn. (4.2) is given by:
n

u h ( x) = a ( x; x y j ) u ( y j ) w j
j =1

Where, n number of nodes/ particles


w quadrature weights.

(4.3)

27

4.4

Window Function

Before examining the mathematical formulation and properties of the window


function, let us discuss the physical meaning of the window function. Window functions,
or weighting functions (as they are generally called), are common to all meshfree
methods.
The selection of the window function a has significance in the RKPM technique.
Because the window function is a wavelet function that employs scale-varying basis
functions, the orthogonality property must be satisfied:

( x; x p ) a ( x; x q ) d = pq

(4.4)

Another important property of the window function is that its integral over the
domain should yield unity. Equation (4.2) likens the shifting property of the delta
function, given by:
u ( x) = ( x s ) u ( s ) ds

(4.5)

If we use the Dirac delta function, all frequencies of a function would be


reproduced. Unfortunately, the delta function by itself cannot be used in the discrete
version (Eqn. 4.3) because of the absolutely fine mesh refinement it would require.
Hence, we try to develop a smoother window function that is very close to the delta
function, such that its integral over the domain is unity.

( x; x y ) d y = 1

(4.6)

It should be noted that the above integral will be slightly less than unity near the
boundaries, which is a major constraint of SPH and other kernel based methods [22]. In
other words, when the point y is close to the boundary, Eqn. (4.6) is not satisfied as the

28

left hand side becomes less than 1. In order to account for this discrepancy, we
incorporate a factor called the boundary correction term , which is 1 in the interior
domains and a non-zero real number near the boundaries. The window function
represented as a is actually a product of the kernel function and a correction function
. From this point forward, we refer to the kernel function as the window function.
1 x y
a ( x ) = ( a )
,
a a

a>0

(4.7)

Thus, a(x) represents the dilation and translation of the window function [22].
The scaling parameter a dilates the function over the domain. This provides the
necessary refinement or resolution. The refinement/scaling/dilation parameter allows us
to view the kernel as an extended projection operator; it can now be interpreted as a
window function. The scaling function determines the extent to which the window
function confines the major contribution to the reproduced solution to the neighborhood
of the particle y = x [7]. A small value of the dilation parameter implies finer resolution.
The resolution decreases with increasing a. The value of this refinement parameter is
directly proportional to the difference between the reproduced and real solutions.
i.e., as

a 0,

u h ( x) u ( x) 0

The ability to control refinement provided by the dilation parameter enables the
implementation of multiresolution analysis, which is an advantage of RKPM.
As mentioned in the previous section on the theory of wavelets, the window
function is a localized function and hence needs to be translated over the domain. This is
done by inserting the argument (x - y) to the window function, where y denotes the point
at which the window function is centered. This translation term enables RKPM to cover
the entire domain without a mesh, unlike FEM which defines elements for the same

29

purpose. Thus,

1 x y

represents a window function centered at the point y and


a a

with support scaled by the dilation parameter a [22].

4.4

Boundary Correction Term

The correction function is what differentiates RKPM from SPH. Since the integral
of the window function is not unity near boundaries, Liu et al. [22] introduced a boundary
correction function that would compensate appropriately. Thus, the boundary correction
term is unity in the interior of the domain but has a finite value near the boundaries to
compensate for the lack of partition of unity property. This makes the use of boundary
correction term necessary for finite domains.
The integral window transform and the approximation can now be written as:
u h ( x) =

x y
u ( y ) d y
a

(a) a

(4.8)

The discretized form can be rewritten as:


n
1 x yj
u h ( x) = j (a )
a a
j =1

u ( y j ) w j

(4.9)

From the above equation, we can deduce that the shape function, N, is written as a
product of the correction function, window function and the quadrature weight.
n

u h ( x) = N j ( x) u ( y j )

(4.10)

j =1

where,

1 x yj
N j ( x) = j (a )
a a

w j

(4.11)

30

Our goal is to develop a shape function such that only a small number of particles
are required to obtain an acceptable solution. This would increase the accuracy of RKPM
over the SPH method.

4.4.1 Computation of the Boundary Correction Term:


We know that any function can be expressed as a product of linearly independent
functions, P(x), and a set of unknown coefficients, d. P(x) is our set of basis functions in
the following derivation for the computation of the unknown coefficients d

u ( x) = P ( x) d

(4.12)

The use of reproducing kernel as a low-pass filter enables RKPM to employ scalevarying basis functions. An example of linear scale-varying bases:

1
P T ( y) = y x
a

1
x x
P T ( x) = 1

a
x2 x
a

(A superscript T denotes transpose)


In Eqn. (4.12) u(x) is the approximation of the unknown function. Since Eqn.
(4.12) is true for any function it should be equivalently true for both u(x) and u(y). Thus,
Eqn. (4.12) can be rewritten as,
u ( y ) = P( y ) d

(4.13)

Since our kernel is likened to a low-pass wavelet filter that not only transforms a
function, but also scales it, we must employ scale-varying basis functions in our
formulation. Hence, Eqn. (4.13) can be rewritten as,

31

y x
u ( y) = P
d
a

(4.13 a)

Pre-multiplying both sides of Eqn. (4.13 a) with the transpose of the basis vector, PT,

y x
T y x y x
PT
u( y) = P
P
d
a
a a

(4.14)

In order to deduce the unknown coefficients d, we apply an integral window transform to


Eqn. 14. Recall here the definition for reproducing kernels as operators that are capable
of reproducing themselves through integration over the domain.

1 y x
T y x
T y x y x1 y x
P
u
(
y
)

a a a y P a P a a a d y
y
y
(4.15)

Observation of the right hand side of Eqn. (4.15) reveals that it is, in fact , a product of a
moment matrix, M, with a set of constant coefficients, d. Since the integration is with
respect to y, the resultant moment matrix will be a function of x, i.e, M(x), though the
moments, as will be deduced later, are constant in the interior. Hence, Eqn. (4.15) written
in terms of the moment matrix is,

1 y x
y x

u( y)
d y = M ( x ) d
a a
a

(4.16)

Rearranging terms in Eqn. (4.16),


1 y x
y x
d = M 1 ( x) P T
u( y)
dy
a a
a
y

(4.17)

We have now computed the co-efficient, d. Next step is to substitute this


expression for d in Eqn. (4.12) which is repeated here in terms of the scale-varying basis
function.

32

x x
u ( x ) = P
d
a

(4.12 a)

This gives us our kernel approximation of the function u(x),


1 y x
x x 1
T y x
u h ( x ) = P
M ( x) P
u( y)
dy
a
a
a
a

(4.18)

Eqn. (4.18) can be rewritten as,


1 y x
y x
u h ( x) = P (0 ) M 1 ( x) P T
u( y)
dy
a a
a
y

(4.18 a)

Comparison of Eqn. (4.18) with Eqn. (4.8) yields the expression for the boundary term
(x,y),

( x, y ) = P(0). M 1 ( x). P T ( y )

(4.19)

Thus, our expression for the continuous reproducing kernel is,


1 y x
u h ( x ) = ( x, y )
u ( y ) dy
a a

(4.19 a)

The evaluation of the boundary correction term is not complete as yet, because we still
need to simplify Eqn. (4.19). The discretized form of the boundary correction term is
given by:
yj x

( x, y j ) = P(0 )M 1 ( x) P T

(4.20)

The moment matrix has been defined as,


y x y x1 y x
M ( x) = P T
P

dy
a a a a
On writing the basis functions in vector/ matrix form:

(4.21)

33

1
M ( x) = y x 1
a

y x 1 y x

dy
a a a
(4.22)

1
=
yx
a

yx
1 y x
a

2
dy
y x a a

The zeroth, first, and second moments of the window function are given by the following
formulae:
1 y x
dy
a

Zeroth moment (scalar), m0 =

First moment (vector),

m1 =

y x1 y x

dy
a a a

(4.23a-c)

y x 1 y x
Second moment (tensor), m11 =

dy
a a a
2

From Eqns. (4.23 a-c), it becomes clear that elements of the M(x) matrix are the zeroth,
first and second moments of the window function, which thereby proves that M(x) is
indeed the moment matrix.
m
M ( x) = 0
m1

m1
m11

(4.24)

The moment matrix, M, is a gram matrix. The determinant of the moment matrix must be
ensured to be non-zero and positive. This can be achieved by controlling the nodal
distance, x.
det( M ( x)) = m0 m11 m1

(4.25)

Mathematically, recall that the definition of moment matrix involved linearly


independent basis functions. The linear independence of these basis functions ensures

34

that the determinant of the moment matrix is always positive [22]. Since the determinant
is positive, the moment matrix is invertible. The inverse of the moment matrix is given
by,
M 1 ( x) =

m1
m0

m11
2
m0 m11 m1 m1

(4.26)

The boundary correction term is then evaluated by substituting the terms evaluated above
in Eqn. (4.20)
1
y x
a

m11
( x, y ) = [1 0]
2
m0 m11 m1 m1

m1
m0

1
[m11
2
m0 m11 m1

1
y x
a

m1 ]

(4.27)

which on simplification yields,

( x, y ) =

m11
m0 m11 m1

y x

m0 m11 m1 a
m1

(4.28)

Previously published works [22, 37, 38] write the above boundary correction function in
terms of two functions C1 and C2, as follows:
x y

( x, y ) = C1 ( x, y ) + C 2 ( x, y ).

where,

C1 =

(4.29)

m11
m0 m11 m1

C2 =

m1
m0 m11 m1

4.5

Shape Function

Once the boundary correction function is formulated, we could state the shape function

35

formula by substituting all the appropriate terms in Eqn. (4.11)


1 x yj
N j ( x) = j (a )
a a

w j

(4.30)

m11
m1
=

2
2
m0 m11 m1
m0 m11 m1

y j x 1 x y j

a a

w j

Thus, the reproduced solution in Eqn. (4.10) becomes,


n
m11
m1

u h ( x) =

2
2
m0 m11 m1
j =1
m0 m11 m1

y j x 1 x y j


a a a

u ( y j ) w j

(4.31)

4.5.1 Derivatives of Shape Function:


In order to build the stiffness matrix, we require derivatives of the shape function.
One advantage of RKPM is that the derivatives can be obtained by direct differentiation
and, if the shape functions are continuous, the derivatives are continuous as well.

1 x yj

N j ( x) = j (a )
a a

1 x yj
w j + j (a )
a a

w j

(4.32)

The derivative of the correction function is:

( x, y ) =

1
m0 m11 m1

m m m 2 m m

1
11
1
0 11

y x 1

m0 m11 + m0 m11 2m1 m1 m11 m1


+ m1

a a
(4.33)

where,

m0 , m1 , m11 are the derivatives of the zeroth, first, and second moments

respectively.

36

Once the window functions, shape functions, and its derivatives are calculated, the
stiffness matrix can be developed, as in the general Galerkin procedure used in numerical
schemes such as FEM. The integral expression for the stiffness and force terms can be
derived from the weak form of the PDE. The stiffness and force matrices are assembled
and the standard Ku = F equation is solved to obtain the solution approximations. The
entire procedure and the final solution retrieval will be clearly explained in the section on
the algorithm and its implementation.

4.6

Choice of Window Function

4.6.1 Gauss, Cubic Spline Kernels

The window function is a kernel that is used reproduce a given function, by


approximating the delta function. Window functions are also called weight functions. As
mentioned, the weight function is non-zero within its sub domain, but is zero outside of
it. Generally, the support size, shape, and functional form of the window function can be
arbitrarily specified [11]. The window function is preferred to be continuous because
continuity of the shape function depends on it. The choice of dilation parameter has its
effect on the window function such that as the dilation parameter tends to zero, the
window function approaches the delta function.

The choice of window function is mostly arbitrary in mesh free particle methods.
There is a plethora of options for a window function, but the two most commonly used
functions in RKPM are the Gaussian kernel and the spline-based kernels. The Gaussian

37

kernel, with its bell-shaped curve, is easily reasoned to mimic the delta function closely.
The Gauss kernel function is also shown to have high convergence rates, owing to their
exponential convergence [6]. The spline function is also a popular choice because it has
compact support and hence is more practical to apply to finite domains. Examples of
popular kernel functions in one dimension are:

( z) =

Gaussian kernel:

( z) =

and

z=

e ( z ) =
2

1
a

x y

2
3 2 3 3
1 z + z
3
2
4

(4.34)

, 0 z 1

2 1
3
(2 z )
3 4

( z) =
0

( z) =

Cubic spline kernel:

where,

,1 z 2

(4.35)

, otherwise

yx
a

a = 2 m x , m is an integer.

Quartic spline kernel [6]:

C
1 6q 2 + 8q 3 3q 4
n
h
( z) =
0

( z) =

, 0 p q p1

(4.36)

, q 1

where, C is determined from the normalization condition,


q = r/h

and

r = ( x i xi )

The derivatives of the kernel/ window function can easily be obtained by using the chain
rule of differential calculus.

38

Figure 4.1 Illustration of the Gaussian and cubic spline kernel functions

Liu et al. [37] have tested different values of the integer m, to alter the dilation
parameter in the reconstruction of a function and concluded that a value of m = 2 results
in a window function close to the regular finite element shape function.
The following figures depict the plots of RKPM shape functions obtained using
the Gauss kernel. The shape functions are plotted over a one dimensional domain,
= {x : 0 x 6}

The domain is represented by 21 nodes and hence, x = 0.3. The dilation parameters and
support radius of the window functions used for the three kernels are indicated in Table
(6.1). The figures (4.2), (4.3) show the plot of shape functions and their derivatives of
nodes 1, 10 and 21 with respect to the other nodes.

39

Gauss kernel:

Figure 4.2 Shape functions of nodes 1, 10, 21

Figure 4.3 Shape function derivatives of nodes 1, 10 and 21

4.6.2 Lagrange Kernel

In the previous sections the properties of the window function were discussed, of
which orthogonality and partition of unity are of importance. Another important, but nonessential property is the Kronecker delta property. Although the kernels described above

40

and in general, the kernels used by far in RKPM do not satisfy the Kronecker delta
property. By this, it is meant that the entire set of RKPM shape functions over the whole
domain satisfy the partition of unity, but the values of the shape functions themselves do
not become zero at all other nodes, hence the Kronecker delta property is not completely
satisfied. This results in one of RKPMs major disadvantage the difficulty in satisfying
essential boundary conditions.
Special techniques have to be adopted in order to satisfy the essential (or
Dirichlet) boundary conditions of the domain, if kernels that do not satisfy the Kronecker
delta property are used. Several techniques have been developed and implemented to
RKPM. The most notable of these are the Lagrange multiplier method, Nitsches method
and the penalty method. Though these methods seem efficient in solving problems, they
have their shortcomings which deteriorate the effectiveness of the reproducing kernel
particle method. These include the additional computational steps involved in
implementing the method to satisfy the essential boundary conditions, the introduction of
errors by the penalty term and the incorporation of additional, otherwise unnecessary
parameters. The satisfaction of Kronecker delta property is critical to solving dynamic
problems. Hence, the Nitsches method and other similar methods would cause
difficulties when dealing with dynamic problems.
The basis for choosing these kernels, as cited by most prior research, is their
smooth behavior; however it is essentially a completely arbitrary choice. Yet, there exists
a lack of enough justification for the emphasis laid on the choice of a smooth kernel, at
the cost of extra computational burden on imposing the essential boundary conditions.
Hence, in this research we examine the feasibility of employing kernels that satisfy the

41

Kronecker delta property to the reproducing kernel particle method.


The option of using a simple divided difference kernel was explored. Our
immediate choice was the Lagrange kernel, which is most widely used in the finite
element method.
In one dimension,

x xi
h

= 1

where, h - nodal distance


The Lagrange kernel satisfies the Kronecker delta property. Thus, when a Lagrange
kernel is used to construct the RKPM shape function, the value of the shape function
vanishes in all other nodes. Hence, when this kernel is used in the RKPM technique, no
special schemes have to be incorporated in order to satisfy the essential boundary
conditions. Once the stiffness matrix and force vector are computed, the imposition of
essential boundary conditions is done in the same way as in the traditional finite element
method.
The Lagrange kernel, with its Kronecker delta property, would also prove
effective in solving dynamic problems, as opposed to the implementation of boundary
conditions using methods based on a penalty term.
4.7

Solution Procedure

A brief outlook at the solution procedure is given here. A detailed description


with a step-by-step flowchart is provided during the description of a two dimensional
solution procedure, discussed in the later sections. The solution algorithm for a onedimensional problem is:

42

Input the dilation parameter, support radius, nodal co-ordinates, and other
constants, according to the problem at hand
Establish the Dirichlet boundary nodes
Set up Gauss-Legendre points over the domain
Loop over each Gauss point
o Search for all the nodes, under whose supports the Gauss point falls by

using the formula xnorm =

(x

xg )

o Loop over each of the associated nodes, i


 Calculate the boundary correction term and derivative
 Define the window function (Lagrange kernel) and derivative
xi x g
 Develop the shape function Ni
a

and its derivative

 Assemble force vector


 Loop over every other associated node, j

Repeat all steps as for node i, to develop the shape function


x j xg
Nj
a

and its derivative

Assemble stiffness matrix




End j loop

o End i loop

End Gauss point loop


In the stiffness matrix, set the terms on the rows and columns corresponding to the
essential boundary nodes equal to zero, except the diagonal terms on those

43

rows/columns, which are set to one


In the force vector, set the terms on the rows corresponding to the essential
boundary nodes equal to the specified boundary conditions at those nodes
Solve [K ] {u} = { f }
xi x j
Compute the shape functions of every node w.r.t. every other node N
a

Once the nodal parameters, u, are obtained, build the approximation according to
Eqn (4.10)

44

CHAPTER 5

RKPM IN TWO DIMENSIONS

5.1

Two Dimensional Formulation

The reproducing kernel particle method can be quite easily extended to two
dimensions, using the same fundamental principles as the single dimensional derivation.
Similar to the 1D case, the solution approximation to a two dimensional problem can be
represented as a convolution of the given function and the window function [38]:
u h ( x , y ) = u ( x , y ) * ( x , y )
= a ( x , y : x x s , y y s ) u ( x s , y s ) d

(5.1)

In terms of the boundary correction term, the above equation can be rewritten as,
1 x xs y y s
u h ( x, y ) = (a, x, y )
,
u ( x s , y s ) d
a a
a

(5.2)

This gives the continuous form of the approximation. The discretized form of the
equation is,
u h ( x, y ) =

N (x, y )u (x
n

j =1

sj

, y sj )

(5.3)

In the discretized form, the shape function is given as,


1 x x sj y y sj
N j ( x, y ) = (a, x, y )
,
a a
a

w j

(5.4)

45

5.2

Boundary Correction Term in 2D

The construction of the boundary correction term in two dimensions is very


similar to that of one dimension, except for a few changes to the basis function and the
moment matrix.
Let us recall that basis function in two dimensions would be:

Linear:

x xj y yj x xj
=
P
,

a a
a
y yj

Bilinear:

x xj

a
x xj y yj
y yj
=
P
,
a
a
a

x x j y y j
a a

(5.6)

Quadratic:

1
x x
j

a
y y
j

a
x x j y y j x x 2
j
=
P
,

a
a
a

x x j y y j
a a

y y 2
j

(5.7)

(5.5)

46

The moment matrix now has additional terms because of the incorporation of an
additional dimension. If a linear basis is opted, the moment matrix will be:
m00 m10 m01
M ( x, y ) = m10 m20 m11
m01 m11 m02

(5.8)

For a bilinear basis, the moment matrix would be,


m00
m
10
M ( x, y ) =
m01

m11

m10 m01 m11


m20 m11 m21
m11 m02 m12

m21 m12 m22

(5.9)

The elements of the moment matrix are defined by,


1 x xs y y s
,
ds
a
a
x xs 1 x xs y y s

m10 =
,
ds
a a a
a
y y s 1 x xs y y s

,
m01 =
ds
a a a
a

m00 =

x xs y y s
m11 =

a a

1 x xs y y s
,

ds
a
a a

x xs 1 x xs y y s
=
,

ds
a
a a a
2

m20

y y s 1 x xs y y s
,
m02 =

ds
a
a a a
2

x xs y y s 1 x xs y y s
,
=


ds
a
a a a a
2

m22

(5.10 a-g)

The moments are integrated by using the composite Trapezoidal rule, which is mentioned
below for reference.

47

b d

f (x, y ) dA = f (x, y ) dy dx
R

a c

1
= x y [ f (a, c ) + f (b, c ) + f (a, d ) + f (b, d ) +
4
m 1

m 1

m 1

m 1

i =1

i =1

i =1

i =1

2 f ( xi , c ) + 2 f ( xi , d ) + 2 f (a, y i ) + 2 f (b, y i ) +
n 1 m 1

4 f ( xi , y i ) ]

(5.11)

j =1 i =1

where, the domain R = {( x, y ): a x b, c y d }


m + 1 = number of nodes in x direction
n + 1 = number of nodes in y direction

For a linear basis, the determinant of the moment matrix is calculated as:
det = m00 (m20 m02 m11 m11 ) m10 (m10 m02 m01 m11 ) + m01 (m10 m11 m01 m20 )

(5.12)

The boundary correction term hence becomes,


x xs y y s
,

a
a

( x, y ) = P(0) M 1 ( x, y ) P T
=

1
x xs
y ys
2
(m10 m02 m01 m11 ) +
(m10 m11 m20 m01 )
m02 m20 m11
det
a
a

(5.13)
If the correction function is represented as,
x xs
y ys
+ 22

a
a

( x, y ) = 1 + 21

21
22

(m

m20 m11
det
(m m m11m01 )
= 02 10
det
(m11m10 m20 m01 )
=
det

where, 1 =

02

(5.14)

48

5.3

Window Function

The two dimensional window function is obtained by multiplying the 1D kernels


in each of the dimensions as follows,
x xs y y s
x xs y y s
,
=

a
a
a a

(5.15)

For instance, the Gauss kernel would be,

1
x xs y y s
,
e
=
a 2a 2
a

2
2
1 x xs y y s

+

2 a a

(5.16)

The 2D Lagrange kernel can be defined by:


x xs
y ys
x xs y y s
,
= 1
1

a
a
a
a

(5.17)

On computation of the window function and the boundary correction term, the remaining
solution procedure is very similar to the one dimensional procedure. The numerical
integration of the stiffness and force terms is enabled by Gauss-Legendre quadrature in
two dimensions.
An illustration of two dimensional shape functions using Lagrangian kernel is presented
in the following figures. The shape functions are plotted over a unit square domain. The
figures depict plots of shape functions of an extreme node and a center node.

49

Figure 5.1 Shape function of node 25 using Lagrangian kernel

Figure 5.2 Shape function of node 13 using Lagrangian kernel

50

5.4

Solution Procedure

The solution procedure of a model two dimensional problem using the reproducing kernel
particle method and a Lagrangian kernel is detailed in this section. The procedure is also
illustrated in the flowchart depicted in Figure (5.3).
The domain is represented by nodes, whose co-ordinates are input, along with
information about the boundary and interior nodes. Once the boundary nodes are
established and identified as Dirichlet and Neumann nodes; the dilation parameter,
support radius and other relevant data are specified.
Integration points (Gauss-Legendre quadrature points) are set throughout the domain and
their weights are specified. Each of these integration points falls within the support of one
or more nodes. Our next step is to compute the shape functions centered at every node
with respect to each of the gauss points within its support to construct the stiffness
matrix. To do this, the Euclidean distance between each node and each Gauss point is
calculated to determine the nodes under whose supports each of the Gaussian integration
points lie. The Euclidean distance between the integration point and the node is computed
by,
xnorm =

(x

x j ) + (y g y j )
2

(5.18)

The boundary correction term and the kernel function are computed for the shape
functions to be determined. The procedure of integration is self-explained in the
flowchart. The stiffness matrix and force vector are assembled. The essential boundary
conditions are then applied as in finite element methods. This is done by setting the
stiffness terms corresponding the essential boundary nodes equal to zero, except those on

51

the diagonal, which are set to 1. Correspondingly, the force terms of the essential
boundary nodes are set to the value of the dependent variable specified on the respective
essential boundary node. Now that the partial differential equation representing our
problem has been reduced to a set of algebraic equations, the solution can be obtained by
solving these equations. This solution gives us the nodal values of the dependent variable.
The approximate solution at every node is then obtained by a summation of the products
of the shape function of the node with respect to every other node and the corresponding
nodal values. This gives:
u h (x ) =

N (x )u
j =1

(5.19)

52

Input a, r, number of
nodes and nodal
coordinates

Establish Dirichlet
and Neumann
boundary nodes

Set the lumped mass


(trapezoidal weights)
of each node

Define 2D Gauss
integration points
over the domain

Loop over each gauss


point,g

Establish all the nodes whose


supports enclose the Gauss point,g,
by calculating Euclidean norm

Loop over each node,


i, associated with
gauss point, g

Formulate window
function and its
derivatives

B
A

Calculate moment
matrix, basis and the
boundary correction
term and its
derivatives

53

Compute shape function,

N ( x, y ) = (a, x, y ) ( x, y ) V
and its derivatives

Formulate and assemble


force vector, f

Loop over every other


node, j

Repeat all steps as for node i. Compute


, and shape function, N and its
derivatives

Formulate and
assemble stiffness
matrix, K

End loop over node j

End loop over node i

End loop over Gausspoint,


g

54

To impose EBCs,
Set K(i,bn) andK(bn,i) = 0.0
And K(bn,bn) = 0
Where bn boundary nodes

Set f(bn) = u(bn)

Solve [K]{u} = {f}

Calculate shape
function of every node
w.r.t. every other node

Arrive at solution by,

u h ( x, y ) =

N ( x, y ) u
i =1

Figure 5.3 Flowchart depicting 2D RKPM solution procedure

55

CHAPTER 6

NUMERICAL EXAMPLES

6.1

One Dimensional Examples

In order to illustrate the validity of employing the Lagrange kernel in the


reproducing kernel particle method, one dimensional examples were solved using the
Lagrange kernel, the Gauss kernel and the cubic spline kernel and all three results have
been compared to the analytical solution.
6.1.1 Axial Deformation of a Bar

A one dimensional bar with uniform body force, q, is considered. The bar is fixed
on one end and is free on the other, i.e., we have an essential boundary condition on end
and a natural boundary condition on the other.

Figure 6.11D bar subjected to body force

56

The governing differential equation for the axial deformation of this bar is given by,
2u
EA 2 = q
x

,0 xl

(6.1)

with the following boundary conditions,


u (0 ) = 0

u (l ) = 0
x

(6.2)

The weak form of the above governing differential equation is,


u
u
u
EA
dx EA
= q dx
+ EA

x x
x x = l
x x = 0 0
0
l

(6.3)

On application of the boundary conditions, the weak form reduces to:


u
EA
dx = q dx
x x
0
0
l

(6.4)

From the above weak form, we can discern the formula for the terms of the stiffness
matrix and force vector as,

K (i, j ) = EA
f (i ) = i q dx

i j
dx
x x

(6.5)

Thus, the above formulae are applied to a steel bar of length 3 m. The Youngs modulus
of the bar is E = 3*10^7 psi. The bar is of uniform cross-sectional area of 1m2.
The following table gives the dilation parameters and radii of influence (support), used
for each of three kernels:

57

Kernel

Dilation parameter

Radius of support

Gauss

0.133

r = delx

Cubic Spline

1.1

r = 3.0*delx*a

Lagrange

1.0

r = delx

Table (6.1) Dilation parameter and radius of support of kernels


Note that for the Gauss kernel, the refinement parameter is actually its standard deviation,
given by the formula [37],
a = 2 j x

(6.6)

The analytical solution for this problem is,


1 (x 3 27 x )

= u (x )
2
EA
Figures (6.2), (6.3) and (6.4) represent the solutions obtained using the reproducing
kernel particle method by employing the Gauss kernel, cubic spline kernel and the
Lagrange kernel respectively.

58

Figure 6.2 Axial deformation of a bar - RKPM (Gauss kernel) vs analytical solution

Figure 6.3 Axial deformation of a bar - RKPM (cubic spline kernel) vs analytical
solution

59

Figure 6.4 Axial deformation of a bar - RKPM (Lagrange kernel) vs analytical


solution

6.1.2 Deformation of a Dirichlet Bar

As a slight variation of the above problem, the axial deformation of a bar with both ends
fixed, or Dirichlet boundary conditions on both ends was also considered.

Figure 6.5 1D bar subject to body force with Dirichlet boundaries on both ends

The governing differential equation and the boundary conditions are,


EA

2u
+ 10 x = 0
x 2

(0 x 10)

u (0) = 0, u (10) = 0

(6.7)

60

The analytical solution for this problem,

( 5 x
u (x ) =

+ 500 x )
3 EA
3

The plots of the RKPM solution versus the analytical solution are given below:

Figure 6.6 Axial deformation of a 1D Dirichlet bar: RKPM solution vs Analytical


solution

6.2

Numerical Example Two Dimensional Poissons equation

As a model problem, we first consider the steady state heat transfer over a two
dimensional plate. The temperature distribution of the plate is determined by using the
reproducing kernel particle method. The governing differential equation for this problem
is the Poissons equation:
2u 2u
k 2 + 2 = q
y
x

(6.7)

61

where, k thermal conductivity


q - internal heat source
u - temperature (dependent variable)
This particular example deals with a unit square domain.
=

{(x, y ): 0 x 1, 0 y 1}

The thermal conductivity of the plate, k = 1. An internal heat source of 50 is supplied.


The boundary conditions are set as follows:
Essential boundary conditions:

u ( x,0) = 0

u ( x,1) = 0

u (0, y ) = 0
x
Natural boundary conditions:

u (1, y ) = 0
x

(6.8)

(6.9)

The weak form of the aforementioned Poissons equation is given by,


u
u u
u
dxdy + k n x + n y ds =
k
+
x x y y
y
x

q dxdy

(6.10)

After application of the boundary conditions the stiffness and force terms are,
N N j N i N j
dxdy
K (i, j ) = k i
+
y y
x x
f (i ) =

(6.11)

q dxdy

The Lagrange kernel is used as the window function with a dilation parameter, a =1.0
and a support radius, r = 0.4. A total of 25 nodes were used to represent the domain. The
whole domain was covered with 256 Gauss- Legendre quadrature points to perform the
integration. For easier comprehension, this could be viewed as spreading the Gauss points
uniformly over the domain in integration zones, with each zone consisting of a 4x4 Gauss

62

quadrature distribution and bordered by four nodes. It should be kept in mind that the
above statement implies that there is a uniform distribution of integration points over the
domain, and not division of the domain into integration zones. This is not necessary in
meshfree methods and is not performed in the actual computation. A linear basis is used
to compute the boundary correction term.
The temperature distribution of over the plate is illustrated in Figure (6.7)

Figure 6.7 Temperature distribution over a 2D plate by Poissons equation, solved


using RKPM with 25 nodes

If the nodal density is increased by adding more nodes to the domain, the solution
improves in accuracy. Figure (6.8) indicates the temperature distribution for the same
problem with 36 nodes representing the domain.

63

Figure 6.8 Temperature distribution over a 2D plate by Poissons equation, solved


using RKPM with 36 nodes

Thus, it can be concluded that increasing the number of nodes leads to more
accurate solutions. Implementation of the above problem using RKPM with a Lagrange
kernel is achieved in much lesser calculation steps than with other kernels that do not
satisfy the Kronecker delta property, such as the Gauss kernel. If the Gauss kernel was
used, the boundary conditions have to be employed through a very tedious process. The
next example portrays more detailed analysis results.

6.3

Numerical Example 2D Laplaces equation

A more thorough representation of results for a model two dimensional Dirichlet


problem is presented in this section. Solution to the 2D Laplace equation is sought by
applying the reproducing kernel particle method using the Lagrange kernel. The
boundary value problem definition is given below:
2u 2u
+ 2 = 0,
x 2
y

0 x 1, 0 y 1

64

The boundary conditions are,


u ( x,0 ) = sin (x )
u ( x,1) = 0
u (0, x ) = 0
u (1, x ) = 0
The analytical solution [39] to this problem is given by,
u = [cosh (y ) coth sinh (y )]sin (x )

The weak form for the Laplace equation is given by,


u
u u
u
dxdy + k n x + n y ds = 0
k
+
x x y y
y
x

The weak form is reduced by considering the boundary conditions and the terms for
stiffness matrix and force vector are derived, similar to the previous examples. The
RKPM shape functions are computed using the Lagrange kernel and the solution to u is
obtained. The problem has been implemented for various nodal densities. A constant
dilation parameter of a = 1.0 has been used for all the cases. Figures (6.9), (6.10), (6.11)
and (6.12) show the plot of the RKPM solution using Lagrange kernel against the
analytical solution.

65

Figure 6.9 RKPM vs Analytical solutions for 25 nodes

Figure 6.10 RKPM vs analytical solution for 100 nodes

66

Figure 6.11 RKPM vs analytical solution for 625 nodes

Figure 6.12 RKPM vs analytical solution for 2500 nodes

The L2 error norm of the results is plotted against the number of nodes in each
direction in Figure (6.13). (It is to be noted that in this example, we consider equal and

67

uniform nodal distribution in the x- and y- directions). The slope of the L2 norm curve is
around 1.11 for RKPM using Lagrange kernel. The L2 error norm of RKPM using the
Lagrange kernel is compared with traditional RKPM using a cubic spline kernel and a
corrected collocation technique [39].

Figure 6.13 Logarithmic plot of L2 norm of error vs nodal density for RKPM using the
Lagrange kernel, cubic spline kernel and a corrected collocation method

6.4

Results Discussion

The results obtained to the one and two dimensional problems using the
reproducing kernel particle method employing the Lagrange kernel appear to be
promising. As shown in the examples involving the axial deformation of a bar, the
solution that resulted from using the Lagrange kernel matches the analytical solution,
equally exact as those that resulted from the Gauss kernel or the cubic spline kernel. This
suggests that the Lagrange kernel may work as effectively as the conventionally popular
kernels used in RKPM. The fact that the application of simple divided difference kernels

68

also works for two dimensional problems, supports the effectiveness of this concept.
Also, it is observed that the solution accuracy is directly proportional to the number of
nodes that represents the domain, as in any numerical method.
The results obtained for the 2D problem with Laplaces equation, are comparable
to those obtained by Wagner et al. [39], who have demonstrated the same problem by
using a corrected collocation method. This also suggests that the Lagrange kernel gives
reasonable results when applied with RKPM. It is to be noted that the Lagrange kernel
used in this particular example was a linear kernel. This lower order kernel contributes to
the shallower slope seen in the error norm for the Lagrange kernel when compared to the
error norm for the higher order cubic spline kernel used in Wagner et al [39]. If a higher
order Lagrange kernel was used in place of the linear kernel, the slope of the L2 error
norm would increase, i.e., a higher rate of convergence would be obtained.
Extension of the formulation to problems involving multiple dimensions is not
very complicated and can be done by using multi-dimensional basis functions. The
complexity of extension is further simplified by the ease of implementation of essential
boundary conditions, because there arises no more a necessity to employ a special
technique in an additional dimension to enforce the boundary condition.
Thus, the major difference between RKPM using Lagrange kernel and traditional
FEM lie in the development of the shape functions and their derivatives. Since the
solution procedure beyond this step is the same as in other Galerkin methods, RKPM
shows promise as a technique to solve problems of higher complexity.

69

CHAPTER 7

CONCLUSION

Research on meshfree methods is progressing successfully to help overcome the


shortcomings faced with traditional mesh-based methods. The scope of their application
in various fields is increasing by the day. It is only a matter of time before meshfree
methods becoming an indispensable component of conventional computational
mechanics.
The reproducing kernel particle method has been studied in detail as part of this
research. Its mathematical framework was explored to help gain an insight into its
formulation, implementation, and limitations. One of the major problems with RKPM is
the difficulty faced in applying the essential boundary conditions. Since the traditional
RKPM shape functions do not satisfy the Kronecker delta property, special methods such
as the Nitsches method and the penalty method are adopted to impose the essential
boundary conditions. Yet, these methods have their shortcomings due to complexity,
computational burden and difficulties arise when solving dynamic problems.
This research sought a solution to this limitation by examining the kernels used to
build the RKPM shape function. The importance of kernel smoothness over the partition
of unity property was deliberated and it was concluded to be of minimal significance.

70

Hence, a simple divided difference kernel, the Lagrange kernel was adopted to replace
the conventionally popular Gaussian and cubic spline kernels. The implementation of
RKPM with a Lagrange kernel served to provide easy imposition of the essential
boundary conditions and efficient computation. Numerical examples were used to
illustrate the capacity of the Lagrange kernel to produce accurate results, without the
difficulties of employing additional techniques to deal with boundary conditions.
These results illustrate a simple solution to an important issue in RKPM. When
applied to complex real-life problems, the reproducing kernel particle method with the
Lagrange kernel promises to yield solutions with a considerable minimization in
computational burden that is experienced with prevailing kernel practice.

71

REFERENCES

[1] C.A. Felippa, A historical Outline of Matrix Structural Analysis: A Play in Three
Acts, Report No. CU-CAS-00-13, June 2000.
[2] J.N. Reddy, An Introduction to the Finite Element Method, Third Edition, Mc-Graw
Hill 2006, ISBN 0-07-246685-5
[3] C.A.Duarte, J.T.Oden, Hp Clouds A Meshless Method to Solve Boundary Value
Problems, TICAM Report 95-05, The University of Texas at Austin, 1995.
[4] Belytschko et al, Meshless methods: An overview and recent developments,
Computer Methods in Applied Mechanics and Engineering, Vol. 139,Issues 1-4
(1996), 3-47.
[5] W.K.Liu, S.Li, Meshfree and particle methods and their applications, Appl Mech
Rev vol 55, no 1, January 2002
[6] S. Li, W.K.Liu, Meshfree Particle Methods, Springer 2004, ISBN 3540222561,
9783540222569
[7] J.J. Monaghan, Why Particle Methods Work, Society for industrial and applied
mechanics, 1982, Pgs 422-433.
[8] J.S. Chen, Preface, Computer Methods in Applied Mechanics and Engineering, Vol.
193, Issues 12-14 (2004), Pgs iii-vi.
[9] M.A. Puso, JS Chen, E. Zywicz, W. Elmer, Meshfree and finite element nodal
integration methods, International Journal for Numerical Methods in Engineering,
Vol. 74, No. 3 (2008), 416-446.

72

[10]

W.K.Liu et al, Multiresolution Reproducing Kernel Particle Method for

Computational Fluid Dynamics, International Journal for Numerical Methods in


Fluids, Vol 24, Issue 12, Pgs 1391 1415, 1997
[11]

T.P.Fries, H.G. Matthies, Classification and Overview of Meshfree Methods,

July 2004
[12]

Gui-Rong Liu, Meshfree methods: Moving Beyond the Finite Element Method,

CRC Press, 2002, ISBN 0849312388


[13]

JS Chen et al, Reproducing Kernel Particle Method for Large Deformation

Analysis of non-linear structures, Computer Methods in Applied mechanics and


Engineering, 139 (1996), 195-227.
[14]

R.A. Gingold, J.J.Monaghan, Kernel Estimates as a Basis for General Particle

Methods in Hydrodynamics, Journal of Computational Physics, 46, 429-453, 1982.


[15]

J.J. Monaghan, Smoothed Particle Hydrodynamics, Annual review of

Astronomy and Astrophysics, 1992, 30:543-574


[16]

L.D.Libersky, P.W.Randles, Smoothed Particle Hydrodynamics: Some recent

improvements and applications, Computer Methods in Applied Mechanics and


Engineering, 139 (1996), 375-408
[17]

J.J.Monaghan, An Introduction to SPH, Computer Physics Communications,

48(1988), 89-96.
[18]

J.S.Chen, C.Pan, C.T. Wu, Large deformation analysis of rubber based on a

reproducing kernel particle method, Computational Mechanics, Volume 19, Issue 3,


pp. 211-227 (1997).
[19]

T.Belytschko et al, A unified stability analysis of meshless particle methods,

73

International Journal for Numerical Methods in Engineering, 43, 1998, 785-819.


[20]

J. Bonet, S.Kulasegaram, Correction and Stabilization of Smooth Particle

Hydrodynamics methods with applications in metal forming simulations,


International Journal for Numerical Methods in Engineering, 2000, Vol 47, Issue 6:
1189 - 1214.
[21]

G.A. Dilts, Moving Least Squares Particle Hydrodynamics- I: Consistency and

Stability, International Journal for Numerical Methods in Engineering, 1999, Vol 44,
1115-1155.
[22]

W.K.Liu, S.Jun, Y.F. Zhang, Reproducing Kernel Particle Methods,

International Journal for Numerical Methods in Fluids, Vol. 20 (1995), 1081-1106.


[23]

P. Lancaster, K. Salkauskas, Surfaces Generated by Moving Least Squares

Methods, Mathematics of Computation, 1981, 37, 141-158.


[24]

A. Gossler, Moving Least Squares: A numerical differentiation method for

irregularly spaced calculation points, Sandia Report, SAND 2001-1669.


[25]

B. Nayroles, G. Touzot, P. Villon, Generalizing the Finite Element Method:

Diffuse approximations and diffuse elements, Computational Mechanics, 1992, Vol.


10, No. 5, 307-318.
[26]

W.K.Liu, S.Li, T. Belytschko, Moving Least Square Reproducing Kernel

Methods: (I) Methodology and Convergence, Computer Methods in Applied


Mechanics and Engineering, 143, 1997, 113-154.
[27]

T. Belytschko, Y.Y.Lu, L.Gu, Element-Free Galerkin Methods, International

Journal for Numerical Methods in Engineering, Vol. 37, 1994, 229-256.


[28]

P.Krysl, T.Belytschko, ESFLIB: A library to compute the Element Free Galerkin

74

shape functions, Computer Methods in Applied Mechanics and Engineering, Vol.


190, No.15, (2001), 2181-2205.
[29]

C.A.Duarte, A review of some meshless methods to solve partial differential

equations, Technical Report 95-06, TICAM, The University of Texas at Austin,


1995.
[30]

N.R. Aluru, G.Li, Finite Cloud Method: a true meshless technique based on a

fixed reproducing kernel approximation, International Journal for Numerical


Methods in Engineering, Vol. 50 (2001), 2373-2410.
[31]

I. Babuska, J.M.Melenk, Partition of Unity Finite Element Method: Basic Theory

and Applications, Computer Methods in Applied Mechanics and Engineering, Vol.


139 (1996), 289-314.
[32]

T.P.Fries, T.Belytschko, The Intrinsic Partition of Unity Method, Computational

Mechanics (2007), 40: 803-814.


[33]

C.A.Duarte, J.T.Oden, An h-p adaptive method using clouds, Computer

Methods in Applied Mechanics and Engineering, Vol. 139 (1996), 237-262.


[34]

H.Lu, JS Chen, Adaptive Galerkin Particle Method, Book Chapter in Meshfree

Methods for Partial Differential Equations, Springer 2002.


[35]

A. Graps, An Introduction to Wavelets, IEEE Computational Science and

Engineering, Vol.2 Num. 2 (1995).


[36]

Y.You, J-S Chen, H.Lu, Filters, Reproducing kernel and Adaptive Meshfree

Method, Computational Mechanics, Vol. 31 (2003), 316-326.


[37]

W.K.Liu, S.Jun, S.Li, J. Adee, T.Belytschko, Reproducing Kerenel Particle

Methods for Structural Dynamics, International Journal for Numerical Methods in

75

Engineering, Vol. 38 (1995), 1655-1679.


[38]

W.K.Liu et al, Overview and Applications of the Reproducing Kernel Particle

Methods, Archives of Computational Methods in Engineering, Vol. 3, 1 (1996), 380.


[39]

G.J. Wagner, W.K. Liu, Application of essential boundary conditions in meshfree

methods: a corrected collocation method, International Journal of Numerical


Methods in Engineering, 2000, Vol 47(8), 1367 1379.
[40]

S. Fernandez-Mendez, A. Huerta, Imposing essential boundary conditions in

meshfree methods, Computer methods in Applied Mechanics and Engineering, 2004,


Vol 193, Issue 12-14, 1257-1275.
[41]

W. K. Liu, J. Gosz, Admissible approximations for essential boundary conditions

in the reproducing kernel particle method, Computational Mechanics, 1996, 19, 120135.
[42]

W.K. Liu, Y. Chen, Wavelet and multiple scale reproducing kernel methods,

International Journal for Numerical Methods in Fluids, 1995, Vol. 21, 901-931.
[43]

Nguyen et al., Meshless methods: A review and computer implementation

aspects,

Mathematics

and

Computers

in

Simulation,

2008,

doi:

10.1016/j.matcom.2008.01.003
[44]

J. Dolbow, T. Belytschko, An Introduction to programming the meshless element

free Galerkin method, Archives of Computational Methods in Engineering, 1998,


Vol. 5, 3, 207-241.
[45]

S. Jun, W.K. Liu, T. Belytschko, Explicit reproducing kernel particle methods

for large deformation problems, International Journal for numerical Methods in

76

Engineering, 1998, Vol. 41, 137-166.


[46]

N. Sukumar, R.W. Wright, Overview and construction of meshfree basis

functions: From moving least squares to entropy approximants, International Journal


for Numerical Methods in Engineering, 2007, Vol 70(2), 181-205.
[47]

R. Schaback, H. Wendland, Kernel techniques: From machine learning to

meshless methods, Acta Numerica, 2006, 1-97.


[48]

J. Dolbow, T. Belytschko, Numerical integration of the Galerkin weak form in

meshfree methods, Computational Mechanics, 1999, 23, 219-230.

Você também pode gostar