Вы находитесь на странице: 1из 54

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/269637307

Evaluation of Accuracy and Stability of the Classical SPH Method Under


Uniaxial Compression

Article  in  Journal of Scientific Computing · December 2014


DOI: 10.1007/s10915-014-9948-4

CITATIONS READS

10 307

2 authors:

Raj Das Paul W. Cleary


University of Auckland The Commonwealth Scientific and Industrial Research Organisation
226 PUBLICATIONS   820 CITATIONS    360 PUBLICATIONS   7,123 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Waves through Nonlinear Metamaterial View project

Friction Stir Welding View project

All content following this page was uploaded by Raj Das on 13 January 2015.

The user has requested enhancement of the downloaded file.


Evaluation of accuracy and stability of the classical SPH method
under uniaxial compression
R. Das1 and P.W. Cleary2
1
Centre for Advanced Composite Materials, Department of Mechanical Engineering, University of
Auckland, Auckland 1010, New Zealand
2
CSIRO Mathematics, Informatics and Statistics, Normanby Road, Clayton, Victoria 3168, Australia
.

Abstract

The accuracy and stability of the classical formulation of the Smoothed Particle
Hydrodynamics (SPH) method for modelling compression of elastic solids is studied to assess
its suitability for predicting solid deformation. SPH has natural advantages for simulating
problems involving compression of deformable solids arising from its ability to handle large
deformation without re-meshing, complex free surface behaviour and tracking of multiple
material interfaces. The ‘classical SPH method’, as originally proposed by Monaghan and co-
workers [1, 2], has become broadly established as a robust method in different areas, Deleted: for

especially involving fluid flows. However, limited attention has been paid to understanding of
its numerical performance for elastic deformation problems. To address this, we evaluate the
classical SPH method to explore its stability, accuracy and convergence and the effect of
numerical parameters on elastic solutions using a generic uniaxial stress test. Short term
transient and long term uniform state SPH solutions agree well with those from the Finite
Element Method (FEM). The SPH elastic deformation solution showed good convergence
with increasing particle resolution. The tensile instability stabilisation method was found to
have little impact on the solution, except for higher values of the correction factor which then
produce small amplitude benign artificial banded stress patterns. The use of artificial viscosity
is able to eliminate the instability and improve the accuracy of the solutions. Overall, the
classical SPH method appears to be robust and suitable for accurate modelling of elastic
solids under compression.
Keywords: Smoothed particle hydrodynamics; Elasticity; Uniaxial compression; Stress waves;
Convergence; Stability.


Corresponding author. address: Department of Mechanical Engineering, University of Auckland, Auckland
1010, New Zealand a. Tel: +64 9 923 5094, Fax: +64 9 373 7479, E-mail: r.das@auckland.ac.nz.
1 Introduction

Smoothed Particle Hydrodynamics (SPH) is a particle based or meshless numerical method


that is used to solve dynamic PDE systems [1, 2]. It was first proposed by Gingold and
Monaghan [3] and Lucy [4] for use in solving astrophysical gas dynamics problems. The
computational domain is discretised to form SPH particles and the field variables are
approximated at each particle location using a suitable interpolation function (known as
smoothing kernel). Using a smoothing process for field variables and their derivatives,
systems of partial differential equations can be converted to SPH forms. For details of the
SPH background theory and its applications, refer to [1, 5-10].

While SPH has traditionally been applied to the modelling of fluid flows [11, 12], in recent
years there has been a growing interest in its application to a wide variety of solid mechanics
problems [13-17]. SPH potentially provides a unified approach for the modelling of coupled
fluid-solid interaction problems, such as rock fracture under hydraulic pressure and fluid flow
through solid porous media. SPH can potentially provide a numerical framework that is
highly suited to solving multi-physics problems in geomechanics [18]. The method also
shows promise for simulation of manufacturing processes involving solid deformation such as
forging, extrusion and drawing [16, 19].

A primary feature of SPH is that as a particle based technique it does not require an
underlying grid structure to represent materials. This avoids difficulties associated with
traditional mesh-based methods (FEM, FVM and BEM), such as maintenance of the integrity
and quality of the mesh under large deformation. The grid-free nature of SPH makes it well
suited to modelling processes which involve large deformations and discontinuities, such as
fracture [20], fragmentation [21], metal forming [16, 19, 22], and nano-machining [23]. SPH Deleted: and

therefore offers several potentially important advantages for simulating geomechanics and
solid deformation processes.

The simplest and most representative compressive stress state is that of uniaxial compression.
So in this paper we use a generic uniaxial compression test as the model configuration with
which to study the numerical performance of SPH. This test is the most widely used
mechanical testing process for characterising solid material behaviour. It is a simple and
versatile method for determining material properties in many material applications [24] and is
representative of the stress conditions found in many complex engineering and geomechanics Deleted: more

problems. In this test, a standard specimen is gripped between the jaws of a tensile testing
machine. One end of the specimen is pushed by a moving piston and the other end is held by a
fixed jaw. The compression in the specimen is measured by extensometers and/or strain
gauges. We will consider different specimen geometries and material properties

We will first introduce the classical SPH formulation of the elasto-dynamic equations and
then detail the test configuration. We will then analyse the stress wave solutions and longer
time behaviour of the SPH solution and then compare these to a high resolution Finite
Element Method (FEM) solution in order to assess the solution accuracy. Convergence with
increasing resolution will then be evaluated. Finally, the effect of the tensile instability
correction and the artificial viscosity on the solution will be evaluated. These will provide a
detailed understanding of the performance of SPH under uniaxial compression. Deleted: clear

2 SPH Method Variants and Stabilisation Approaches

We will first review the different variants of SPH that are commonly used by research
communities, and then discuss the studies undertaken so far in relation to the accuracy,
convergence and stability aspects of the different SPH formulations.

2.1 Different variants of the SPH method

The SPH method developed by Gingold and Monaghan [3] and its related variants for
different applications, which essentially use the same basic formulation (of non-normalised
kernel), are termed as ‘classical SPH’ here. This ‘classical SPH’ formulation has been
modified by many researchers in an attempt to improve the accuracy and stability of the
method. The commonly used variants of SPH can be broadly classified into two basic
categories: (a) Corrected SPH (CSPH) and (b) Stress point SPH.

The concept of CSPH was first proposed by Liu et. al. [25] in connection with the
development of Re-producing Kernel Particle Method (RKPM). The ideas of kernel and
gradient corrections (renormalisations) were borrowed from RKPM and subsequently applied
to SPH [26, 27]. Considerable research has since then been undertaken in applying CSPH for
a wide range of solid problems, notably by Bonet and Kulasegaram [10, 28, 29] and others
[30]. In this technique, a combination of the following approaches are used to improve the
accuracy of classical SPH: (a) Zero order kernel correction, (b) First order kernel correction,
(c) Zero order gradient correction, (d) First order gradient correction. The choice of the
correction(s) depends on specific problems and affects the computational efficiency.
The accuracy and stability of CSPH have not been thoroughly examined for practical and
complex (3D) problems. The examples used in the literature to demonstrate the effectiveness
of CSPH are mostly simple 1D and 2D problems [10, 26, 28, 30], where tensile instability
was removed by using the CSPH corrections. The use of CSPH increases the computational
time because of the need to solve for the correction matrix at each interpolation location for
every time step. Also, CSPH has limited utilities for many applications, especially those
involving high deformation and disintegration, such as material forming and
fracture/fragmentation. While using CSPH for these problems, specific configurations of SPH
particles near or at the boundaries (e.g. three particles in a straight line) may lead to a singular
kernel or gradient correction matrix, and consequent failure of the respective correction. In
such cases, an attempt to overcome the singularity (e.g. by introducing tolerance factors) can
contribute to additional inaccuracies in the solutions.

The Stress point SPH method, proposed initially by Dyka and Ingel [31], has been used by
many researchers to improve the accuracy and stability of SPH [32-34]. This involves the use
of a staggered particle arrangement similar to that used in the finite volume method. The
analysis domain consists of two sets of particles: master particles and slave particles. The
master particles are the SPH particles (also called velocity points) and are used for evaluation
of kinematic variables (displacement, velocity, etc.). The slave particles are called stress
points. The pressure, internal forces and stresses are calculated at these locations. This method
has been used and shown to work for limited 1D and 2D problems (in simple wave
propagation and impact) [31, 32]. In practice, it is difficult to create particles with correct
positioning of the velocity and stress points for complex 2D shapes and particularly for
arbitrary 3D bodies, especially when using an automated particle/mesh generation program.
Also, maintaining the relative positions of master and slave particles in high deformation
problems is often difficult. This limits the applicability of this technique for realistic
problems, as acknowledged in [33].

So neither CSPH nor Stress point SPH has yet to become established as robust methods for
realistic solid deformation problems, particularly for complex geometries in three dimensions.
As a result, despite using the CSPH or Stress point SPH formulation, a number of additional
measures are sometimes required to ensure an accurate and stable solution. For example,
Vidal et al. [30] found that CSPH exhibits the ‘sawtooth’ type instability for high strain-rate
problems, and they applied a new stabilisation technique called ‘Hessian difference
stabilisation’, where a Laplacian filter was added to the gradient approximation to remove the
instability. For Stress point SPH, Belytschko and Xiao [35, 36] concluded that the location
and the number of stress points (slave nodes) play an important role in ensuring a stable and
accurate scheme for stress point integrations, even in 2D domains. It could not be ruled out
that some specific relative arrangements of the master and slave nodes can even introduce
additional instabilities.

As a result of various difficulties posed by CSPH and Stress point SPH, the classical SPH Deleted: stress point SPH and

method, originally proposed by Monaghan [1, 2], is regaining popularity and different
variants of it are being developed. The artificial viscosity in the conventional SPH method
was modified by Shaw and co-workers [37-40] to remove spurious high frequency
oscillations in solutions. In this variant of SPH, the effect of the artificial viscosity term on the
change in acceleration (with and without artificial viscosity) of the SPH particles was
evaluated. An acceleration correction algorithm was proposed for use in the momentum
equation so that the artificial viscosity term is derived by minimising the spatial variation of
the high oscillations in the acceleration [40]. This was found to reduce the large energy
dissipation often caused by artificial viscosity, while retaining its advantages of stabilising the
computations. This acceleration correction term was further optimised for hypervelocity
impact problems in [37, 38], where high frequency spurious oscillations are often observed,
particularly for shocks and sharp discontinuities. A novel Galerkin SPH formulation was
proposed in [41, 42] for large deformation structural mechanics applications. This method
employs a moving-least-squares approximation along with stabilised nodal integration.
However, the accuracy and stability of this method were not examined in those paper. Deleted: i

2.2 Accuracy and convergence for solid deformation applications

In most of the SPH literature concerning solid deformation problems, the accuracy of CSPH
and Stress point SPH solutions have been assessed using mostly simple 1D and (in few cases) Deleted: s

2D test problems. The typical problems analysed are 1D stress wave solutions in bars/rods in
comparison to FE solutions, see [25, 26, 29-34]. For classical SPH, there has been little
detailed quantitative evaluation of accuracy either against analytical or other numerical
solutions, such as FEM. The accuracy studies, conducted to date, are mostly limited to
qualitative comparisons (e.g. in [43]). In this paper, we will study uniaxial compression in a
rectangular plate to allow for the reflection and lateral superposition of the waves, which play
an important role in many solid deformation and dynamic fracture problems. We will evaluate
the accuracy of the short-term (early stage) 2D wave propagation solutions and long-term
uniform state solutions, predicted by classical SPH, against FE solutions.
Little attention has been paid to testing the convergence of different variants of SPH till date.
Vignjevic et al. [34] studied convergence of Stress point SPH using the problem of 1D Deleted: s

symmetrical impact of two aluminium bars. A particle resolution study with only two
different resolutions (202 and 4202 particles) was performed, and the value of the kinetic
energy was shown to approach the equivalent finite element (using LS-DYNA) results with
refinement. Apart from this, the convergence of neither CSPH nor classical SPH has been
examined rigorously for elastic solid problems. In this paper, we will therefore perform a
convergence study of classical SPH with varying resolutions by comparing against both
analytical and finite element solutions.

2.3 Tensile instability in SPH

The origin of ‘so called’ tensile instability has received much attention in SPH. It has been
marked as a serious demerit of the SPH method, and can adversely affect the solution of solid
deformation problems by producing non-physical clustering of particles in the zone of
material under tension, leading to erroneous numerical fracture. The tensile instability in SPH
was first studied by Swegle et al. [44]. Afterwards, many studies have been undertaken to
develop alternative SPH formulations and stabilisation techniques to mitigate this instability.
There are a number of established techniques for moderating the tensile instability in SPH:

 Principal stress modification in classical SPH [15],

 SPH kernel and/or gradient renormalisation (Corrected SPH) [10, 25-30],

 Use of additional set of integration points or stress points, similar to a staggered grid
(Stress point SPH) [31, 32], [33, 34],

 Use of higher order kernels and alternating kernels [31, 44, 45],

 Use of stability filters [30, 35, 36].

The first method, termed as the ‘principal stress correction’, was initially proposed by
Monaghan [46] for fluid problems and then extended to elastic solid problems by Gray et al
[15]. This technique has been found to be effective by Melean et al. [47] for droplet formation
in viscous liquids and by Cleary et al. [16] for material forming applications. This method
involves explicitly removing a fraction of the tensile force in the principle frame of each
particle to control the instability. This method has been shown to be robust for both solid [13-
16, 43, 48] and fluid [1, 11, 12, 49-54] problems, and it does not affect computational
efficiency, unlike some of the alternative methods. For example, both the CSPH and Stress
point SPH methods increase computational time because of the requirement of correction Deleted: P
matrix inversion and integration at additional (stress) points, respectively. This makes the
principle stress correction method attractive for large scale engineering and geomechanical
problems where computational expense is a major issue.

From the above review, it is noted that only limited assessment of the numerical performance Deleted: clear

of the various SPH formulations has been undertaken to date for elastic solid deformation. So
far there is no established SPH formulation that can be applied to a broad range of
applications. As a result, the classical SPH still continues to be a popular and widely used
method for a variety of problems because of its robustness, simplicity and computational
efficiency. However, despite its widespread use, detailed evaluation of the accuracy and
stability of the classical SPH for solid mechanics problems remains inadequate.

The aim of the present paper is not to develop a new formulation of SPH or new methods for
suppressing or reducing the tensile instability. Rather, we seek to evaluate the ‘classical SPH’
method. The primary objective is to thoroughly assess the accuracy and convergence of
‘classical SPH’ for elastic solid problems. Accuracy will be determined by comparison with
the mesh-based finite element method. We therefore use the principal stress correction
technique as proposed by Gray et al. [15], and investigate the effect of the choice of tensile
correction coefficient on the solution quality for elastic solid deformation problems, which
again has not been addressed before. This will enable us to identify suitable ranges of the
correction for ensuring accuracy and stability of solid deformation solutions.

First, we will evaluate the numerical performance of the classical SPH method for predicting
transient stress wave propagation and for long term stress behaviour in elastic solids. This will
then provide a basis for expanded use of the method for solving engineering problems
involving solid deformation and potentially fracture under compressive loads. Accurate
transient behaviour modelling is important because fluctuating stress can trigger generation
and propagation of localised damage or flaws, which then play a critical role in initiating
fracture failure. It is also important for systems with rapidly changing stress states, such as
those found in metal forging.

3 SPH Formulation of Elastodynamics

SPH has been developed for and applied to modelling solid deformation problems in a range
of applications [13-16, 43, 48]. It is based on the use of local interpolations from surrounding
discrete particles to construct continuous field approximations. This is the basis of the spatial
discretisation of the governing equations. The basis of SPH is the following identity to
express any spatial function A(r), which can be exactly written as:

A  r    A  r     r  r   d r  (1)

where A(r´) is the value of A at r = r´, and δ(r - r´) is the Dirac delta function. The volume
integral is over either a 2D or 3D material volume, depending on the dimensionality of the
problem domain. In the SPH approximation, the volume integral is approximated as a sum of
values of A over neighbouring particles b (Ab). The elemental volume dΩ r' is given by mb/ρb,
where mb and ρb are the mass and the density of particle b.
The Dirac delta function is not differentiable and cannot thus represent an approximated
distribution of continuous field (state) variables, and so instead a continuous function (weight
function or kernel function) is used in the approximation of continuous field variables in the
SPH method. Hence, the interpolated value of a function A at any position r can be
expressed using SPH smoothing (replacing the Dirac delta function in Equation (1) by a
weight function) as [1]:
Ab
Ar    mb W r  rb , h  (2)
b b
where the sum is over all particles b within a radius 2h of r. Here W(r, h) is a C2 spline based
interpolation or smoothing kernel, that approximates the shape of a Gaussian function, but has
compact support with radius 2h. The gradient of function A is given by differentiating the
interpolation equation (2) to give:
Ab
Ar    mb W r  rb , h  (3)
b b
The gradient of the kernel function as formulated in equation (3) is based on its central
function property. Using these interpolation formulae and suitable Taylor series expansions
for second order derivatives, one is able to convert parabolic partial differential equations into
ordinary differential equations for the motion of the particles and the rates of change of their
properties.

3.1 Continuity equation

The continuity equation for an elastoplastic solid is given as:


d
    v (4)
dt
where  is the density and v is the velocity. For use in the SPH form we denote the position
vector from particle b to particle a by rab  ra  rb , and let Wab  W rab , h  be the
interpolation kernel with smoothing length h evaluated for the distance rab . The SPH

discretisation of the continuity equation from [1] is then:


d a
  mb  v a  v b    aWab (5)
dt b

where a is the density of particle a with velocity va, and mb is the mass of particle b with
velocity vb. This form of the continuity equation is Galilean invariant (since the positions and
velocities appear only as differences), has good numerical conservation properties, and is not
affected by density discontinuities or free surfaces.

3.2 Momentum equation and stress evolution equation

The momentum equation used for predicting elastic deformation of solids is:
dv 1
 σ  g (6)
dt 
where v is the velocity, g denotes the body force (gravity), and  is the stress tensor. The
stress tensor can be written as:
σ  P I  S (7)
where P is the pressure and S is the deviatoric stress tensor. Assuming Hooke’s law with
shear modulus G, the evolution equation for the deviatoric stress S is calculated using the
Jaumann rate equation [15] as:
dS ij  1 
 2G   ij   ij  ij   S ik  kj   ik S kj (8)
dt  3 
where the components of the strain tensor ε are given by:

1  v i v j 
 ij     (9)
 x x i
j
2 
and the rotation tensor that accounts for the large rotational effect is given by:

1  v i v j 

Ω ij   (10)
2  x j x i 
The SPH discretisation of the momentum equation (6) is:
dv a  σa 
σb
dt
  m  
b  Π ab I    aWab  g
2

 b2
(11)
b  a 
where a and b are the stress tensors of particles a and b, ab is an artificial viscous stress
term that produces a shear and bulk viscosity, and g is the gravity. For particle a, the SPH
equations for the components of strain rate ε a and the rotation tensor Ωa are:
. ij 1 mb i i Wab Wab
a  
2 b b
[(vb - v a )
x aj
 (vbj - v aj )
x ai
] (12)
1 m W W
 b [(vbi - vai ) x abj  (vbj - vaj ) x abi ]
Ωaij 
2 b b
(13)
a a
An Improved Euler explicit integration scheme was used [2], with the time step Δts
determined by the Courant condition:
t s  min0.5h c a  (14) Formatted: Centered
a
where ca is the sound of speed of particle a, and it is calculated from the material bulk
modulus and reference density by ca  K /  0 .

3.3 Equation of state

We use an equation of state where the elastic pressure is proportional to the change in density:
Pa  ca2   a  0  (15)
where 0 is the reference density and a is the current density. Deleted: ,
Deleted: and c is the speed of sound in
We use the real speed of sound for the material in solid deformation analysis using SPH so as the solid material

to accurately resolve the short-term stress wave propagation. This is in contrast to fluid flow Formatted: Subscript

problems where the sound speed is treated as a numerical parameter and chosen to increase
computational speed. In fact one approach to solve incompressible fluid flow problems using
the SPH methodology is assuming the fluid to be ‘quasi-compressible’ (weekly compressible)
[49]. The incompressibility is attained by setting the fluid sound speed much higher (up to 10
times) than the characteristic (~largest) velocity scales in the flow and thus results in a low
Mach number (~0.1). This ensures that the density variations are small (~1%) leading to a
nearly incompressible flow solution. Deleted: (~1%)

3.4 Stabilisation methods

3.4.1 Tensile instability correction

SPH shows instabilities for solid deformation of materials under tension. We use the approach
proposed by Monaghan [46] and subsequently extended for solids by Gray et. al. [15], to Deleted: ,

counteract this. This involves the addition of a correcting artificial stress to the tensile
(positive) principal stress components. The form of the artificial stress is based on a linear
perturbation analysis of the governing partial differential equations. The calculation of the
artificial stress involves a scaling parameter ψ termed as the ‘tensile coefficient’, which
controls the magnitude of the artificial stress and controls the numerical instability (for further
details refer to [15]). A value of ψ = 0.1 was found to be a suitable for the present study. Deleted: value

Implementation of tensile instability correction involves modifying the momentum equation


by adding the artificial stress terms as:
dv a σ σ 
  mb  a2  b2  Π ab I  (R a  R b ) f abn    aWab  g (16)
dt b  a b 
where Ra and Rb are the artificial stress tensors of particles a and b in the world co-ordinate
frame, fab = W(rab)/W(Δp), where Δp is the SPH particle spacing, n is an index with an
optimum value of 4.0 [15] that ensures that the effect of the artificial stress is limited to the
nearest neighbouring particles.

3.4.2 Artificial viscosity

SPH also uses an artificial viscosity for stabilising the solution. For elastic solids, the
following form of the artificial viscous pressure, proposed by Monaghan and Gingold [55], is Deleted: ,

used in the SPH momentum equation (11):


   c ab    ab
2

 if ( v a  v b )  (ra  rb )  0
  ab
Π ab  (17)
0 if ( v a  v b )  (ra  rb )  0


where
h( v a  v b )  (ra  rb ) (a  b )
 ab  2
and  ab  .
ra  rb 2

Here α is the coefficient of the linear term that produces a shear and bulk viscosity, and β is
the coefficient of the quadratic term which is approximately equivalent to the Von Newmann-
Richtmyer viscosity. Here we use traditional values from [46] of α = 1.0 and β = 2.0, except
when we explore the solution sensitivity to these parameters.

3.5 Boundary conditions

In the SPH method, solid boundaries are also modelled using particles. These “boundary
particles” are assigned properties similar to the interior “solid” particles, including mass,
position, density, elastic modulus, etc. These particles are subsequently included in the SPH
interpolation using the ‘smoothing’ kernel function.
One aspect in the SPH formulation is the edge effect in implementing boundary conditions.
The SPH particles on the free edges do not have neighbours beyond the free boundary, which
leads to kernel sum deficiency when interpolating at a particle on or close to a free boundary.
This makes the implementation of general boundary conditions difficult.

In this work, we followed an approach similar to the ‘ghost particle’ method [56-60]. In this
method, two sets of particles are considered for interpolation at a given SPH particle (called
the interpolation particle, denoted by suffix ‘a’) whose domain of influence exceeds the
boundary subjected to a specified boundary condition. The ‘interior’ particles are the set of
particles (denoted by suffix ‘b’) that exist within the problem domain or at the boundary
(including the interpolation particle itself). A set of exterior particles, also called ghost
particles (denoted by suffix ‘c’), are constructed at each time step mirroring the interior
particles about the boundary and are positioned beyond the boundary to create a ‘fictitious’
extension of the boundary (or extended domain). This extended boundary is treated as a
material continuum and included in the SPH solution process.

A specified distribution of the state variables is used for the extended boundary to enable a
reflecting condition. When the state variables (e.g. density and velocity) are interpolated for a
specific SPH particle (ath particle), either near or at the boundary, the domain of influence
(neighbourhood) for kernel interpolation consists of both the interior ‘real’ (bth particle) and
exterior ‘ghost’ (cth particle) SPH particles. Thus, the sphere of the domain of influence for the
specified ath particle becomes complete, and the SPH density and momentum equations (5 and 11)
include contributions from the exterior particles lying at the extended domain.

The material properties (e.g. Young’s modulus and Poisson’s ratio) of the exterior particles
are set to be the same as those of the interpolation particle a. The evolving density of the
exterior particles are assigned to be the same as that of the interpolation particle a. The
velocity of an exterior particle is calculated by reflecting the velocity of the associated
‘mirror’ interior SPH particle about the original free boundary. The SPH continuity and
momentum equations are modified near the boundary incorporating the contributions from the
exterior particles as:

d a
  mb  v a  v b    aWab   mc  v a  v c    aWac (18)
dt b c

dv a  σ a σb   σa σc 
dt
 m b  2  2  Πab I    aWab   mc  2  2  Πac I    aWac  g
   
(19)
b  a b  c  a c 

where suffix ‘b’ represents the sum over all the interior (real) particles (including those on the
boundary) and suffix ‘c’ represents the sum over all the exterior (ghost) particles. So the
second term in equations 18 and 19 represents the boundary effects in the discretized
governing equations. We here assign ρc = ρa, and σc is calculated using the deviatoric stress
and strain rate-velocity equations (7, 8, 9, 10, 12 and 13).
The velocity of the neighboring exterior particle c is based on the velocity of the
corresponding interior ‘mirror’ particle about the boundary and the specified boundary
condition. For a stationery boundary, the velocity component(s) of the exterior particle
tangential to the boundary is set to be equal to that of the mirror particle, and the component
normal to the boundary is set to be equal in magnitude, but opposite in sense to that of the
mirror particle. For a 2D boundary (curved or straight), let ns and ts be the unit vectors
normal and tangential to the boundary at the point of intersection of the boundary and the
straight line joining the interior and the exterior particle pair (particles b and c). Then the
velocity components of particle c are:

vct  v m  t s
(20)
vcn   v m  n s

where vct and vcn are the tangential and normal components of the velocity of particle c, and
vm is the velocity of the ‘mirror’ interior particle.

For a moving boundary, let vs be the velocity of the point of intersection of the boundary and
the straight line joining the interior and the exterior particle pair (particles b and c). The above
equations are modified by adding components vs along ns and ts so as to produce the correct
specified velocity at the boundary. Deleted: correct

vct  v m  t s  v s  t s
(21)
vcn   v m  n s  v s  n s
This method of implementing the boundary conditions is accurate in that it reproduces the
correct extension of density and velocity beyond the boundary.

4 SPH Stress and Deformation Solutions for a Uniaxial Test

4.1 Test Configuration for Uniaxial Compression

A uniaxial tester is modelled using SPH (as seen in Figure 1). It uses an axially loaded
rectangular specimen of width 82 mm and height 140 mm. The specimen has the properties of
a typical sandstone (Crossley sandstone) with bulk modulus of 12.2 GPa, shear modulus of
2.67 GPa and density of 2300 kg/m3. The geometry and material of the specimen were taken
from [61]. It is held fixed at the bottom by a rigid plate and the load is applied through a
piston on the top. This simulates the requisite boundary conditions for a typical uniaxial
compression test. For the base SPH simulation, the specimen domain was discretised with
particles of resolution (spacing) 1 mm, giving a total of 12,040 particles for these two
dimensional simulations. Based on the material properties, the sound speed was 2303 m/s.
The particles were initially placed in regular square grid pattern.

In a uniaxial test, the load on the specimen is usually applied through a piston that is
hydraulically controlled to ensure that the loaded end is deformed at the required constant
rate, thus causing a uniform longitudinal strain in the specimen. This is termed ‘uniform
velocity loading’, also known as ‘constant strain rate loading’. This type of loading is
particularly suitable where the possibility of acceleration or deceleration of the loading piston
can cause sudden fluctuation in the applied pressure. The motion of the piston at a specified
constant rate prevents any accelerated motion and ensures a uniform rate of deformation of
the loaded face of the specimen. In this test the piston was moved vertically downwards at a
constant velocity of 1.5 mm/s, while the bottom end of the specimen was kept fixed by
placing it on a rigid plate. Also the relative displacement between the top surface of the
specimen and the piston was constrained.

In this study we use von Mises stress as the criterion for analysing the stress field. It combines
the normal and shear components of the deviatoric stress tensor, and is a commonly used
criterion to assess failure strength of materials. The von Mises stress σvm can be expressed in
2D in terms of principle stresses σ1 and σ2 (from [62]):

 vm   12   22   1  2 (22)

4.2 Early response

In the early stage of loading, we observe the early transient elastic stress wave propagation
within the specimen. Figure 2 shows the SPH prediction of wave propagation through the
specimen. As a consequence of the current boundary conditions, the waves initiate from the
top of the specimen and propagate downwards. A rapid variation in the stress pattern is
observed. On reaching the bottom surface, the elastic waves reflect back upwards from the
rigid bottom plate (Figure 2), and interfere with the (newly generated) incident waves from
the top due to the continued compression by the upper piston. This creates a wave pattern by
the superposition of the incident waves and the reflected waves. The ‘superposed’ waves
propagate further up through the specimen and are then reflected from the moving top piston.
These reflected waves again interact with the waves reflected from the bottom plate. The
superposition of the waves alternately reflected from the top piston and bottom plate
continues. This phenomenon leads to a complex interacting stress wave pattern in the
specimen. The variation in the amplitude of the superposed elastic waves leads to spatial
fluctuation in the von Mises stress distribution.

4.3 Uniform response

Uniform state response is reached when stress state becomes spatially uniformly distributed
throughout the specimen at any given time. The interacting waves reach a uniform state
around 3.5 ms for this problem. Figure 3 shows the uniform state stress distribution, which is
extremely even. The pattern of new and reflected waves fills the specimen and the stress
fluctuations decline, leading to a uniform state which has little change in the spatial
distribution of stress with time. However, the magnitude of the stresses at all points continues
to increase uniformly with time as the specimen is steadily compressed. At the corners of the
specimen, there are regions of (theoretical) stress singularities due to sharp geometry changes.
This induces and maintains ‘localised’ high stresses at the corners throughout the simulation,
which is a physically expected phenomenon.

4.4 Stress variation at representative points

To allow quantitative assessment of the stress field, we select three representative points in
the specimen. These are shown in Figure 4. The reasons for choosing these specific points are:

 Point A (0 mm, 70 mm) lies on the vertical and horizontal planes of symmetry (planes
v-v’ and h-h’), and this will enable analysis of the stress field at the intersection of the
two symmetry planes.

 Point B (21.5 mm, 103.5 mm) lies on neither of the symmetry planes. Stress variation
at point B will represent stress field at an arbitrary generic location with no symmetry
involved.

 Point C (21.5 mm, 70 mm) lies on a single symmetry plane (plane h-h’), and will help
monitor the stress field at a location on a plane of symmetry.

These representative locations characterise the stress variation taking into account the
problem symmetry and allow detailed comparison and quantitative assessment of errors of
SPH solutions against other numerical (FEM) and analytical solutions.

Figure 5 shows the variation of the von Mises stress at the three representative locations as the
specimen is loaded over short, medium and long time scales. Figure 5a shows the
instantaneous response of the structure for a very short period of 100 μs when the load is just
applied. The stress at point B rises first (blue line in Figure 5a). This is followed by the rise in
stress levels at points A and C as the initial stress waves first reach point B and then points A
and C. This is expected as point B is nearer the top edge. As points A and C are at same
distance from the top edge, the rise in stress levels at these locations is observed almost
simultaneously.

Figure 5b shows the stress variation over moderate time (up to 3.5 ms). After the
instantaneous sharp rise and the early oscillations in the stress levels, the subsequent transient
stress pattern (up to ~3.5 ms) exhibits declining stress fluctuations. This is because the
amplitude of the elastic waves diminishes rapidly and the response becomes increasingly
linear, reaching a uniform state. The pattern of these oscillations is complex due to the
superposition of the reflected waves and the newly generated waves. The fluctuations
gradually decline to produce a uniform state stress variation.

In a uniaxial test, the ideal uniform state response for a linear elastic structure should consist
of a uniform stress distribution at all locations (apart from the loading region where contact
mechanics plays a dominant role and affects the local stress field). Figure 5c shows the
response of the system over a long time (100 ms), from which a uniform state pattern can be
observed. Once the system has reached uniform state (t > ~3.5 ms), the stresses at points A, B
and C are found to be the same at any given time and increase linearly with load (and
therefore time) as expected. This demonstrates that the SPH method is producing the expected
uniform (spatial) stress distribution at uniform state and linear elastic structural behaviour.

5 Comparison of SPH and FEM Stress Solutions

The specimen geometry used for the previous uniaxial test simulation had a finite width and
height, and as such it has no exact elasto-dynamic solution. The FEM was therefore used to
model the uniaxial test to provide an alternative numerical solution with which to compare the
SPH solutions. The following measures were taken to ensure that the finite element analysis
provides an accurate and stable solution.

 For FE modelling of 2D elasto-dynamic problems, plane stress elements are


commonly used. They are known to be most accurate and stable for transient dynamics
problems [63], whereas the plane strain elements are usually more error prone. So in
this study the FE model consists of plane stress elements.

 Higher order shapes functions interpolate the stress field more accurately within an
element. This is particularly relevant to modelling stress wave propagation, in which
the internal stresses may vary significantly. Therefore, second order (eight-node)
quadrilateral elements were used (instead of linear elements) to improve the accuracy
of the FE solutions.

 We used a structured quadrilateral mesh, rather than an unstructured triangular mesh,


to further improve the accuracy of the FE solutions.

 The FE solution was obtained using an implicit dynamic analysis, which provides
more stability than an explicit method.

The mesh resolution with FEM is chosen to match the 1 mm particle resolution of the SPH
solution so as to enable direct comparison of the two methods. In a later case we will use a
much finer mesh to use as a reference solution for studying the convergence rate of SPH. For
both methods, the same specimen and boundary conditions were used.

Figure 6 shows the comparison between the SPH and the FE predictions of the von Mises
stress over time at point A. They agree well for both the early and uniform state stages of
compression. The maximum relative differences over the simulation were 4.9%, 5.5%, and
7.1% for points A, B and C, respectively. While using plane strain elements for this problem,
the difference in the von Mises stress predicted by SPH and FEM was quite large (11% at
point A) at the asymptotic limit indicating that the accuracy of the FEM solution is as much
unknown as it is for the SPH solution.

In the early stage, the FE solutions show considerable oscillations (Figure 6a) in comparison
with the SPH solutions which are smooth and non-oscillatory. These oscillations appear to be
Gibbs phenomena following the rapid initial rise in stress. The absence of such unphysical
oscillations in the SPH solution is a beneficial feature of this method. Figure 6b compares the Deleted: positive
Deleted: or
SPH and FE solutions over the long time (in the uniform state) for point A. The solutions
Deleted: M
agree well and show a steady linear increase in stress with no oscillations. The asymptotic
differences are 3.3%, 1.8% and 2.0% at the three representative locations.

Figure 7 compares the very early response of the stress at the representative locations. The FE
solution follows the same trend of early stress rise and its subsequent variation, as that of
SPH. The SPH solution shows a smooth behaviour, whilst the FE solution exhibits
considerable oscillations. The primary reasons for the oscillations in the early transient FE
solution can be attributed to the behaviour of plane stress elements used for modelling the
problem. These elements have some drawbacks when analysing elasto-dynamic problems as
they become over-stiff while simulating bending and in the case of incompressible solids,
tend to exhibit a phenomenon known as ‘locking’. Locking results from over-constrained
conditions and insufficient active degrees of freedom [64-67]. The inability to adequately
simulate bending and locking may lead to early oscillations in the stress solutions obtained in
an elastic analysis, as seen in Figures 6a and 7. One effective remedy to locking of the FEM
solution is the use of ‘reduced-integration’ elements which are computationally attractive
because they need evaluation of the matrices at a single point (centroid) of an element.
However, reduced-integration elements have a serious drawback as they give rise to spurious
energy modes, (known as ‘hourglass modes’), in which an element deforms to assume the
shape of an hourglass [68, 69]. This can produce oscillatory behaviour and considerably affect
the accuracy of the solutions, and consequently these elements were not used in the present
study. Rather, the current work used an FE model that employed quadratic quadrilateral
isoparametric elements with incompatible modes turned on, which minimises the effect of
locking.

In contrast SPH, being a mesh-free method, does not suffer from such problems because the
SPH particles have no underlying mesh structure to generate such spurious solution modes.
Also, the SPH interpolation uses more neighbouring particles including contributions from
further away leading to an inherently smoother solution.

6 Effect of Specimen Geometry and Material Property on the SPH Stress


Solution

Next we study the stress response of uniaxial test specimens with different geometries and
material properties. This is necessary to establish the robustness, accuracy and stability of the
SPH method for different test conditions since these influence the spatial and temporal
variations in stress and deformation. It is also important to know if the method works Deleted: that we

uniformly well for such variations.

6.1 Effect of different specimen geometries on the stress distribution

Here we investigate the effect of the geometry of the specimen on the stress field. The
specimen used earlier will be termed the ‘base specimen’. We consider rectangular specimens
with three different aspect ratios of the geometry and evaluate the variation in von Mises
stress with time. The aspect ratio is defined as the ratio of the width (W) to the height (H) of a
specimen. The loading and boundary condition for these specimens were the same as before.
The widths of the specimens were chosen to be twice, equal and half that of the base
specimen, respectively, giving aspect ratios of 0.8, 1.6, and 3.2.
Figure 8 shows the differences in the von Mises stress (at the centre) for the lower (0.8) and
higher aspect ratio (3.2) cases from that (1.6) of the base case. The differences are moderate in
early times, being 1.72% and 1.23% for the two cases at 1 ms. These reduce rapidly with
continued vertical compression, becoming less than 0.3% at 100 ms. Over longer times, the
von Mises stresses at the centre of the specimens are 7.982, 7.986, and 7.963 MPa (at 100 ms)
for the aspect ratios 0.8, 1.6, and 3.2, respectively. So the stresses in the specimens of
different geometries were virtually the same over longer times. The stresses of the higher and
lower aspect ratio specimens match that of the base specimen (which was verified against the
FEM solution in section 5). This demonstrates that the SPH solutions are robust and accurate Deleted: 6

and not affected by the geometry of the problem.

The time variation of the SPH stress solutions also accurately demonstrates the interaction
between the vertical stress due to compression and the horizontal stress due to lateral bulging.
The von Mises stress depends on both the vertical and horizontal stresses. Under uniaxial
compression, the horizontal (tensile) stress depends on the vertical compression and the lateral
width of the specimen. In early stages, the vertical and horizontal (lateral) stress waves are of
comparable magnitudes. So the von Mises stress is moderately sensitive to the horizontal
component of the stress and hence on the aspect ratio (width) of the specimen, as observed in
Figure 8. Over longer times, the vertical compressive stress becomes considerably larger than
horizontal (lateral) stress, and consequently the effect of the horizontal stress on the von
Mises stress becomes negligible. The vertical compressive strains for all three specimens are
the same because of their same height and same vertical compression. So the resulting vertical
compressive stresses are nearly the same (due to comparatively small lateral strains). As a
result, the von Mises stresses, being primarily dependent on the corresponding vertical
stresses, are essentially identical for the specimens of different aspect ratios. This establishes
that the SPH method produces stable and equally accurate stress solutions for a range of
specimen geometries under uniaxial compression and captures the influence of the lateral
strains (hence lateral dimension) on the stresses in the direction of compression.

6.2 Effect of different specimen material properties on the stress distribution

Next we examine whether the SPH predicted stress solutions are stable and equivalently
accurate for variations in materials properties. We use the same base case specimen as before
but with three different sets of material properties. All the specimens were subjected to the
same loading and boundary conditions as used before. For uniaxial loading, the predominant
deformation field is one dimensional; so Young’s modulus is the relevant material property
that controls the one-dimensional stiffness and hence the structural response of the specimen
for linear elastic material behaviour. So we vary the Young’s modulus with representative
values chosen to match rock (sandstone), metal (steel), and diamond with values of 7.47, 200,
and 950 GPa respectively.

Figure 9 shows the variation in the von Mises stress at the centre of the specimens with time.
For all of the materials, the stress variation is linear and stable. The von Mises stresses at the
specimen centres are 7.986, 194.7, 924.8 MPa (at 100 ms), which correlated well with the
equivalent FEM solutions and are nearly proportional to the respective Young’s moduli. This
stress variation is stable and consistent with the response of linear elastic materials under uni-
directional loading conditions and is an equally accurate solution for uniaxial compression
problems for a range of elastic properties.

7 Gravity Loading Case

Next we examine the stress response of the specimen under gravity only. The specimen, Deleted: just

discretised again as 1 mm SPH particles, was placed on a rigid platform under gravity. Figure Deleted: only

10 shows the distribution of the von Mises stress. The stress rapidly equilibrates to give an
almost linear vertical variation. The closer a location is to the bottom plate, the higher is the
stress due to the greater weight of the material above it.

Figure 11 shows the von Mises stress distributions along various horizontal and vertical
planes in the specimen. The stress variation along the vertical planes PP’ and QQ’ (shown in
Figure 11a) is consistent with the graded stress contour in Figure 10. The linearly increasing
stress variation from the top to the bottom surface demonstrates the ability of the current SPH
technique in predicting static solutions. The stress patterns along the horizontal planes taken
at different heights are shown in Figure 11b. The stress along each horizontal section remains
constant (planes TT’ and UU’), since each point on a horizontal section supports the same
weight of the material above. However, near the bottom of the specimen, the von Mises stress
remains slightly higher in the middle than at the two sides (e.g. along planes SS’ and RR’).
This occurs as a result of different constraining conditions on the material in the interior and
near the sides (free edges). When the specimen is vertically compressed, it deforms sidewise
to maintain incompressibility causing lateral bulging. Lateral movement is restricted in the
middle resulting in higher stresses, whereas the material near the free boundaries is more able
to move in response to the compression, relieving the stresses there. This effect is stronger
lower down because of the increasing weight above.
To verify the stress distribution predicted by SPH, this load case was also analysed using
FEM, with a fine resolution mesh size of 0.25 mm. Figure 12 compares the stress field under
gravity along plane PP’ (shown in Figure 11a) with the corresponding FE solution. The SPH
solution agrees very well with that obtained by FEM (within a maximum difference of 2%). Deleted: d

8 Solution Convergence with Increasing Particle Resolution

Next we evaluate the order of convergence with increasing particle resolution. The uniaxial
test was simulated with four other particle resolutions, 1.5, 2, 2.5 and 3 mm. The accuracy of
the SPH solutions was assessed using a very fine resolution (0.25 mm mesh) FE solution as a
reference solution. This fine-mesh FE solution was compared with the coarse-mesh (1 mm)
FE one. The differences at points A, B and C were uniform in time and were 0.030%, 0.005%
and 0.030%, respectively. This indicates that the FE solutions have converged well, and the
fine-mesh FE solution should be accurate enough to be taken as the reference solution.

The von Mises stress obtained for each SPH resolution case was compared with the fine mesh
FE solution at locations A, B and C. For all the particle resolutions the stress increases
linearly but there are small differences in the gradients of the stress variation. Figure 13 shows
the percentage difference of each of the SPH solutions from the FE solution at point A. The
differences fluctuate initially before becoming uniform by 5 ms. These occur during the early
wave propagation and are due to the Gibbs oscillations in the FEM solution. The difference is
largest for the coarsest SPH resolution and decreases monotonically with increasing
resolution. Table 1 gives the differences, averaged over the analysis domain, at three different
times during the simulation. The finest particle resolution leads to the lowest difference, with
the differences from the FE solution of 2.4% for 1 mm resolution up to 9.6% for 3 mm
resolution. This indicates that the SPH solution converges towards the fine mesh FE solution
as the particle spacing is progressively reduced.

Figure 14 shows the relative difference between the SPH and FE solutions on a log scale
versus the particle resolution. The relationship between the difference and the resolution is
linear and very close to constant with time (see Table 1). The order of accuracy, as estimated
from the gradient of the line of best fit through the accuracy-resolution plot, is 1.3.

In principle, the function approximations used in this SPH formulation are second order
accurate. The integral representation of a function is approximated in SPH using a smoothing
kernel. For the method to be second order accurate, the kernel approximation should satisfy a
normalisation condition that requires the integral of the kernel function over the region of
support to be unity. This is achieved when the region of support is entirely contained within
the problem domain, i.e. there are particles everywhere in the region of support. At the free
boundaries, a portion of the region of support lies outside the problem domain, and the kernel
approximation is truncated because of the absence of particles on one or more sides. As a
result, the normalisation condition is no longer satisfied at the boundaries, which contributes
to the degradation in the order of accuracy so that the method performs midway between a
first and a second order scheme. For fluid flows the errors are typically within the reasonable
range. For heat transfer they have been found to reduce the order of accuracy [9], and we Deleted: ,

observe here a similar loss of convergence accuracy for elastic solids. In the following
section, we investigate whether the boundary effect is responsible for the reduction in the
order of accuracy in the present case.

9 Comparison with Analytical Solution for an Infinite Width Specimen

It is noteworthy that the boundary condition implementation in classical SPH formulation


influences the solutions on the edges and surfaces. In classical SPH, field variables at a given
SPH particle (interpolation point) is computed by interpolation over a circular (in 2D) or a
spherical (in 3D) domain of influence. This can lead to unphysical solutions when evaluating
field variables (e.g. displacements and stresses) and their gradients, specifically for the first
layer of particles located on the boundaries (edges and surfaces) due to kernel truncation. For
the uniaxial test problem, this is reflected by the different stress values at the SPH particles on
the boundaries of the rectangular specimen (in Figure 2) compared to those of the adjacent
interior SPH particles. This effect is caused by the fact that the summation during
interpolation at a boundary SPH particle is over a semicircle or a hemisphere of particles. So a
set of interpolation points are not available for inclusion near a boundary, which affects the
calculation of interpolated values for the boundary particles, and is known as kernel
truncation error. To address this, one approach could be addition of multiple ‘fictitious’ layers
(usually two layers) of SPH particles at the boundary. This is demonstrated here by a specific
case for the piston, whereby the piston is modelled with two layers of SPH particles. This
provides sufficient SPH particles around the real boundary particles of the object, and thus
reduces the numerical boundary effects.

Also, evaluation of the accuracy with resolution for the uniaxial test indicates that the
boundaries adversely affect the solution. In this uniaxial test configuration, there are two sets
of boundaries, the top and bottom boundaries and the side boundaries, and they affect the
solution differently. For the particles on the side boundaries, the region of support of the
kernel extends over two layers of absent particles beyond the edge particles. For the particles
on the top and bottom boundaries, the region of support has a single layer of particles absent
because of the presence of the moving piston and the fixed support. As a result, the
normalisation errors (kernel truncation errors) at the side boundaries are comparatively larger
and therefore have a greater effect on the accuracy than the top and the bottom boundaries.

To test the inherent accuracy of the solution in the interior of the problem domain and the
validity of our explanation for the lower order of convergence, we consider a simplified
problem of a 2D specimen with an infinite width. This eliminates the effect of side edges. The
geometry and loading conditions are otherwise the same as shown in Figure 1, but with the
side boundaries now being periodic. This problem has an analytic solution for the vertical
principle stress when in the uniform state:

(23)
vt
 y,exact  E
L
where σy,exact is the analytical vertical principle stress, E is the Young’s modulus, v is the
uniform velocity of compression, t is the time of compression, L is the height of the specimen.

We consider two different SPH models to assess how the top and bottom boundaries affect the
solution. In the first case the top piston and the bottom plate are modelled using a single layer
of SPH particles. In the second case, two layers of particles are used, ensuring that the kernel
approximation satisfies the normalisation condition at the top and bottom boundaries of the
specimen, and the effect of kernel truncation at the free edges on the SPH solution is
completely eliminated.

Figure 15a shows the comparison of the SPH prediction of the vertical stress for a particle
resolution of 0.5 mm with the analytical solution for the single layer boundary case. The SPH
stress solution agrees very well with the analytical result. Figure 15b shows the difference of
the SPH solution from the analytical solution with time. After 10 ms the transient SPH
solution reaches a uniform state and can be compared to the exact solution. The difference
rapidly reaches an asymptotic value of 0.24%. The behaviour for the two layer solution is
similar and is not shown.

The order of convergence was again estimated by calculating the difference of the SPH
solutions from the FE solution for particle resolutions from 0.75 to 3.0 mm. Figure 16 shows
the variation of the difference with particle spacing (on a log scale) for both the single and Deleted: in

double layer cases. For the single layer objects, the asymptotic errors decrease from 7.1%
down to 0.6%. As the particle spacing is decreased, the accuracy of the SPH solution
improves and the solution converges towards the exact solution. The relationship between the
error and the resolution is linear. The order of convergence, estimated from the gradient of the
line of best fit, was 1.92. The order of convergence for the original finite uniaxial
configuration was only 1.3. It therefore improves significantly when the effects of side edges
are eliminated. For the double layer case, the asymptotic errors from the analytical solution
decline from 6.8% down to 0.5% and the corresponding order of convergence was 1.97. So
eliminating the kernel normalisation error at the top and bottom boundaries further improves
both the solution accuracy and the order of convergence to very close to its theoretical second
order level.

The difference in the order of convergence for these two cases establishes that for the
conventional SPH formulation of elastic solid deformation, the free edges can considerably
affect the solution accuracy. The accuracy obtained in the absence of boundary effects is very
close to the theoretical second order accuracy. This suggests that SPH variants that re-
normalise the kernel to account for the boundary weight discrepancy can reasonably be
expected to restore the second order accuracy if this is needed.

It is noteworthy that the boundary conditions were implemented following the method
outlined in Section 3.5 on the top and bottom edges of the specimen, which alleviates the
kernel truncation error. However, the kernel truncation still occurs at the free edges of the
finite width specimen along with the violation of the kernel normalization condition. This
affects the gradient calculations at the boundary, and leads to different stress values at the
boundary from that of the interior of the domain. However, the SPH solution at the interior of
the domain is not significantly affected when compared to the FE solution, as observed in the Deleted: with

present study. Similar features will not be observed in the FE solution as FEM uses a different
interpolation method (shape functions) compared to that of SPH (kernel functions).

10 Effect of the Tensile Instability Correction on SPH Solution Accuracy

In this paper, we focus on the principal stress modification approach proposed by Gray et al.
[15] for the tensile instability reduction, as discussed in section 2.3. This technique has been
found to be effective in producing good results for a variety of simple material deformation
problems [15, 16]. However, the correction factor needs to be carefully chosen to obtain a
stable solution with desired accuracy. An understanding of the trade-off between stability and
accuracy is important because a higher value of tensile coefficient may reduce instabilities in
some cases, but may also adversely affect the accuracy of the solution. This has not been
explored previously for material deformation problems.

From the work of Monaghan [46], it appears that the value of the tensile coefficient depends
on the physics of the problem. Based on the examples studied in [15], it was suggested that a
value of 0.3 usually provides stable solutions for 2D elasticity problems. However, the
authors did not evaluate the validity or accuracy of this solution with varying tensile
coefficients. We therefore analyse the uniaxial test problem to determine how the stress field
is affected by varying the tensile coefficient, using the values of 0.1, 0.2, 0.3 and 0.4.

10.1 Banded stress artefacts

We consider the von Mises stress variation along the horizontal mid-plane of the specimen
(plane TT’ in Figure 11b) for different tensile coefficients. Figure 17 shows the stress
variation at different times for tensile coefficients ψ = 0.2 and 0.3. For ψ = 0.1 (used in the
previous sections) the stress field is uniform and stable. The solution does not quantitatively
change much when increasing to ψ = 0.2 (Figure 17a). The stress increases as the compressive
load on the specimen increases with the spatial distribution being unaltered. When the value
of tensile coefficient is increased to 0.3 though, mesoscale banded stress patterns with
alternate high and low stress regions appeared during the approach of the uniform state
condition. This resulted in the fluctuating stress field shown in Figure 17b.

Figure 18 compares the stress distributions in the specimen for different tensile coefficients
and shows the formation of banded stress pattern for ψ = 0.3. The nature and time of
appearance of the stress bands depend on the particle resolution and distribution,
representative geometry and loading conditions. For the uniaxial compressive loading
configuration and a particle resolution of 1 mm set up in a regular square lattice pattern, the
banded structure first becomes visible at about 8 ms, whereas for ψ = 0.1, such a banded
structure does not appear at all.

Figure 19a shows the well-developed artificial banded stress distributions at 20 ms. The
spatial fluctuations remain nearly constant in amplitude over time. Detailed examination of
the SPH solutions suggests that the presence of these numerical artefacts does not seem to
affect the real underlying stress behaviour. The stress bands are prominent when the
background stress is low, as in Figure 19a. As the magnitude of the background compressive
stress increases, the relative contribution of the banded numerical stress declines and the
bands slowly fade, as seen in Figure 19b. Deleted: from view
In Figure 20, the SPH solutions for different tensile coefficients are compared against the fine
mesh FE solution at the centre of the specimen (point A). For the lower values of 0.1 and 0.2,
the solution difference is constant over time (discounting the initial fluctuations) and close to
3%. However, for ψ = 0.3 the difference varies substantially. At first it increases reaching
around 5% by 12 ms. The difference then rapidly drops and becomes negative reaching a peak
of near -4% at 20 ms. This corresponds to the initial appearance of the banded stress
structures. The error from the banded structure then declines with time and is quite close to Deleted: exponentially

the level for the lower tensile instability coefficients by 100 ms. The artificial banded stress
field that is superimposed on the physical stress field sharply degrades the accuracy of the
solutions.

Table 2 gives the difference between the SPH and FE stress solutions for point C at three
different times. This shows similar behaviour. The errors are consistently close to 2% for ψ =
0.1 and 0.2. For ψ = 0.3 there are large fluctuations reaching an extreme value of around -7%.
Over longer times (t = 100 ms), due to the increasing background stress, the relative
contribution of the constant magnitude banded stress fluctuations to the stress field declines
and the solution has comparable accuracy to that of the lower values of ψ. This shows that the
stress fluctuations are essentially constant in magnitude, do not grow and are essentially
benign. They do not seem to affect the evolution of the uniform state. For pure deformation
problems the stress artefacts have few consequences. However, for fracture modelling, the
banded stress pattern could lead to erroneous prediction of local damage initiation and
growth. For these problems, it is better to minimise or eliminate these artefacts, by choosing a
suitable value of the tensile coefficient. This is confirmed by Figure 21 which shows the von
Mises stress along the horizontal mid-plane for the three values of the tensile coefficient at
100 ms. For ψ = 0.1 and 0.2 the stress is constant across the domain. But for ψ = 0.3, there are
random stress fluctuations on the scale of the particles. The locally averaging stress is also
shown. It is also constant and is very close to the values for lower ψ. So although for higher ψ
there are high frequency oscillations that lead to locally inaccurate solutions, the filtered stress
fields are very close to those obtained with lower tensile coefficients.

Use of even higher values of the tensile coefficient results in extremely poor solutions with
very high stress fluctuations, suggesting that a value around 0.3 represents the limiting value
of the tensile coefficient that can be sensibly used in the present problem, which is the value
also used by Gray et al. [15] for 2D solid deformation analysis. Similar studies should be
conducted to determine an optimum trade-off value of the tensile coefficient that will result in
a stable solution with the required accuracy for different problems.
10.2 Effect of particle arrangement on the solution

The effect of high tensile coefficient on the solution quality (where a high value needs to be
used to stabilise the solution) can be moderated by the choice of the SPH particle
arrangement. In the previous case, the SPH particles were arranged in a regular fashion on a
square lattice (Figure 22a) which was aligned with the principal stress-strain directions. One
can imagine that this may contribute to the generation of the artificial banded stress pattern. Deleted: easily

To evaluate the importance of alignment of principal stresses with the particle structure, we Deleted: microstructure

consider two more different particle distributions, namely a diamond lattice (Figure 22b) and
a slightly disordered lattice (Figure 22c). The diamond lattice was generated by orienting the
regular square lattice structure at an angle of 45o to the direction of loading. The disordered
lattice structure was generated from the particle distribution of a uniform square lattice by
introducing slight randomness in the particle positions. This makes the structure more
‘isotropic’, and should alter the stress wave propagation paths. The degree of randomness is
controlled using a randomisation factor, which is the maximum allowable displacement in the
particle positions from the uniform spacing of a square lattice. For this case, we used a
randomisation factor of 0.5% of the particle spacing. The stress field was then calculated for
all three sets of particle arrangements with a tensile coefficient of 0.3.

Figure 23 shows the stress distributions for the different particle arrangements at 20 ms. The
diamond structure has eliminated the artificial banded stress pattern. However, the slightly
disordered lattice did not eliminate the banded stress pattern. Rather the presence of the
randomness in the particle positions altered the original longitudinal high stress bands to a
more quasi-random form. This shows that the introduction of noise in the particle positions
does not fundamentally change the alignment of the principle stresses and the preferred
connectivity in the particle arrangements.

Figure 24 shows the von Mises stress variation along the horizontal mid-plane for the
diamond lattice at 20, 60 and 100 ms. The uniformity of the stress at each time confirms the
elimination of the artificial banded stress pattern. The variation in von Mises stress with time
is linear and is indistinguishable for the three representative locations, which is similar to that
observed before for the square lattice with low tensile coefficients (base case solution). The
linear increase of the stress with time indicates that the diamond lattice produces the correct
elastic structural response. The change in the SPH particle structure to be not aligned with the Deleted: microstructure

principal stresses has suppressed the onset of banded stress pattern without any loss of
accuracy.
Figure 25 shows the relative deviation in von Mises stress from the FE solution for the three
different particle distributions at point A. For the square and disordered lattices, the onset of
banded stress pattern is noticeable with a higher accompanying difference from the FE Deleted: clearly

solution (up to 7%). The banded structure is most notable when the real compressive stress is Deleted: visible

low (around 20 ms). The magnitude of the relative difference then decreases steadily over
time with the increase in the overall stress level and asymptotes to 2.9%. The diamond pattern
does not have such fluctuations and there is no transition to a banded stress field.

In summary, the use of the square lattice particle structure produces erroneous banded stress
patterns in the early stage. Over longer times, it demonstrates the best accuracy for this
problem. The diamond lattice is effective in suppressing the banded stresses and has
reasonable comparative accuracy. The disordered lattice does not suppress the banded pattern
and is less accurate than the regularly aligned lattices.

11 Effect of artificial viscosity on the stability and accuracy of the solution

In SPH, an artificial viscosity is used to stabilise the solution [55]. The coefficients α and β in
the artificial viscous term (in Equation 17) are usually chosen to be 1.0 and 2.0 for elastic
solids and are generally found to produce stable solutions for the examples considered in [2].
Accordingly, these values were used for the earlier parts of this study. The effect of these
coefficients on the quality of the solution has not yet been explored for elastic deformation
problems. The coefficient α governs the linear term in the artificial viscosity and β scales the
quadratic term. Here we investigate how these coefficients affect the stability and accuracy of
the stress solution for the uniaxial test problem. Deleted: er

First, the effect of changing α while β is kept the same as used in the base case simulation (β =
2.0) is examined. The SPH solution remains stable and essentially invariant for 0.2  α  1.4.
Figure 26 shows the difference from the base case solution (α = 1.0) of the von Mises stress
for α = 0.2 and 1.4 at the specimen centre (point A). It is slightly higher in the early stages
with the maximum values being 2.1% and 4.5% for the two cases (Figure 26a). The difference
declines with time and becomes very small over longer times (Figure 26b). For the uniform
state, the two solutions (with α = 0.2 and 1.4) are almost identical to the base case solution
with the differences being less than 1 part in a million by 100 ms.

A value of α lower than 0.2 does not significantly affect the solution in the early stages.
However, as the uniform state is approached, the stress deviates from the base case solution,
indicating instabilities. Lower values of α ( 0.1) result in more rapid onset of instability. For
example, when the linear term is omitted from the artificial viscosity by setting α = 0, the
solution becomes immediately unstable. When the value of α was increased above 1.4, the
solution also rapidly becomes unstable. Further investigation shows that the range of stability Deleted: quickly

for α is independent of β. For uniaxial compression the stability range for α is from 0.2 to 1.4.

Next, the effect of the quadratic viscosity term on the solution is examined. The uniaxial test
is analysed for a variety of values of β, with α being kept constant at α = 1.0. Figure 27 shows
the difference in the von Mises stress at the specimen centre (point A) for β = 0.2 and 20 from
the base case solution (with β = 2.0). The SPH solution remains stable for both the cases, and
the accuracy was insensitive to the value of the coefficient β. The maximum differences in the
solution for β = 0.2 and 20 from the base case solution are 0.003% and 0.015% respectively.
Over longer times (after 20 ms), the solutions using the various values of β become
indistinguishable.

12 Advantages of SPH for Specific Applications

A fully mesh-less method, such as SPH, offers many advantages for modelling high
deformation and discontinuous processes, such as material forming, fracture and
fragmentation. The use of SPH eliminates the time-consuming effort of mesh generation and
refinement. Moreover, SPH can handle complex free surface deformations, which enables
prediction of surface contacts during metal forming and self-contact of surfaces after folding. Deleted: naturally predicts

SPH uses a Lagrangian approach, in which the solution frame of reference moves with the
particles. This allows history dependent properties of materials to be tracked [12]. In SPH, a
given particle always represents a specific mass of material and carries that information with
it. This is a critical attribute of Lagrangian methods. This means that information on the
precise state of each piece of material is recorded at all times, and the history of each piece of
material deformation and property evolution is built into the particle data. So SPH can predict
the thermal and mechanical history of the workpiece in a metal forming application without
the need for data mapping between evolving (discretised) geometric configurations. This
provides significant capability to track properties, such as cumulative plastic strain, damage,
metal composition, trapped gas, metallic phase, structure, grain flow, and surface oxide. Some Deleted: microstructure

or all of these properties can then be used to feed back into the flow dynamics using suitable
constitutive models. This also facilitates easy incorporation of rate-dependent deformation
behaviour, and so the effect of process parameters, such as axial feed rate, on the thermo-
mechanical responses can be included.
Another advantage of SPH is that it can easily handle coupled physical processes, such as
elastoplastic flow and heat transfer, in a unified analysis framework. The modelling of all
components of a multi-physics application using the same numerical framework means that
the physical laws, interfacial boundary conditions, and associated evolution of the material
properties and parameters can be seamlessly incorporated into an integrated analysis system.
SPH can also easily incorporate geometric nonlinearity by using a large strain formulation and
material nonlinearity due to its Lagrangian formulation. All these analysis features are
essential requirements of modelling material forming processes.

One particular area of application of SPH is the damage and fracture mechanics. Fracturing
processes are generally driven by the stress-strain history of the material. Traditional Eulerian
methods can experience difficulties in capturing this history on a node by node basis and in
predicting the evolution of damage. The history tracking ability of SPH should, in principle,
allow prediction of the damage initiation and crack propagation, and thus the dynamics of
damage evolution can be included explicitly.With grid based methods, damage propagation
often requires re-meshing to accommodate for the localised change in geometry. Automated
re-meshing can be computationally expensive and can lead to mesh distortion and inaccurate
results, especially in gradient computations [70]. The mesh-free nature of SPH enables large
scale deformation to be naturally handled without the need to re-mesh. So the combined
meshless-Lagrangian features of SPH make it ideally suited to modelling solids problems,
involving discontinuous large scale deformations and disintegrations.

13 Conclusions

This study has evaluated the numerical performance of the traditional (classical) SPH method
for predicting solid deformation under uniaxial compression conditions. The present paper has Deleted: using

demonstrated that SPH has good accuracy and stability properties for modelling both at the
scale of individual stress waves and for overall solid deformation. Specific conclusions
include:

 SPH is able to predict stress wave propagation and material deformation in uniaxial
compression. It can effectively model the early stress wave propagation through the
elastic specimen. The generation, reflection and superposition of the elastic waves are
well captured. The SPH solutions can predict attainment of spatial uniform stress state
condition and show no instability.
 SPH solutions have very good accuracy, as judged against both an analytical and a Deleted: The

fine-mesh finite element solution. In the early stages, the SPH stress predictions are
smooth whilst the FE solutions show some Gibbs like oscillations, which may be due
to locking and/or hour-glassing. However, the underlying trends in the early stress
wave solutions match well for both the methods. The uniform state response obtained
using SPH agrees well with that predicted by FEM.

 SPH stress solutions are accurate and stable for different specimen dimensions, and for Deleted: The

specimens made of different materials with a wide range of variation of properties. So


the SPH method is robust for elasto-dynamic solid deformation analysis under uniaxial
compression.

 SPH stress predictions for static load cases have very good accuracy, as evidenced by Deleted: The

stable behaviour and close agreement with FEM, demonstrating the ability of SPH to
simulate static stress analysis problems.

 SPH demonstrates adequate convergence characteristics for use in structural and


engineering analysis. Accuracy increases as the particle spacing is reduced as one
would expect. However, for the finite width test specimen, the order of convergence
was degraded by the presence of kernel normalisation errors at the boundaries. This
reduces the order of accuracy to only 1.3. The order of convergence was improved to
1.94 when an infinite width specimen (which eliminates the side boundaries errors)
was used. The order of convergence was further improved to 1.97 when the edge
effects at the top and bottom boundaries were removed. This is very close to the
theoretical second order of accuracy possessed by the spatial discretisation.

 SPH was found to be stable for uniaxial compression. The difference between the SPH Deleted: The
Deleted: method
and FE results remains essentially constant with time in the uniform state. For all of
the SPH solutions with different particle resolutions, the stress response is linear
elastic as expected. Deleted: (
Deleted: it should be)
 Uniaxial compression does not strongly exhibit the tensile instability, and so a low
value of tensile instability correction coefficient can be used to produce a stable SPH
solution. When higher values of the tensile coefficient are used, artificial stress bands
appear superimposed on the real stress field. The magnitude of the banded stress
fluctuations is relatively small, does not vary much with time, and does not seem to
affect the overall dynamics of the problem. These bands are prominent when the
background stress is low, and they become a smaller proportion of the total stress as
the real stress increases and so fade with time.

 Alignment of the particle structure with the principal axes of the stress field is a key Deleted: microstructure

factor in the generation of artificial stress bands. Re-orienting the SPH lattice pattern
so that the particle structure is not aligned to the principal stresses eliminates the Deleted: microstructure

banded stress pattern. If tensile instability is to be eliminated via a higher tensile


coefficient, aligning the particle distributions with the principal stress field should be
avoided.

 The linear term in the SPH artificial viscosity is found to play an important role in
stabilising SPH solid deformation solutions. For uniaxial compression, the values of α
leading to a stable and accurate solution behaviour ranges from 0.2 to 1.4. Within this
range, the stability and accuracy of the solutions are essentially insensitive to the
specific values used. The quadratic term in the artificial viscosity does not affect either
the stability or the accuracy of the SPH solutions for elastic deformation problems.
14 Reference

[1] Monaghan, J.J., 1992. Smoothed particle hydrodynamics. Ann. Rev. Astron.
Astrophys., vol. 30, pp. 543-574.
[2] Monaghan, J.J., 2005. Smoothed particle hydrodynamics. Rep. Prog. Phys., vol. 68,
pp. 1703–1759.
[3] Gingold, R.A., and Monaghan, J.J., 1977. Smoothed particle hydrodynamics - Theory
and application to non-spherical stars. MNRAS, vol. 181, pp. 375–389.
[4] Lucy, L.B., 1977. A numerical approach to the testing of the fission hypothesis.
Astron. J., vol. 82, pp. 1013–1024.
[5] Monaghan, J.J., and Price, D.J., 2001. Variational principles for relativistic smoothed
particle hydrodynamics. Monthly Notices of the Royal Astronomical Society, vol. 328
(2), pp. 381-392.
[6] Cleary, P.W., 1998. Modelling confined multi-material heat and mass flows using
SPH. Applied Mathematical Modelling, vol. 22 (12), pp. 981-993.
[7] Cleary, P.W., Ha, J., Prakash, M., and Nguyen, T., 2005. Simulation of casting
complex shaped objects using SPH. In in San Francisco, CA, United States. pp. 317-
326. Minerals, Metals and Materials Society, Warrendale, PA 15086, United States.
[8] Cummins, S.J., and Rudman, M.J., 1998. Truly incompressible SPH. In in
Washington, DC, USA. pp. 8. ASME, Fairfield, NJ, USA.
[9] Cleary, P.W., and Monaghan, J.J., 1999. Conduction modelling using smoothed
particle hydrodynamics. Journal of Computational Physics, vol. 148 (1), pp. 227-264.
[10] Bonet, J., and Kulasegaram, S., 2002. A simplified approach to enhance the
performance of smooth particle hydrodynamics methods. Applied Mathematics and
Computation (New York), vol. 126 (2-3), pp. 133-155.
[11] Cleary, P., Ha, J., Alguine, V., and Nguyen, T., 2002. Flow modelling in casting
processes. Applied Mathematical Modelling, vol. 26 (2), pp. 171-190.
[12] Cleary, P.W., Prakash, M., Ha, J., Stokes, N., and Scott, C., 2007. Smooth particle
hydrodynamics: status and future potential. Progress in Computational Fluid
Dynamics, vol. 7 (2-4), pp. 70-90.
[13] Libersky, L.D., and Petschek, A.G., 1990. Smooth particle hydrodynamics with
strength of materials. In Advances in the Free-Lagrange Method, ed. Trease and
Crowley, Berlin: Springer.
[14] Strength Modeling in SPHC. Wingate, C.A., and Fisher, H.N., 1993. Los Alamos
National Laboratory, Report No. LA-UR-93-3942.
[15] Gray, J.P., Monaghan, J.J., and Swift, R.P., 2001. SPH elastic dynamics. Computer
Methods in Applied Mechanics and Engineering, vol. 190 (49-50), pp. 6641-6662.
[16] Cleary, P.W., Prakash, M., and Ha, J., 2006. Novel applications of smoothed particle
hydrodynamics (SPH) in metal forming. Journal of Materials Processing Technology,
vol. 177 (1-3), pp. 41-48.
[17] Das, R., and Cleary, P.W., 2008. The potential for SPH modelling of solid
deformation and fracture. In IUTAM Proceedings book series volume on "Theoretical,
Computational and Modelling Aspects of Inelastic Media", ed. D. Reddy, pp. 287-
296, Capetown, South Africa: Springer.
[18] Karekal, S., Das, R., Mosse, L., and Cleary, P.W., 2011. Application of a mesh-free
continuum method for simulation of rock caving processes. International Journal of
Rock Mechanics and Mining Sciences, vol. 48 (5), pp. 703-711.
[19] Cleary, P.W., Prakash, M., Das, R., and Ha, J., 2012. Modelling of metal forging using
SPH. Applied Mathematical Modelling, vol. 36 (8), pp. 3836-3855.
[20] Das, R., and Cleary, P.W., 2013. A mesh-free approach for fracture modelling of
gravity dams under earthquake. International Journal of Fracture, vol. 179 (1-2), pp.
9-33.
[21] Das, R., and Cleary, P.W., 2010. Effect of rock shapes on brittle fracture using
Smoothed Particle Hydrodynamics. Theoretical and Applied Fracture Mechanics,
vol. 53 (1), pp. 47-60.
[22] Fagan, T., Das, R., Lemiale, V., and Estrin, Y., 2012. Modelling of equal channel
angular pressing using a mesh-free method. Journal of Materials Science, vol. 47
(11), pp. 4514-4519.
[23] Islam, S., Ibrahim, R., Das, R., and Fagan, T., 2012. Novel approach for modelling of
nanomachining using a mesh-less method. Applied Mathematical Modelling, vol. 36
(11), pp. 5589–5602.
[24] Bradley, G.L., Chang, P.C., and McKenna, G.B., 2001. Rubber modeling using
uniaxial test data. Journal of Applied Polymer Science, vol. 81 (4), pp. 837-848.
[25] Liu, W.K., Jun, S., Li, S., Adee, J., and Belytschko, T., 1995. Reproducing kernel
particle methods for structural dynamics. International Journal for Numerical
Methods in Engineering, vol. 38 (10), pp. 1655-1679.
[26] Chen, J.K., Beraun, J.E., and Jih, C.J., 1999. Improvement for tensile instability in
smoothed particle hydrodynamics. Computational Mechanics, vol. 23 (4), pp. 279-
287.
[27] Liu, M.B., and Liu, G.R., 2006. Restoring particle consistency in smoothed particle
hydrodynamics. Applied Numerical Mathematics, vol. 56 (1), pp. 19-36.
[28] Bonet, J., and Kulasegaram, S., 2000. Correction and stabilization of smooth particle
hydrodynamics methods with applications in metal forming simulations. International
Journal for Numerical Methods in Engineering, vol. 47 (6), pp. 1189-1214.
[29] Bonet, J., and Kulasegaram, S., 2001. Remarks on tension instability of Eulerian and
Lagrangian corrected smooth particle hydrodynamics (CSPH) methods. International
Journal for Numerical Methods in Engineering, vol. 52 (11), pp. 1203-1220.
[30] Vidal, Y., Bonet, J., and Huerta, A., 2007. Stabilized updated Lagrangian corrected
SPH for explicit dynamic problems. International Journal for Numerical Methods in
Engineering, vol. 69 (13), pp. 2687-2710.
[31] Dyka, C.T., and Ingel, R.P., 1995. An approach for tension instability in smoothed
particle hydrodynamics. Computers and Structures vol. 57, pp. 573-580.
[32] Dyka, C.T., Randles, P.W., and Ingel, R.P., 1997. Stress points for tension instability
in SPH. International Journal for Numerical Methods in Engineering, vol. 40 (13),
pp. 2325-2341.
[33] Randles, P.W., and Libersky, L.D., 2000. Normalized SPH with stress points.
International Journal for Numerical Methods in Engineering, vol. 48 (10), pp. 1445-
1462.
[34] Vignjevic, R., Campbell, J., and Libersky, L., 2000. A treatment of zero-energy modes
in the smoothed particle hydrodynamics method. Computer Methods in Applied
Mechanics and Engineering, vol. 184 (1), pp. 67-85.
[35] Belytschko, T., and Xiao, S., 2002. Stability analysis of particle methods with
corrected derivatives. Computers and Mathematics with Applications, vol. 43 (3-5),
pp. 329-350.
[36] Xiao, S.R., and Belytschko, T., 2005. Material stability analysis of particle methods.
Advances in Computational Mathematics, vol. 23 (1-2), pp. 171-90.
[37] Shaw, A., and Roy, D., 2012. Stabilized SPH-based simulations of impact dynamics
using acceleration-corrected artificial viscosity. International Journal of Impact
Engineering, vol. 48, pp. 98-106.
[38] Shaw, A., Roy, D., and Reid, S.R., 2011. Optimised form of acceleration correction
algorithm within SPH-based simulations of impact mechanics. International Journal
of Solids and Structures, vol. 48 (25-26), pp. 3484-3498.
[39] Shaw, A., and Reid, S.R., 2009. Applications of SPH with the acceleration correction
algorithm in structural impact computations. Current Science, vol. 97 (8), pp. 1177-
1186.
[40] Shaw, A., and Reid, S.R., 2009. Heuristic acceleration correction algorithm for use in
SPH computations in impact mechanics. Computer Methods in Applied Mechanics
and Engineering, vol. 198 (49-52), pp. 3962-3974.
[41] Wang, S.L.N., 2011. A large-deformation Galerkin SPH method for fracture. Journal
of Engineering Mathematics, vol. 71 (3), pp. 305-318.
[42] Wong, S., and Shie, Y., 2008. Large Deformation Analysis with Galerkin based
Smoothed Particle Hydrodynamics. Cmes-Computer Modeling in Engineering &
Sciences, vol. 36 (2), pp. 97-118.
[43] Gray, J.P., and Monaghan, J.J., 2004. Numerical modelling of stress fields and fracture
around magma chambers. Journal of Volcanology and Geothermal Research, vol.
135, pp. 259– 283.
[44] Swegle, J.W., Hicks, D.L., and Attaway, S.W., 1995. Smoothed Particle
Hydrodynamics Stability Analysis. Journal of Computational Physics, vol. 116 (1),
pp. 123-134.
[45] Morris, J.P., 1996. A study of the stability properties of smooth particle
hydrodynamics. Publications of the Astronomical Society of Australia, vol. 13 (1), pp.
97-102.
[46] Monaghan, J.J., 2000. SPH without a tensile instability. Journal of Computational
Physics, vol. 159 (2), pp. 290-311.
[47] Melean, Y., Sigalotti, L.D.G., and Hasmy, A., 2004. On the SPH tensile instability in
forming viscous liquid drops. Computer Physics Communications, vol. 157 (3), pp.
191-200.
[48] Liu, Z.S., Swaddiwudhipong, S., and Koh, C.G., 2004. High velocity impact dynamic
response of structures using SPH method. International Journal of Computational
Engineering Science, vol. 5 (2), pp. 315-326.
[49] Monaghan, J.J., 1994. Simulating free surface flows with SPH. J. Computat. Phys.,
vol. 110, pp. 399–406.
[50] Kulasegaram, S., Bonet, J., Lewis, R.W., and Profit, M., 2003. High pressure die
casting simulation using a Lagrangian particle method. Communications in Numerical
Methods in Engineering, vol. 19 (9), pp. 679-687.
[51] Cedric, T., Janssen, L.P.B.M., and Pep, E., 2005. Smoothed particle hydrodynamics
model for phase separating fluid mixtures. I. General equations. Physical Review E
(Statistical, Nonlinear, and Soft Matter Physics), vol. 72 (1), pp. 016713.
[52] Cleary, P.W., Ha, J., Prakash, M., and Nguyen, T., 2006. 3D SPH flow predictions and
validation for high pressure die casting of automotive components. Applied
Mathematical Modelling, vol. 30 (11), pp. 1406-1427.
[53] Fang, J., Owens, R.G., Tacher, L., and Parriaux, A., 2006. A numerical study of the
SPH method for simulating transient viscoelastic free surface flows. Journal of Non-
Newtonian Fluid Mechanics, vol. 139 (1-2), pp. 68-84.
[54] Imaeda, Y., and Inutsuka, S.-i., 2002. Shear Flows in Smoothed Particle
Hydrodynamics. The Astrophysical Journal, vol. 569 (1), pp. 501-518.
[55] Monaghan, J.J., and Gingold, R.A., 1983. Shock simulation by the particle method
SPH. Journal of Computational Physics, vol. 52 (2), pp. 374-389.
[56] Monaco, A.D., Manenti, S., Gallati, M., Sibilla, S., Agate, G., and Guandalini, R.,
2011. SPH modeling of solid boundaries through a semi-analytic approach.
Engineering Applications of Computational Fluid Mechanics, vol. 5 (1), pp. 1-15.
[57] Libersky, L.D., Randles, P.W., Carney, T.C., and Dickinson, D.L., 1997. Recent
improvements in SPH modeling of hypervelocity impact. International Journal of
Impact Engineering, vol. 20 (6-10 pt 2), pp. 525-532.
[58] Libersky, L.D., Petscheck, A.G., Carney, T.C., Hipp, J.R., and Allahdadi, F.A., 1993.
High strain Lagrangian hydrodynamics- a three dimensional SPH code for dynamic
material response. Journal of Computational Physics, vol. 109 (1), pp. 67-75.
[59] Randles, P.W., and Libersky, L.D., 1996. Smoothed particle hydrodynamics: Some
recent improvements and applications. Computer Methods in Applied Mechanics and
Engineering, vol. 139 (1-4), pp. 375-408.
[60] Mayrhofer, A., Rogers, B.D., Violeau, D., and Ferrand, M., 2013. Investigation of
wall bounded flows using SPH and the unified semi-analytical wall boundary
conditions. Computer Physics Communications vol. 184 (11), pp. 2515-2527.
[61] Wu, B., and Tan, C.P., 2002. Sand Production Prediction of Gas Field: Methodology
and Laboratory Verification. In SPE Asia Pacific Oil & Gas Conference and
Exhibition, in Melbourne, Australia.
[62] Timoshenko, S., and Goodier, J.N., 1984. Theory of elasticity. 3rd ed. McGraw-Hill.
[63] Bathe, K.J., 1995. Finite Element Procedures Prentice-Hall, Englewood Cliffs.
[64] Babuska, I., and Suri, M., 1992. On locking and robustness in the finite element
method. SIAM Journal on Numerical Analysis, vol. 29 (5), pp. 1261-1293.
[65] Suri, M., Babuska, I., and Schwab, C., 1995. Locking effects in the finite element
approximation of plate models. Mathematics of Computation, vol. 64 (210), pp. 461.
[66] Ozkul, T.A., and Ture, U., 2004. The transition from thin plates to moderately thick
plates by using finite element analysis and the shear locking problem. Thin-Walled
Structures, vol. 42 (10), pp. 1405-1430.
[67] Suri, M., 1996. Analytical and computational assessment of locking in the hp finite
element method. Computer Methods in Applied Mechanics and Engineering, vol. 133
(3-4), pp. 347-371.
[68] Hansbo, P., 1998. New approach to quadrature for finite elements incorporating
hourglass control as a special case. Computer Methods in Applied Mechanics and
Engineering, vol. 158 (3-4), pp. 301-309.
[69] Reese, S., and Wriggers, P., 2000. Stabilization technique to avoid hourglassing in
finite elasticity. International Journal for Numerical Methods in Engineering, vol. 48
(1), pp. 79-109.
[70] Fernandez-Mendez, S., Bonet, J., and Huerta, A., 2005. Continuous blending of SPH
with finite elements. Computers and Structures, vol. 83, pp. 1448-1458.
FIGURES

Moving piston

Specimen

Fixed end plate

Figure 1: Uniaxial test configuration with uniform velocity loading.


(a) t = 40 μs (b) t = 50 μs

(c) t = 60 μs (d) t = 80 μs

(e) t = 90 μs (f) t = 110 μs


Figure 2: Stress wave propagation from the top piston towards the bottom plate (von Mises stresses
are in Pa).
(g) t = 120 μs (h) t = 160 μs

(i) t = 190 μs (j) t = 240 μs

(k) t = 280 μs (l) t = 380 μs


Figure 2 (ctd): Reflected wave propagation from the bottom plate and their interaction with the waves
originating from the piston at the top.
(a) t = 40 ms (b) t = 70 ms
Figure 3: Uniform state response with continuously increasing stress pattern with time (and therefore
load) shows little change in stress pattern over the entire specimen (stresses are in Pa).

h h’
A C

v’
Figure 4: Representative points in the specimen used for monitoring von Mises stress variation.
12
Point A
von Mises Stress (KPa)

10 Point B
Point C
8

0
0 20 40 60 80 100
Time (s)
(a) Very early stress variation with time.

0.30
Point A
von Mises Stress (MPa)

0.25 Point B
Point C
0.20

0.15

0.10

0.05

0
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Time (ms)
(b) Early stress variation with time.

9
Point A
von Mises Stress (MPa)

8
Point B
7 Point C
6
5
4
3
2
1
0
0 20 40 60 80 100
Time (ms)
(c) Stress variation over longer times when the system has reached uniform state.

Figure 5: Stress variation in response to loading at the representative points.


0.30
SPH
von Mises Stress (MPa)

0.25 FE

0.20

0.15

0.10

0.05

0
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
(a) Time (ms)
9
SPH
von Mises Stress (MPa)

8
FE
7
6
5
4
3
2
1
0
0 20 40 60 80 100
(b) Time (ms)
Figure 6: Comparison of SPH and FEM values of the von Mises stress at the specimen centre (point
A): a) early stage, and b) in the long term uniform state.
12000
SPH
von Mises Stress (Pa)

10000 FE

8000

6000

4000

2000

0
0 20 40 60 80 100
(a) Point A Time (s)
12000
SPH
von Mises Stress (Pa)

10000 FE

8000

6000

4000

2000

0
0 20 40 60 80 100
(b) Point B
Time (s)
12000
SPH
von Mises Stress (Pa)

10000 FE

8000

6000

4000

2000

0
0 20 40 60 80 100
(c) Point C Time (s)

Figure 7: Very early response of the SPH and FE solutions at the three representative locations.
Difference in VM Stress (%) 14
Diff. between H/W = 1.6 & 0.8
12 Diff. between H/W = 1.6 & 3.2
10

0
0 2 4 6 8 10
Time (ms)
Figure 8: Differences in the von Mises stress (at the centre) for the lower (W/H = 0.8) and higher
aspect ratio (W/H = 3.2) specimens from that of the base specimen (W/H = 1.6).

1000
E = 7.50 GPa
von Mises Stress (MPa)

800 E = 200 GPa


E = 950 GPa
600

400

200

0
0 20 40 60 80 100
Time (ms)
Figure 9: Comparison of the time variation of von Mises stress distribution for three different materials.
(a) t = 50 ms (b) t = 100 ms
Figure 10: von Mises stress distribution in the specimen subjected to gravity loading (stresses are in
Pa).

P Q

3000
von Mises Stress (Pa)

2500 PP’ plane

2000
QQ’ plane
1500

1000

500

0
0 20 40 60 80 100 120 140
Height (mm) y

(a) x

P’ Q’
3000
RR’ plane
von Mises Stress (Pa)

2500

U U’
2000
SS’ plane
1500
T T’
1000 TT’ plane

500
UU’ plane
S S’
0
-40 -30 -20 -10 0 10 20 30 40
R y R’
X (mm)
(b) x

Figure 11: von Mises stress variation in the specimen under gravity at t = 50 ms along; (a) vertical
planes at planes PP’ (x = 0 mm) and QQ’ (x = 20 mm), and (b) horizontal planes at various heights:
RR’ (y = 20 mm), SS’ (y = 35 mm), TT’ (y = 70 mm), and UU’ (y = 105 mm).
3000
SPH
von Mises Stress (Pa)

2500 FE

2000

1500

1000

500

0
0 20 40 60 80 100 120 140
Height (mm)
Figure 12: Comparison of SPH and FEM prediction of von Mises stresses at various heights (along
plane PP’) at t = 50 ms.

1 mm 1.5 mm 2 mm 2.5 mm 3 mm
14
Difference in VM Stress (%)

12

10

0
0 20 40 60 80 100
Time (ms)
Figure 13: Difference in von Mises stress from the FE solution at point A for different SPH particle
resolutions.

10

8
Difference (%)

1.0 1.5 2.0 2.5 3.0


Particle Resolution (mm)

Figure 14: Variation of the difference between the SPH and FE solutions with particle spacing at 100
ms.
9
Analytical
8
Vertical Stress (MPa)

SPH
7
6
5
4
3
2
1
0
0 20 40 60 80 100
(a) Time (ms)
4
Difference in Stress (%)

0
0 20 40 60 80 100
(b) Time (ms)
Figure 15: (a) Comparison of the von Mises stress obtained using SPH with the analytical solution for
an infinite width specimen, and (b) variation of the difference from the analytical solution with time. Deleted: V

8
6 Single layer object
Double layer object
4
Difference (%)

0.5 1 1.5 2 2.5 3


Particle Resolution (mm)
Figure 16: SPH solution accuracy with particle resolution for an infinite width specimen as compared
against the analytical solution.
12
20 ms
von Mises Stress (MPa)

10 60 ms
100 ms
8

0
-40 -30 -20 -10 0 10 20 30 40
(a) X (mm)
12
20 ms
von Mises Stress (MPa)

10 60 ms
100 ms
8

0
-40 -30 -20 -10 0 10 20 30 40
(b) X (mm)

Figure 17: von Mises stress distributions along the horizontal mid-plane are compared for different
values of tensile coefficient: a) tensile coefficient = 0.2, b) tensile coefficient = 0.3. The results for
tensile coefficient = 0.1 are similar to a).

a) Tensile coefficient = 0.1 b) Tensile coefficient = 0.3


Figure 18: von Mises stress (in Pa) distributions at 20 ms for different tensile coefficients.
(a) t = 20 ms (b) t = 100 ms
Figure 19: Comparison of von Mises stress (in Pa) distributions at early (20 ms) and longer times (100
ms) for tensile coefficient = 0.3.

6
Difference in VM Stress (%)

-2

-4
Tensile coeff. 0.1
-6 Tensile coeff. 0.2
Tensile coeff. 0.3
-8
0 20 40 60 80 100
Time (ms)
Figure 20: Difference of the SPH solution from the FE results at point A for different tensile
coefficients.
12
Tensile coeff. 0.1
von Mises Stress (MPa)

11
Tensile coeff. 0.2
10 Tensile coeff. 0.3
Tensile coeff. 0.3 (avg.)
9
8
7
6
5
4
-40 -30 -20 -10 0 10 20 30 40
X (mm)
Figure 21: von Mises stress distributions along the horizontal mid-plane are compared for different
values of tensile coefficient, the magenta line represents the (locally) averaged stress distribution
obtained from ψ = 0.3 curve (the green curve) at 100 ms.

a) b) c)
Figure 22: Distribution of particles relative to loading direction in three different cases: a) Square
lattice, b) Diamond lattice, and c) Disordered lattice.

(a) Square lattice (b) Diamond lattice (c) Disordered lattice


Figure 23: Effect of particle distribution on the von Mises stress pattern (in Pa) when a high tensile
coefficient (0.3) is used, banded stress pattern appears at an early stage for the square and
disordered lattice (at 20 ms) when the background stress is low.
12
20 ms
von Mises Stress (MPa)

10 60 ms
100 ms
8

0
-40 -30 -20 -10 0 10 20 30 40
X (mm)
Figure 24: Comparison of von Mises stress distributions along the horizontal mid-plane at various
times for the diamond lattice with ψ = 0.3. Deleted: simulation

8
Difference in VM Stress (%)

-2
Square lattice
-4 Diamond lattice
Disordered lattice
-6
0 20 40 60 80 100
Time (ms)
Figure 25: Deviation in von Mises stress from the FE solution at point A with a tensile coefficient of 0.3
for different particle lattice structures.

6
Difference in VM Stress (%)

Diff. between  = 1.0 & 0.2


4 Diff. between  = 1.0 & 1.4

-2

-4
0 2 4 6 8 10
Time (ms)
(a) t = 0 - 10 ms
6
Difference in VM Stress (%)
Diff. between  = 1.0 & 0.2
4 Diff. between  = 1.0 & 1.4

-2

-4
0 20 40 60 80 100
Time (ms)
(b) t = 0 - 100 ms

Figure 26: Difference in von Mises stress obtained using α = 0.2 and 1.4 from the base case solution
(with α = 1.0) for a constant β = 2.0.

0.02
Difference in VM Stress (%)

Diff. between  = 2.0 & 0.2


0.015 Diff. between  = 2.0 & 20.0

0.01

0.005

-0.005

-0.01
0 2 4 6 8 10
Time (ms)
(a) t = 0 - 10 ms

0.02
Difference in VM Stress (%)

Diff. between  = 2.0 & 0.2


0.015 Diff. between  = 2.0 & 20.0

0.01

0.005

-0.005

-0.01
0 20 40 60 80 100
Time (ms)
(b) t = 0 - 100 ms

Figure 27: Difference in von Mises stress obtained using β = 0.2 and 20.0 from the base case solution
(with β = 2.0) for a constant α = 1.0.
TABLES

Table 1: Deviation in von Mises stress (averaged over the problem domain) between the FE and SPH
solutions with time for different particle resolutions.

Difference (%)
Particle
Resolution (mm)
t = 20 ms t = 60 ms t = 100 ms

1 2.35 2.35 2.35 Deleted: mm

1.5 3.53 3.54 3.55 Deleted: mm

2 5.39 5.37 5.37 Deleted: mm

2.5 7.35 7.33 7.32 Deleted: mm

3 9.61 9.59 9.57 Deleted: mm

Table 2: Difference in von Mises stress between the FE and SPH solutions at point C for different
tensile coefficients.

% Difference for point C


Time (ms)
ψ = 0.1 ψ = 0.2 ψ = 0.3
20 1.96 1.97 -6.69

60 1.96 1.97 1.07

100 1.96 1.97 1.44

View publication stats

Вам также может понравиться