You are on page 1of 170

A STRAIN-RATE DEPENDENT TENSILE DAMAGE MODEL FOR

BRITTLE MATERIALS UNDER IMPACT LOADING

SAGAR DAS

A thesis submitted in fulfilment of the requirements for the degree of Doctor of

Philosophy in the School of Civil Engineering at

The University of Sydney

2016
ACKNOWLEDGEMENTS

I would like to express the sincere gratitude to my advisor, Dr. Luming Shen, School of Civil

Engineering at the University of Sydney, for his valuable guidance, inspiration, generous

time rendered to me to prepare this thesis. His intellectual guidance had made me overcome

all the academic obstacles during my study in The University of Sydney. His constructive

advice and guidance throughout this research have been the significant influence for me to

reach this far.

I want to extend my gratitude to my co-advisors, Dr. Giang D. Nguyen, School of Civil,

Environmental and Mining Engineering at The University of Adelaide and Yixiang Gan,

School of Civil Engineering at the University of Sydney for their encouragement and support

to the various stages of this work. A special thanks to Dr. E.A. Flores-Johnson for many

valuable discussions and the fruitful directions to solve different issues that raised during the

FEM modelling with Abaqus. I would like to thank my Annual Progress Review committee,

especially Dr. Gwenaelle Proust and Dr. Pierre Rognon for their valuable suggestions for this

research.

I want to thank all the technical staff in J.W. Roderick Materials and Structures lab, The

University of Sydney, for their kind help and time both in training and to get the necessary

setups to run the experiments. Among them, special thanks go to Mr. Garry Towell and Mr.

Paul Burrell for their kind directions and the skills that were essential to run the experimental

facilities in the laboratory.

I must acknowledge to all the post graduate students (research) in our school for their

friendship and advice during my study period. Among them, the contribution from Ling Li,

i
Mrintunjay Sahu, Ghasem Nagib, Mahfuzur Rahman, Imran Kabir, Tanim Ahmed, Arif

Khan, Arghya Das, Shaheen Mahmood, Chongpu Zhai and Fahim Ahmed are highly

appreciated.

I would like to extend my sincere gratitude to my parents, Mr. Shyamal Chandra Das and

Mrs. Pritikana Das and my brother, Mr Shekhar Das for their affection, love and

unconditional supports throughout the every step in my life. From my childhood, their

encouragements have given me the valuable directions to reach so far. A very special thanks

to my wife, the lovely soul-mate Mrs Sharmistha Roy (Tuli) for her tolerance, understanding

and the inspiration that helped me to keep going strong in the midst of several challenges.

Finally, I greatly acknowledge the financial support from the University of Sydney

International Scholarship (USydIS) to bear my educational and living expenses during my

study in The University of Sydney.

ii
TABLE OF CONTENTS

ACKNOWLEDGEMENTS i

TABLE OF CONTENTS iii

LIST OF FIGURES vii

LIST OF TABLES xiii

LIST OF SYMBOLS xiv

LIST OF ABBREVIATIONS xviii

ABSTRACT xix

CHAPTER 1: INTRODUCTION 1

1.1 BACKGROUND 1

1.2 PROBLEM STATEMENT AND RESEARCH CONTRIBUTION 5

1.3 OBJECTIVE 5

1.4 METHODOLOGY 6

1.5 THESIS OUTLINE 6

CHAPTER 2: LITERATURE REVIEW 8

2.1 INTRODUCTION 8

2.2 DYNAMIC CHARACTERISTICS OF MATERIAL 8

2.3 THEORY OF SPLIT HOPKINSON PRESSURE BAR EXPERIMENT 12

2.4 DYNAMIC TENSILE EXPERIMENT 15

iii
2.4.1Dynamic Brazilian Disc test 15

2.4.2 Spalling test 19

2.5 TENSILE DAMAGE MODELS 22

2.6 FRAGMENT SIZE MODELS 23

2.7 DYNAMIC SIMULATION WITH BRAZILIAN DISC 27

2.8 DRUCKER PRAGER MODEL 29

2.9 SUMMARY 31

CHAPTER 3: MECHANICAL CHARACTERISATION OF SANDSTONE 33

3.1 INTRODUCTION 33

3.2 MATERIAL AND SAMPLE PREPARATION 34

3.3 EXPERIMENTAL TECHNIQUE 35


3.3.1 Quasi-static Testing 35

3.3.2 Dynamic Brazilian Disc test with SHPB 37

3.4 RESULT AND DISCUSSION 39

3.4.1 Quasi-static behaviour of sandstone 39

3.4.2 Dynamic behaviour of sandstone 44

3.5 SUMMARY 55

CHAPTER 4: CONSTITUTIVE MODELLING: DAMAGE MODEL


AND FRAGMENT SIZE MODEL 56

4.1 INTRODUCTION 56

4.2 CONSTITUTIVE MODELING FOR DYNAMIC TENSION 56

4.3 FRAGMENT SIZE MODEL 60

4.3.1 Determination of strain energy 61

4.3.2 Determination of surface energy 63

iv
4.3.3 Determination of fragment size 63

4.4 DETERMINATION OF MODEL PARAMETERS 68

4.4.1 Model parameter α and β 68

4.4.2 Determination of and 71

4.5 SUMMARY 72

CHAPTER 5: CALIBRATION AND UTILIZATION OF THE MODEL 73

5.1 INTRODUCTION 73

5.2. IMPLEMENTATION OF DAMAGE MODEL TO SUBROUTINE 73

5.3 SIMULATION OF DYNAMIC RESPONSE FOR OIL SHALE 76

5.4 SIMULATION OF DYNAMIC RESPONSE FOR CONCRETE 82

5.4.1 Spall experiment of concrete 82

5.4.2 Mesh convergence study 84

5.4.3 Spall simulation with damage model 86

5.5 VALIDATION OF FRAGMENT SIZE MODEL 92

5.6 SUMMARY 94

CHAPTER 6: APPLICATION: SIMULATION OF DYNAMIC


BRAZILIAN DISC TEST 95

6.1 INTRODUCTION 95

6.2 MODELLING GEOMETRY OF FEM STUDY 95

6.3 DYNAMIC LOADING AND BOUNDARY CONDITION FOR SHPB 97

6.4 MATERIAL MODEL 98

6.5 MESH CONVERGENCE STUDY 98

6.6 RESULTS AND DISCUSSION 100

v
6.6.1 Damage contour and fracture pattern 100

6.6.2 Fracture Strength and strain rate 109

6.7 SUMMARY 114

CHAPTER 7: CONCLUSION AND FUTURE WORK 115

7.1 CONCLUSION 115

7.2 FUTURE WORK 118

REFERENCES 120

APPENDIX 127

vi
LIST OF FIGURES

Fig. 2.1 Dynamic fragmentation of brittle solid under tensile loading (Meyers,

1994): (a) Material with pre-existed flaws (b) Interaction of microcracks

to from new cracks (c) Fragmentation of brittle solid (d) Unloaded

regions around the crack.

Fig 2.2 Crack formation of brittle material under compressive loading (Meyers,

1994).

Fig 2.3 Geometrical and microstructural sites for the initiation shear failure

(Meyers et al., 1994): (a) strain concentration (b) Inhomogeneity in

external loading (c) Fracture of second phase particle (d) Pile-up release

by shear loading.

Fig. 2.4 1D stress wave propagation for split Hopkinson pressure bar theory.

Fig. 2.5 Schematic of Brazilian disc test with split Hopkinson pressure bar testing

facility.

Fig. 2.6 Dynamic tensile test with Brazilian disc sample, (a) Fracture pattern at

high strain rate (Zhang and Zhao, 2013) (b) Stress distribution (Tension

and compression) from experiment (Chen et al., 2014a) (c) Stress

distribution (Tension and compression) from simulation (Zhou et al.,

2013).

Fig. 2.7 Schematic of spalling test with Hopkinson bar testing facility.

Fig. 2.8 Typical demonstration of particle velocity history at the free end of

vii
sample in spall experiment.

Fig. 2.9 Yield surface of P- stress plane.

Fig. 3.1 Experimental setup for quasi static compressive test using Instron

machine.

Fig. 3.2 Experimental setup for quasi static tension test for Brazilian disc sample

Fig. 3.3 Schematic view of Brazilian Disc test with Split Hopkinson Pressure bar

testing facility.

Fig. 3.4 A Brazilian disc sample is placed between the input and output bars

keeping the axis of gauges normal to loading axis of the bars.

Fig. 3.5 Stress strain curve obtained from sandstone at strain rate 10-5 s-1.

Fig. 3.6 Recovered sample after the quasi static compression test.

Fig. 3.7 Time load history of Brazilian disc test at quasi static loading condition.

Fig. 3.8 Recovered sample after the quasi static tension test.

Fig. 3.9 Strain history of steel bars at impact velocity 15.2 m/s.

Fig. 3.10 Recovered Brazilian disc sample for impact velocity of 11.1 m/s.

Fig. 3.11 Recovered Brazilian disc sample for impact velocity of 15.2 m/s.

Fig. 3.12 Recovered Brazilian disc sample for impact velocity of 20.4 m/s.

viii
Fig. 3.13 Force history at the two ends of the sample at impact velocity 15.2 m/s.

Fig. 3.14 Strain and strain rate history at the sample centre for impact velocity 15.2

m/s.

Fig. 3.15 Strain rate history at the centre of BD sample at projectile impact (a) 11.1

m/s (b) 20.4 m/s.

Fig. 3.16 Force history of the output bar at impact velocity 11.1 m/s.

Fig. 3.17 Force history of the output bar at impact velocity 15.2 m/s.

Fig. 3.18 Force history of the output bar at impact velocity 20.4 m/s.

Fig. 3.19 A plot between dynamic fracture strength and strain rate in log-log space.

Fig. 3.20 A plot between tensile dynamic increase factor (TDIF) with strain rate.

Fig. 4.1 Typical mean tension vs volumetric strain curve obtained using the

proposed damage model.

Fig. 4.2 Fragment formation hypothesis by Grady (1988). (a) Spherical fragment,

(b) Time-mean tension plot.

Fig. 4.3 Fragment formation hypothesis in current model (a) Prolate spheroid

fragment, (b) Time-mean tension response.

Fig. 5.1 Flowchart for the damage model for the numerical implementation.

Fig. 5.2 Strain rate vs fracture stress for oil shale (Grady and Kipp, 1980)

Fig. 5.3 Numerical simulation of oil shale under uniaxial loading (a) Schematic

ix
with loading and boundary condition (b) FEM meshing of the cylinder

Fig. 5.4 Axial stress-strain curves of oil shale under different uniaxial tensile

strain rates.

Fig. 5.5 Uniaxial stress and damage histories for oil shale at strain rate of 10 s-1.

Fig. 5.6 The effect of microcrack growth rate (α) on the stress-strain curves of oil

shale under tensile strain rate 10 s-1.

Fig. 5.7 The effect of strain rate sensitivity parameter β on the stress-strain curves

of oil shale under tensile strain rate 10 s-1.

Fig. 5.8 Schematic of spall experiment using Hopkinson bars.

Fig. 5.9 Logarithmic plot of fracture stress versus strain rate obtained from spall

experiment (Schuler et al., 2006)

Fig. 5.10 Mesh convergence study for impact velocity 7.6 m/s, with mesh size (a)

8 mm (b) 4 mm (c) 2 mm and (d) 1 mm

Fig. 5.11 Axial Stress history to first macrocrack (right hand side of the sample)
for different mesh sizes under impact 7.6 m/s

Fig. 5.12 Damage evolution of spall test by projectile impact 7.6 m/s at time (a)

105μs (b) 112 μs (c) 150 μs

Fig. 5.13 Comparison of experimental fracture locations (Schuler et al., 2006) and

simulated damage contour under projectile impact velocities of (a) 4.1

m/s, (b) 7.6 m/s and (c) 11.1 m/s.

Fig. 5.14 Uniaxial stress-strain curves at the fracture planes under projectile impact

x
velocities.

Fig. 5.15 Tensile strength and strain rate evolution for an impact velocity of 7.6

m/s

Fig. 5.16 Histories of axial strain rate of concrete at three differen timpact

velocities

Fig. 6.1 The FEM model of Brazilian disc simulation with split Hopkinson

pressure bar setup.

Fig. 6.2 Experimental obtained incident stress pulses at the mod length location

of input bar for three impact velocities.

Fig. 6.3 Damage contour of the disc at 150 ms by element size (a) 0.8 mm (b) 0.4

mm (c) 0.2 mm and (d) 0.1 mm.

Fig. 6.4 Mesh convergence study for Brazilian disc test for different mesh sizes.

Fig. 6.5 Tensile stress distribution of the sample at time 36 ms under impact

velocity 11.1.

Fig. 6.6 Tensile stress evolution at the two ends of the disc at impact velocity

11.1 m/s.

Fig. 6.7 Tensile stress evolution at the two ends of the disc at impact velocity

15.2 m/s.

Fig. 6.8 Tensile stress evolution at the two ends of the disc at impact velocity

20.4 m/s.

xi
Fig. 6.9 Damage evolution of Brazilian disc under impact velocity of 11.1 m/s at

time: (a) 0.35 ms (b) 0.38 ms and (c) 0.5 ms.

Fig. 6.10 Damage evolution of Brazilian disc under impact velocity of 15.2 m/s at

time: (a) 0.34 ms (b) 0.36 ms and (c) 0.5 ms.

Fig. 6.11 Damage evolution of Brazilian disc under impact velocity of 20.4 m/s at

time: (a) 32 ms, (b) 35 ms and (c) 0.50 ms.

Fig. 6.12 Damage contour at the end of simulation (at time of 0.50 ms) under

impact velocity (a) 11.1 m/s (b) 15.2 m/s and (c) 20.4 m/s.

Fig. 6.13 Tensile stress (σy) evolution at the centre of Brazilian disc at three impact

velocities.

Fig. 6.14 The strain rate history at the centre of the Brazilian disc for three impact

velocities.

Fig. 6.15 Evolution of tensile strength and strain rate at the centre of the disc at the

impact of 11.1 m/s.

Fig. 6.16 Evolution of tensile strength and strain rate at the centre of the disc at the

impact of 15.2 m/s.

Fig. 6.17 Evolution of tensile strength and strain rate at the centre of the disc at the

impact of 20.4 m/s.

Fig. 6.18 Strain rate and fracture strength data for dynamic Brazilian disc test

which are obtained from experiment and simulation.

xii
LIST OF TABLES

Table 3.1 Result obtained from quasi static compression experiment of sandstone

Table 3.2 Results obtained from quasistatic tension test with Brazilian disc

samples.

Table 3.3 The summary of dynamic Brazilian disc test with different strain rates.

Table 5.1 Comparison of fracture strength and fracture location obtained from

experiments (Schuler et al., 2006) and numerical simulations.

Table 5.2 Fragment sizes predicted by the proposed model and estimated from

experiments by Schuler et al. (2006)

Table 6.1 Comparison of fracture strength and strain rate obtained from

experiments and numerical simulations.

xiii
LIST OF SYMBOLS

A Cross sectional area of a sample

Ab Cross sectional area of a bar

As Specific surface area of a fragment

Aprolate Specific surface area of a prolate spheroid fragment

a The length of a fragment

a1 Diameter of a fragment

aeq Equivalent diameter of the fragment

b The width of a fragment

Cd Microcrack density

̇ Microcrack density rate

Cg Crack propagation velocity

c Elastic wave speed of a material

cb Elastic wave speed of a bar

ck Bulk wave speed of a material

D Damage parameter

d Cohesion of a material

ds Diameter of the sample

E Young’s modulus of a material

Eb Young’s modulus of a bar

F Force

F1 Force obtained by 1-wave method

F2 Force obtained by 2-wave method

F3 Force obtained by 3-wave method

Fd Dynamic force in output bar

xiv
f1 ( ̅ Effective Poisson’s ratio function

f1(v) Poisson’s ratio function

G Shear modulus of a material

̅ Effective shear modulus

I1 First stress invariant

J1 First invariant of deviatory stress

J2 Second invariant of deviatory stress

J3 Third invariant of deviatory stress

K Bulk modulus of the material

̅ Effective bulk modulus

KIc Fracture toughness of a material

kDP Ratio between tensile and compressive triaxial strengths

m The ratio between fragment width and fragment length

N Number of microcrack

n fragment shape constant

P Mean stress or pressure

Pd Mean tension or pressure at dynamic loading

Pst Mean tension or pressure at quasi static loading

q Equivalent stress

R Initial radius of the dilating volume under a dynamic loading

Sf Surface area of a fragment

t Time

Spall time

tcr Time required to reach the critical strain

tf Time required to fracture. i.e. complete separation of the body

Time required to reach the peak strength

tm Time required to reach the maximum stress

xv
u Displacement

u1 Displacement at sample input bar end

u2 Displacement at sample output bar end

uI Displacement caused by the incident pulse in input bar

uR Displacement caused by the reflected pulse in input bar

uT Displacement caused by the transmitted pulse in output bar

Vf Volume of a fragment

ν Poisson’s ratio

̅ Effective Poisson’s ratio

vv Velocity

vej Ejection velocity

vpull Pullback velocity

α The model parameter for the damage model

β The model parameter for the damage model

Kronecker delta

ɛ strain

ɛcr Volumetric strain at the static tensile strength of the material

ɛij Strain tensor

ɛv Volumetric strain

ɛy uniaxial strain along the axis y

̅ Equivalent plastic strain

̇ Volumetric strain rate

̇ Uniaxial strain rate

̇ The numerical average of strain rate at three major axes

Fracture surface energy per unit area

ρ Density of the material

Փ Slope of the linear yield surface in the pressure-pseudo effective stress plane

xvi
Flow potential for Drucker-Prager model

Ψ Dilation angle

Pseudo effective stress

σc Yield strength of the material at uniaxial compression

σst Quasi static tensile strength of the material under uniaxial loading

σtd Dynamic static tensile strength of the material under uniaxial loading

σij Stress tensor

Elastic strain energy until the critical strain of the material

Strain energy between critical strain and the volumetric strain to which the
peak strength is reached

Strain energy between volumetric strain at peak strength and the volumetric
stain to fracture.

Elastic energy

xvii
ABBREVIATIONS

2D Two-dimensional

3D Three-dimensional

FEM Finite element method

LVDT Linear variable differential transformer

SPHB Split Hopkinson pressure bar

S-I Sample input bar face

S-O Sample output bar face

TDIF Tensile dynamic increase factor

xviii
A STRAIN-RATE DEPENDENT TENSILE DAMAGE MODEL FOR

BRITTLE MATERIALS UNDER IMPACT LOADING

ABSTRACT

Brittle materials are often subjected to high strain rate impact load, which could be imposed

due to intentional demolition purposes or during ballistic impact on protective structures.

Fragments of different sizes are generally observed by such impact, which are directly related

to the strain rate experienced by the material at different locations. This thesis presents a rate-

dependent constitutive model to predict such dynamic behaviour of brittle solid under tensile

loading. A three-parameter rate dependent tensile damage model, under continuum

mechanics framework, is developed for simulating the fragmentation of brittle materials

subjected to high strain-rate loading. The damage model is formulated under the assumption

that the isotropic and homogeneous material contains initial microcracks and the microcrack

induced damage increases when a critical volumetric strain is exceeded. Considering the

microcrack induced damage and energy into account, a quantitative and direct method is

developed to determine the fragment size under a constant strain rate loading. In this model,

instead of assuming spherical fragment, more realistic prolate spheroid fragment is assumed,

which eventually determines the more accurate surface energy from a fragment. In addition,

complete strain energy (until the fracture of the material) is considered which improves the

global energy balance in predicting the size of a fragment. The parameters of this model can

be conveniently calibrated by experimental data on fracture strength and strain rate. The

proposed rate-dependent model is validated by a spall experiment of concrete and with a

xix
dynamic Brazilian disc experiment of sandstone. Both of these experiments are numerically

simulated with the proposed model and the experimental observations are compared with the

simulation. The predicted strain rate, fracture strength, fracture location and fragment size are

in very good agreement with those obtained in the experiments.

xx
CHAPTER 1

INTRODUCTION

1.1 BACKGROUND

Dynamic response of brittle material plays a significant role in engineering applications. The

scope ranges from protective structures design with concrete, ceramic and glass (Luccioni et

al., 2004; Medvedovski, 2010; Mohammed et al., 1996; Norville and Conrath, 2006 and

Sands et al., 2009), to demolition related applications such as mining, in-situ bed preparation

of oil shale (Boade et al., 1981; Esen et al., 2003; and Kanchibolta, 2003), etc. The dynamic

load could be imposed to a structure in the form of transient impulse loading such as collision

of two bodies, blast explosion, rapid temperature change, high energy radiation etc.

Sometimes, those external dynamic loads are applied for the purpose of demolition, whereas

other applications encompass to protect those from the impulse load. Whatever is the purpose

of application, in addition to essential experimental investigations, an accurate constitute

model to study the dynamic failure process is necessary for understanding such high strain

rates problems through numerical analysis. In addition, several engineering applications such

as aggregate and mining industries postulate the demolition of brittle material in a fashion

that fragments of a specific distribution has to be achieved by a high strain rate loading. A

fragment size prediction model could be a convenient tool to design such high strain rate

problem with the aid of numerical analysis.

1
The dynamic characteristics of brittle materials such as strength and the degree of

fragmentation were observed to be dependent on strain rate. Brittle materials show higher

strength with the increase of strain rates; specifically the strength is more rate sensitive to

tension than compression loading (Suaris and Shah, 1984). Moreover, the fragment size

distribution was observed to be a function of strain rate. At low strain rate loading, the

interactions of microcracks have such characteristics that only a few large fragments

dominated the fragmentation process. In contrast, fragmentation process under high strain

rate was more complex with numerous small cracks initiating at internal material defects,

then propagating at high speed, and finally coalescing to form many small fragments (Ma and

An, 2008; Taylor et al., 1986; and Vocialta and Molinari, 2015). In order to better understand

the dynamic fracture process of brittle solid, several approaches such as theoretical,

experimental and numerical investigations had brought considerable attention in literature

(Bieniawski, 1967; Clifton, 2000; Curran et al., 1987; Cotsovos and Pavlović, 2008;Grady,

1988; Sharon and Fineberg, 1999; Huang et al., 2006; Ma et al., 1998; Shen, 2009; Wang et

al., 2007; and Wu and Shen, 2013).

Several rate-dependent damage models were formulated concerning microcrack interaction

and growth under dynamic loading (Fahrenthold, 1991; Grady and Kipp, 1980; Hamdi et al.,

2011; Liu and Katsabanis, 1997; Lu and Xu, 2004; Suaris and Shah, 1984; Taylor et al.,

1986; Wang et al., 2008; Yang et al., 1996 and Yazdchi et al., 1996). Although several

damage models are available in literature to capture fragmentation of brittle solid under high

strain rate loading, inclusion of empirical constants and assumptions have deviated some

models from sound mechanical explanation to some extent. For example, Hamdi et al. (2011)

considered that dynamic tensile strength of the material was reached at a damage value of

0.2, and based on this numeric value they determined one of the model parameters. However,

2
the numeric value of fragmentation limit (0.2) was originated from a match between the

results from a blast test and the predictions from a numerical simulation for oil shale (Grady

and Kipp, 1980 and Taylor et al., 1986) without any further explanation to justify this value.

Liu & Katsabanis, (1997) derived a numerical damage value of 0.632 to determine one of the

model constants that determines microcrack growth rate in the model. This value was termed

as fragmentation limit (a threshold value of damage or microcrack density in dynamic

loading when fragments are imminent) of material and was originated from an assumption

that microcracks coalescing to form a fragment occurred when microcrack density became 1.

In other work (Zhang et al., 2006), fragmentation limit was considered as 0.22 and the

corresponding microcrack density was determined as 0.25 in fragmentation model.

Calibration of a rate-dependent model usually requires considerable efforts, particularly for

models that contained several parameters that must be determined from complex and

unconventional experiment setups. Not to mention that some models contained parameters

whose value were difficult to obtain with sophisticated experimental facility (Li et al., 2004).

Moreover, it was found that some models postulated an indirect approach to get model

constants, where iterations of several parameters had to be performed to match experimental

data. For example, a dynamic damage evolution law in brittle solid was developed by

analysing attenuation of sound wave caused by the growth of microcracks (Wang et al.,

2008). The calibration of this model (Wang et al., 2008) was not straightforward and it

required iterations of four model parameters by a numerical study to match experimental

data. Saksala, (2010) developed a viscoplastic damage model for rocks under impact loading.

The model required nine model parameters, which are not readily available from

experimental efforts. This model (Saksala, 2010) was implemented to a numerical study

(Saksala et al., 2013) where it was observed that it required a parameter study with model

3
parameters, since some parameters are difficult to obtain by experiment. The complex nature

in dynamic experiments and the scarcity of specialized experiment facility had led to the

indirect approaches in model calibration. Hence, a damage model, which is straightforward in

calibration and easy in determining the model parameters with conventional experimental

facilities, is desired.

By considering the conservation of total energy, several fragment size prediction models

were developed (Glenn and Chudnovsky, 1986; Grady, 1988, 1982; Zhang et al., 2006). In

those models, the average fragment size was expressed as a function of strain rate. In fact,

several fragmentation size models (Glenn and Chudnovsky, 1986; Grady and Kipp, 1980;

Grady, 1988, 1982; Lu and Xu, 2004; Zhang et al., 2006) assumed that fragments were

spherical, which inevitably underestimated the surface area of the fragment as most

fragments have irregular shapes. As a result, the predicted fragment size was not accurate.

Also the conventional energy balance assumptions (Glenn and Chudnovsky, 1986; Grady,

1988, 1982; Zhang et al., 2006) neglected the strain energy contribution during the post-peak

stress softening of a material. This is in contrast with experimental observations which

revealed that quasi brittle materials exhibit significant strain energy between strain at peak

stress and strain at failure (Albertini and Montagnani, 1994). However, very few studies had

tried to address this issue in order to better estimate the fragment size. Among those few

works, the significance of the fragment shape was discussed by Rouabhi et al. (2005) where a

power law relation of elastic strain energy with fragment size was introduced to

accommodate the irregularity of the fragments in their model. In order to calibrate the

additional parameter for incorporating fragment shape irregularity in their model (Rouabhi et

al., 2005), a cylinder blast study which demanded specialised facility for explosion was

required.

4
1.2 PROBLEM STATEMENTAND RESEARCH CONTRIBUTION

It has been found that some damage models were subjected to assumptions that include

empirical constants (such as fragmentation limit) in model formulation. Moreover, there had

been a conflicting explanation to identify these empirical constants in modelling of brittle

solids. Besides, some damage models have been difficult to calibrate as there were no straight

forward ways to determine model constants. Eventually, an indirect approach by iteration of

parameters had been adopted instead of using direct experimental data for those constants. On

the other hand, it has been observed that the fragment size prediction can be improved when

the irregularity in fragment shapes and the complete strain energy until fracture of a fragment

could be considered in formulation. Addressing such ambiguity in the sense of empirical

constants in the modelling and the efforts that are required to calibrate a model, a damage

model which is straight forward in calibration and sound in formulation is very desirable. In

addition, a model which is convenient to use yet would not compromise with high prediction

capability would be beneficial. Therefore, the major contribution of this research would be to

develop a convenient tool to simulate the fracture and fragment size under a dynamic tensile

load with a high prediction capability.

1.3 OBJECTIVE

The aim of this thesis is to develop a constitutive model to simulate high strain rate impact

phenomena for brittle solid. In order to capture the strain rate dependent fracture in brittle

solid, a tensile damage model has been developed in this thesis. The objective of the thesis

has also been extended to ascertain the size of a fragment under high strain rate tensile

5
loading. Apart from the modelling, the predictions of the models were compared to the

experimental observations.

1.4 METHODOLOGY

The proposed damage model is developed within the formulation of continuum damage

mechanics where it is utilised to predict high strain rate fracture behaviour of brittle solid.

The proposed damage model considers the degradation of the material stiffness due to

interaction and nucleation of microcracks under dynamic tensile loading. The fragment size

prediction model is developed by balancing the energy during the dynamic fragmentation

process of the solid. The microcrack induced damage is also brought into consideration in the

fragment size model. In order to improve the prediction of the fragment size, the conventional

energy balance is improved with two important aspects: a) The post strength peak strain

energy is brought into consideration and b) Conventional sphere fragment assumption is

replaced with a prolate spheroid assumption which could predict the surface area of a

fragment more accurately. The validations of the models are shown by the numerical

implementation of the models to dynamic tensile experiments and comparing the

experimental result with the numerical predictions.

1.5 THESIS OUTLINE

Chapter 2 contains the review of relevant literature for this study. The dynamic tensile

behaviour of brittle solid is explained followed by the experiment techniques in dynamic

tension testings. The review of microcrack induced damage models and fragment size

distribution models are provided. Numerical studies with dynamic Brazilian disc test are also

6
summarized. Finally a brief review of Drucker Prager model is provided. Chapter 3 shows

both static and dynamic experiment of sandstone. Quasi static behaviour of sandstone is

studied in both tension and compression loading. The dynamic tensile characteristic of

sandstone is studied with Brazilian disc test. Chapter 4 is devoted to the constitutive

modelling for the dynamic fracture and fragmentation of brittle solid under impact loading. A

rate dependent tensile damage model is developed in addition to a fragment size prediction

model in this chapter. The details of formulation and the way to determine the model

constants are described in this chapter, following by Chapters 5 and 6 that show the

numerical implementation and validation of the models. Chapter 5 shows a numerical

implementation of the constitutive model to a spall experiment of concrete, and the both

observations (numerical and experimental) are compared. The scope of Chapter 6 is to

numerically study the dynamic behaviour of sandstone (Brazilian disc specimens) with the

same experimental setup used in Chapter 3. The numerical results are compared with

experimental observation to validate the damage model. Along with comparing the

simulation with experiment results, the macrocrack initiation and propagation in Brazilian

disc is studied in this chapter with the proposed damage model. A summary of the entire

work is provided in Chapter 7.

7
CHAPTER 2

LITERATURE REVIEW

2.1 INTRODUCTION

This chapter serves as background information which is mostly related to experimental

facility and modelling approach that have been utilised in this work. A brief review of

dynamic mechanical characteristics of brittle material is discussed, followed by the

experimental theory of split Hopkinson pressure bar (SHPB), and implementation of

SHPB to the Brazilian disc and spalling tests. The rate dependent tensile damage models

that are relevant to this work and have been developed by monitoring growth of

microcrcaks under the applied loadings are briefly reviewed. Fragment size distribution

models formulated by the energy balance during fragmentation are discussed. Finally, a

review of dynamic tensile study with a Brazilian disc sample is performed, followed by a

short discussion of the Drucker-Prager model.

2.2 DYNAMIC CHARACTERISTICS OF MATERIAL

Dynamic failure and fragmentation of brittle solid have been a topic of interest during the

last few decades. The experimental or combinations of experimental and theoretical study

revealed the failure process from microstructure observation to macrostructure response of

brittle material at dynamic loading conditions (Cho et al., 2003; Grady and Kipp, 1979;

8
Nemat-Nasser and Deng, 1994; and Shockey et al., 1974). The dynamic failure of brittle

materials can be classified into three major categories namely tensile, compressive and

shear failure. In this section the microstructural aspects in the failure of these three

categories are briefly described.

Interaction and nucleation of pre-existent micro defects play a significant influence of

tensile failure of material. Among the several experimental studies, a noteworthy

observation of dynamic tensile failure was reported by Shockey et al. (1974) who

explained the fragmentation process in four major stages: activation of inherent flaws, the

growth of cracks, coalescence of cracks and fragmentation of the solid. A similar

explanation was concluded for the dynamic fragmentation by Meyers (1994) and is shown

in Figure 2.1. Figure 2.1(a) shows the pre-existed flaws of brittle solid which under tensile

loading propagate (Shown in Figure 2.1(b)) with a velocity and intercept with each other to

form fragments (Shown in Figure 2.1(c)). During the failure process, each growing crack

generates an unloading region to which no cracks are further generated as shown in Figure

2.1(d). On the other hand, compressive failure is generally caused by the unstable growth

of the compression induced tensile cracks. By the virtue of inhomogeneity in the solid,

localised region of tension is developed by the remote compression applied to it. The

tensile cracks that are generated by this compressive load are termed as wing cracks.

Figure 2.2 shows the wing crack developed by the far field compression load induced to

the material. Failure of material under compressive loading is caused by the growth and

interaction of those wing cracks. Based on such interaction of wing cracks, Nemat-Nasser

and Deng, (1994); and Paliwal and Ramesh, (2008) developed a micromechanical model

to study the dynamic failure of material under compressive loading.

9
New cracks from Unloaded
Microcracks Fragment
interactions region

(a) (b) (c) (d)

Figure 2.1. Dynamic fragmentation of brittle solid under tensile loading (Meyers, 1994):

(a) Material with pre-existed flaws (b) Interaction of microcracks to from new cracks (c)

Fragmentation of brittle solid (d) Unloaded regions around the crack.

Pre-existed
microcrack

Wing crack

(a) (b)
Figure 2.2. Crack formation of brittle material under compressive loading (Meyers, 1994).

Shear failure of material can be triggered by the external, geometrical factors or internal

microstructure reasons (Meyers, 1994a). Based on geometric consideration, the

concentration of stain can localise and may cause the material to fail under shear (See

Figures 2.3(a) and 2.3(b)). The microstructural sites for the shear concentration are shown

in Figures 2.3(c) and 2.3(d). In fact, the microstructural site of shear localization is

preceded by the local softening by some mechanism. When the grain contains a fractured

second phase particle, the fracture may be initiated by the sliding of the fractured

10
interfaces. Another microstuructral explanation could be the dislocation piles-ups that can

pierce a gain and can lead to shear band formation. However, a comprehensive review of

such dynamic shear failure can be found in Meyer, (1994a).

(a) (b)

Second phase
particle

(c)

Grain

(d)

Slipping plane
caused by pile-up

Figure 2.3 Geometrical and microstructural sites for the initiation shear failure (Meyers et

al., 1994): (a) strain concentration (b) Inhomogeneity in external loading (c) Fracture of

second phase particle (d) Pile-up release by shear loading.

11
2.3 THEORY OF SPLIT HOPKINSON PRESSURE BAR EXPERIMENT

Split Hopkinson pressure bar or Kolsky bar testing facility is one of the most widely used

loading techniques which was developed by Kolsky (1949). A comprehensive review of

this testing method including historical background and recent advances of the method can

be found in Chen and Song (2011). The theory of the method relies on the one dimensional

stress wave propagation in thin and straight bar. The bar is assumed to have uniform cross

section and homogeneous throughout the length. It is assumed in split Hopkinson bar

theory that the sample deforms uniformly and the effect of inertia and friction are

negligible. During the experiment, 1D stress pulse has to be generated from a liner elastic

impact and the wave propagation is assumed to be dispersion free. Figure 2.4 shows the

propagation of the incident (ɛI), reflected (ɛR) and transmitted (ɛT) strain wave in input and

output bars in the split Hopkinson bar theory. The sample is sandwiched between input and

output bars as shown in Figure 2.4. The displacements of the sample-bar interface are

denoted by u1 and u2.

Impacted end u1 u2

ɛI ɛR ɛT

Input bar Sample Output bar

Figure 2.4. 1D stress wave propagation for the split Hopkinson pressure bar theory.

The Equation of 1D wave propagation is shown in Eq. (2.1)

12
Here, x is the displacement along the loading axis, t is time and cb is the elastic wave speed

of input and output bars. The solution of Eq. (2.1) for input bar can be regarded as:

The solution for the output bar can be written as:

Here, f1, f2 and f3 are three arbitrary functions for the displacement caused by incident

wave (uI) reflected wave (uR) and transmitted wave (uT), respectively. The strain can be

defined by

Differentiating Eq. (2.2) with respect to x and t, we have Eqs. (2.5) and (2.6), respectively

̇ ( )

Differentiating Eq. (2.3) we have:

Eqs. (2.6) and (2.7) are valid everywhere for input bar and output bar respectively. If the

stress wave propagation in the sample is neglected then the average strain rate ( ̇ is as

follows:

̇ ̇
̇

Here, Hs denotes the sample height. Substituting Eqs. (2.6) and (2.7) in Eq. (2.8), we have:

13
̇

The assumption of the split Hopkinson bar is that the specimen is in stress/force

equilibrium. Therefore, the force that develops at the input bar-sample (F1) and output bar-

sample (F2) face should be equal. The force at two ends of the bars are determined from a

two-wave and one-wave method as shown in Eqs. (2.10) and (2.11), respectively.

Here, Ab and Eb are the cross section and Young’s modulus of the Hopkinson bars.

Eventually, the force equilibrium condition from Eqs. (2.10) and (2.11) yields,

Using Eq. (2.13), the force that is developed at the sample-bar interfaces can be obtained

from three-wave method.

With the condition of equilibrium, the strain rate in one wave analysis can be obtained

from Eq. (2.9) and Eq. (2.12) as:

Assuming isotropic material and a constant cross section of the sample (A) throughout the

loading, the stress in sample ( can be obtained from Eq. (2.15)

14
And integrating Eq. (2.13) the strain in the sample ( ) can be obtained as:

2.4 DYNAMIC TENSILE EXPERIMENT

Generally, dynamic tensile experiments for brittle materials are difficult to perform

compared to the quasistatic experiments. The difficulty is attributed from the brittle

fracture characteristics where propagation of stress wave before and after the fracture is the

typical interest to obtain the dynamic behaviour. Based on the wave propagation theory,

several dynamic experimental techniques have been established and implemented in the

dynamic characterisation of the material. Among those methods, direct tension test

(Asprone et al., 2009; Cadoni, 2010; and Howe et al., 1974), dynamic Brazilian disc test

(Chen et al., 2013a, 2014a, 2014b; Gomez et al., 2001; Grantham et al., 2004; Wang et al.,

2009; Zhang and Zhao, 2013), spalling test (Brara et al., 2001; Cho et al., 2003; and

Schuler et al., 2006), gas gun (Grady and Kipp, 1979; and Wang et al., 2008) test methods

were largely used to obtain the dynamic tensile behaviour of brittle solid. In this work, the

dynamic Brazilian disc test and spalling test are briefly reviewed.

2.4.1 Dynamic Brazilian Disc Test

The dynamic Brazilian disc test is generally conducted by a split Hopkinson pressure bar

facility via keeping the sample in between Input bar and Output bar. Figure 2.5 shows the

schematic of the Brazilian disc test with the split Hopkinson pressure bar testing facility.

The direction of ɛI, ɛR and ɛT follow the same feature as shown in Figure 2.4. Next, using

15
the one-wave, two-wave or three-wave method (Eqs. (2.11), (2.10) and (2.13)

respectively), the maximum force to which the sample fails (Fd) is obtained. After that, the

dynamic tensile strength (σtd) is obtained from the following expression (ASTM, 2008;

Zhou et al., 2012):

Here, Ds is the sample diameter and Hs is the height of the Brazilian disc. In case of quasi

static tensile experiment, the quasistatic tensile stress ( ) has a similar form as Eq. (2.17)

where the Fd is replaced with the quasistatic failure load of the material (ASTM, 2008).

Strain gauges
Striker Bar Input Bar Output Bar
gauges gauges gauges

Sample

Figure 2.5. Schematic of the Brazilian disc test with the split Hopkinson pressure bar

testing facility.

A valid dynamic experiment with the Brazilian disc test must have a diametrical splitting

and the first crack has to be originated from the disc centre (Rocco et al., 2001; and Wang

et al., 2009). Along with this diametrical splitting, the shear failures at the two interface

ends of the sample are observed in the dynamic Brazilian disc test. Figure 2.6 (a) shows

the crack pattern in the Brazilian disc test where the shear and tensile cracks after the

Brazilian disc experiment are shown (Zhang and Zhao, 2013). The distribution of tensile

and compressive strain (Chen et al., 2014a) in the sample during the peak strength is

shown in Figure 2.6(b). The strain distribution in the study (Figure 2.6(b)) was captured by

16
the strain gauges that were attached on different locations of sample surface. Figure 2.6(c)

shows a numerical study that describes the tensile and compressive stress distribution of

the Brazilian disc under dynamic loading (Zhou et al., 2013).

(a) (b)
Y (mm)

Sample input Sample output


bar end bar end

Z (mm)
(c)

Figure 2.6. Dynamic tensile test with Brazilian disc sample, (a) Fracture pattern at high

strain rate (Zhang and Zhao, 2013) (b) Stress distribution (Tension and compression) from

experiment (Chen et al., 2014a) (c) Stress distribution (Tension and compression) from

simulation (Zhou et al., 2013).

17
Experimental study with the Brazilian disc test was done with the split Hopkinson pressure

bar facility to study the dynamic behaviour of brittle solid at high strain rate loading (Chen

et al., 2014a; Chen et al., 2014b; Chen et al., 2013a; Dai and Xia, 2010; Grantham et al.,

2004; Gui et al., 2016; Saksala et al., 2013; Wang et al., 2009, Zhang and Zhao, 2013).

The dynamic Brazilian disc test of brittle solid was reported in term of loading rates (Dai

and Xia, 2010; Gui et al., 2016; Saksala et al., 2013) and strain rates (Chen et al., 2013a,

2014b; Grantham et al., 2004; Wang et al., 2009; Zhang and Zhao, 2013) in literature.

Usefulness of dynamic strength in term of strain rate is more favourable since it represents

the fracture strength characteristics with respect to the deformation rate of the material and

it has a considerable application in engineering purposes. The strain rate determination was

observed to be calculated from three major methods in dynamic Brazilin disc test: a)

measuring strain history using the strain gauges at the sample centre (Chen et al., 2014a;

Wang et al., 2009), b) use of the digital image correlation technique (Chen et al., 2013a;

Grantham et al., 2004; Zhang and Zhao, 2013) and c) the strain rate calculated from the

transmitted bar signal (Chen et al., 2014b; Gomez et al., 2001).

In the first method, (outputting strain from the sample by gauges) the strain gauge is

attached to the sample centre which is also located on the point where the first crack

should appear theoretically (Rocco et al., 2001; Wang et al., 2009). The strain gauge is

attached to the disc centre such a way that strain gauge axis remains perpendicular to the

loading axis. Therefore, this strain gauge axis is the direction to which the tensile strain is

developed in the biaxial strain distribution of the disc. The strain history is captured by the

gauge and the strain rate is obtained from Eq. (2.18). Moreover, throughout analysing the

strain rate history, the strain rate at fracture can be obtained from this method.

18
̇

On the other hand, digital image correlation has been used in determining the strain/strain

rate on the fracture plane by high speed photography. Throughout the failure process, a

number of images are analysed to obtain the deformation, and eventually the strain and

strain rate are calculated from the image analysis. However, it should be noted that a

specialised paint containing particles that sparkles at the presence of light has to be applied

on the sample. A comprehensive description of the method can be found in Berfield et al.

(2007).

In the transmitted-bar-signal method, the value of force peak and the duration of the pulse

to load peak, which is obtained from the output/transmitted bar signal, is used to determine

the average strain rate of the sample. Assuming a constant strain rate deformation across

the loading axis (its direction perpendicular to the loading axis along Hopkinson bar),

average strain rate is estimated from the Eq. (2.19) (Chen et al., 2014b; Gomez et al.,

2001). Here, E is the Young’s modulus, and tm is the duration of loading until the fracture

strength.

2.4.2 Spalling test

The experimental approach to study the dynamic behaviour of brittle material with spalling

phenomena was used for different types of brittle solids (Brara et al., 2001; Cho et al.,

2003; Galvez et al., 2002; Schuler et al., 2006; Tedesco and Ross, 1998; Wu et al., 2005).

The spall experiment to study dynamic tensile behaviour of brittle material consists of a

19
striker hitting an input bar in one end and a sample is in contact with the other end of the

input bar. Figure 2.7 shows the schematic of spalling test with a Hopkinson bar testing

facility. The purpose of this experimental setup is that the striker hitting the input bar

generates a compressive pulse that travels along the bar and eventually transmitted to the

sample as a compressive wave with a peak lower than the compressive strength of the

material. This wave is eventually reflected by the free end on the sample producing a

tensile pulse. Other than using the split Hopkinson pressure bar facility, detonation of

explosive is also used to generate the stress pulse in sample in spall test (Kubota et al.,

2008).

Striker Input bar Specimen

Figure 2.7. Schematic of spalling test with Hopkinson bar testing facility.

There are basically three methods available in the analysing of the spalling test to date,

namely, wave shifting method (Galvez Diaz-Rubio et al., 2002; Tedesco and Ross, 1998;

Wu et al., 2005), ejection velocity method (Brara et al., 2001) and pull back velocity

method (Schuler et al., 2006) in determining the dynamic tensile strength. Among those

methods, the latter two are analysed from the velocity that is measured directly from the

sample, while the first method uses the split Hopkinson pressure bar theory to shift the

waves to reconstruct the stress field along the sample. In the ejection velocity method, the

special interest is to obtain the separation velocity (ejection velocity) of fragmented sample

on the fracture plane. In order to obtain the ejection velocity of fragments, a high speed

camera is required so that the particle velocity on the fragment surface (vej) can be

obtained (Brara et al., 2001). Then the dynamic fracture stress (σtd) is obtained from the

following expression (Brara et al., 2001).

20
Here, ρ is the density and c is the elastic wave speed of the material.

Rear face velocity

vpull

Time

Figure 2.8. Typical demonstration of particle velocity history at the free end of sample in

spall experiment.

The method that was developed by Schuler et al. (2006) made use of pull back velocity

(obtained from rear end of the sample) to determine the dynamic fracture strength. The

method has a similarity to plate impact test where the pullback velocity was used to obtain

the fracture strength. A high speed accelerometer/laser vibration meter should be used at

the rear end of the sample to determine the free vibration at the sample end. The typical

demonstration of the velocity history and the magnitude of pull back velocity at the free

end are shown in Figure 2.8. Using the pullback velocity (vpull) the dynamic tensile stress is

obtained as

21
The average strain rate can be estimated from the displacement velocity curve from the

sample length (L) as

̇ ∫

Here, vv(t) is the velocity history at the rear end of sample at time t, and tm is the time

required to reach the velocity peak.

2.5 TENSILE DAMAGE MODEL

Rate dependent tensile damage models of brittle solid have been a topic of interest during

the last few decades (Fahrenthold, 1991; Grady and Kipp, 1980; Hamdi et al., 2011; Liu

and Katsabanis, 1997; Lu and Xu, 2004; Suaris and Shah, 1984; Taylor et al., 1986; Wang

et al., 2008; Yang et al., 1996; Yazdchi et al., 1996). The difficulties in calibrating the rate

dependent models are discussed in Chapter 1. In fact, a model that uses minimal number of

parameters that can be calibrated directly from experiments with common facilities is

desirable. For instance, several two-parameter/three-parameter damage models, which are

formulated from a power law distribution of volumetric strain against the number of

microcracks, are straightforward in calibration and effective in describing high strain rate

behaviour (Grady and Kipp, 1980; Liu and Katsabanis, 1997; Taylor et al., 1986; Yang et

al., 1996). This power law function had either Weibull distribution for the number of

microcracks (Grady and Kipp, 1980; Taylor et al., 1986) or empirical relation between

microcrack density and tensile volumetric strain (Liu and Katsabanis, 1997; Yang et al.,

1996). A semi-empirical damage model for blast study was developed by Yang et al.,

(1996) where dynamic characteristics, namely, the minimum time requirement to initiate

fracture (Li et al., 1994) and the dilatancy (Throne, 1990) of brittle material were

22
considered in damage modelling. In order to accommodate these two phenomena in brittle

solids, they assumed that damage started to accumulate in the material when its

extensional volumetric strain exceeded this critical strain value, at which the material

reached its static tensile strength. Liu and Katsabanis, (1997) developed a damage model

that could be used for blast analysis of rocks. The microcrack density expression used in

their model had a power law dependence on the tensile volumetric strain. One of the

material constants in their model, which was related to the growth rate of microcracks,

could be determined from the minimum damage value that characterises the onset of

fragmentation, the so-called fragmentation limit, under dynamic loading.

2.6 FRAGMENT SIZE MODELS

The theoretical models, which intended to correlate the fragment size under a constant

strain rate, were suggested to the energy balance framework (Glenn and Chudnovsky,

1986; Grady, 1988, 1982; Rouabhi et al., 2005; Yew and Taylor, 1994 and Zhang et al.,

2006). Concerning stress wave propagation in brittle solids, several numerical models were

developed for fragmentation of brittle solids (Drugan, 2001; Shenoy and Kim, 2003; Yarin

et al., 2000; Zhou et al., 2006). Both finite element and discrete element methods were

used for fragment size study under dynamic loading condition (Levy and Molinari, 2010;

Maiti et al., 2005; Miller et al., 1999; Raghupathy et al., 2006; Vocialta and Molinari,

2015).

The energy models make a balance between external energy (obtained from external

impulse load) and internal energy in predicting the fragment size. Generally, the strain

energy and kinetic energy are considered as the external source of energy which is

balanced with surface energy of the fragments. The surface energy is obtained from the

23
newly created surface area of the fragments. The available fragment size models are

formulated against a constant strain rate deformation, and this relative simple co-relation

makes those models a useful tool in the fragment size assessment in macro level analysis.

However, it should be noted that the deficiency of energy consideration in the

conventional energy-based models is discussed in Chapter 1. A short description of

energy-based fragment size models is provided in the following paragraphs of this section.

Grady (1982) developed a fragment size model in which the fragment size was governed

by the energy balance between kinetic energy and surface energy of the newly created

fragments. The contribution of elastic energy was overlooked in this model and fragments

were assumed to be spherical in size for this model. The fragment diameter (a1) can be

obtained as follows (Grady, 1982).

[ ]
̇

Here, KIc denotes the fracture toughness (mode I) of the solid and ̇ is the numerical

average of strain rates that are developed in three major axes. The governing equations of

thermodynamic were used in the fragment size model by Yew and Taylor (1994) that,

eventually recovered Grady’s (1982) fragment size expression. Glenn and Chudnovsky

(1986) identified that neglecting the significant strain energy underestimated the amount of

dissipating energy required to create the new surface of fragments. They (Glenn and

Chudnovsky, 1986) added the strain energy part in the energy balance and enriched

Grady's (1982) model. Since then, their model has been a reference frame in several

research applications and the fragment size. The diameter of the fragment predicted by

Glenn and Chudnovsky (1986) is given in Eq. (2.24).

24
( ) ( )

Where,

( )

( )
̇

( )
̇

Here, is the bulk wave speed √ of the material and R is the initial radius of the

dilating volume under a dynamic loading. However, the way to determine R under a

dynamic loading was not explicitly described by Glenn and Chudnovsky (1986).

Later, Grady (1988) developed a fragment size prediction model by equating the elastic

strain energy with surface energy of a fragment. They showed that the kinetic energy had

an insignificant quantity (the kinetic energy is about 1/15 times of elastic strain energy) to

create new surface of a fragment and it could be neglected. The communication horizon

concept was developed by Grady (1988) and it was introduced in the formulation that co-

related the spalling time with fragment dimension of brittle material. Since then, several

researches (Hild et al., 2003; Levy and Molinari, 2010; Wagner et al., 1992) had used this

concept (communication horizon) as a reference frame in modelling fragment size with

fracture time of brittle solids. However, the fragmentation model by Grady (1988) and

Glenn and Chudnovsky (1986) had compromised the elastic wave speed ( √ with

25
bulk wave speed ( √ and declared that accuracy in high order were not pursued in

their models. Here, E and K are the Young’s modulus and bulk modulus of the solid,

respectively. The predicted fragment diameter in Grady (1988) had a following expression,


[ ]
̇

Where, ̇ is the volumetric strain rate. The fragment size models that have been described

so far did not consider the internal cracks effect in the fragmentation process. In fact, it

was found that when internal defects were included in the energy balance model, the

predicted fragment sizes were greater than those derived from without any internal defect

(Yatom and Ruppin, 1989).

Fragment size models that included the internal defects in modelling were derived from

continuum damage mechanics (Zhang et al. 2006), and with the combination of continuum

and thermodynamic framework (Rouabhi et al. (2005). An empirical threshold value

(Damage parameter, Df = 0.22) was used as a fragmentation limit on the basis that the peak

strength is obtained at this damage value (Zhang et al., 2006). Assuming the fracture

happens instantaneously at the strength peak, Zhang et al. (2006) modelled the dissipated

elastic strain energy by to create new surface of a fragment under a constant strain

rate. Here, ɛ represents the uniaxial strain, and Df is the damage value at strength peak that

was assumed 0.22 as mentioned earlier. However, the excess of elastic energy [

] was assumed to throw the fragments apart. The fragment size model by Rouabhi et

al. (2005) was formulated in thermodynamic principle coupled with energy model and it

was developed in the theory of continuum mechanics. Their model assumed the fragments

with irregular shapes which eventually improved the surface energy determination.

26
2.7 DYNAMIC SIMULATION WITH BRAZILIAN DISC

In recent years, a number of researches conducted numerical studies on dynamic Brazilian

disc test to study dynamic behaviour of brittle solids at high strain rates. Among the

studies, 2D numerical simulation (FEM) of concrete was conducted by Hughes et al.

(1993) to study the fracture initiation and propagation in the disc under high strain rate

loading. The objective of the study was to observe the effect of rate-dependent fracture

strength on crack initiation and propagation of Brazilian disc. A non-liner material model

for concrete was used in their study where cutoff stress values were used in tension and

compression loading. The cutoff value of tension was obtained from the experimental

observation under an impact velocity of split Hopkinson bar experiment. In fact, a cut-off

tension value, which was observed in experiment under a given impact (strain rate), was

considered as the reference of fracture value under the studied impact, and in subsequent

simulations the cut-off fracture strength was varied. The simulation with actual fracture

strength (obtained from experiment) showed the first crack appeared at sample centre.

However, the fracture time in Brazilian disc was shortened by the reduction of fracture

strength under a given impact, and under such condition the fracture tended to initiate

somewhere between sample centre and incident bar end of the sample.

A 2D discrete element study by Zhou and Hao (2008) was concerned with dynamic tensile

behaviour of concrete with the special attention to aggregate distribution and influence on

interfacial transition zone (ITZ) on concrete. The numerical study was conducted with a

Brazilian disc sample with different strain rates. They found that the crack pattern, fracture

time and dynamic tensile strength were affected by the ITZ material property and

aggregate position in the sample. It was observed that the crack pattern could deviate from

27
the well-defined splitting based on coarse aggregate position, while the ITZ property

controlled the fracture initiation time and fracture strength of the sample.

Zhu and Tang (2006) numerically studied the failure process of Brazilian disc samples at

static and dynamic loading condition. In dynamic condition, the particular attention was to

observe the failure pattern in Brazilian disc when compressive waves of different

amplitudes were applied to the disc sample. They found that, compared to static loading;

more tensile cracks were developed, in addition to major tensile crack along the sample

diameter. The macrocrack evolution for cement mortar was studied numerically by Chen et

al. (2014b). This 2D simulation by a split Hopkinson bar setup could capture both shear

failure near sample-bar interfaces and the tensile cracking of Brazilian disc sample. Their

(Chen et al., 2014a) numerical study showed that the first crack initiated between input bar

and sample centre.

The dynamic Brazilian disc study by Saksala et al. (2013) was conducted in both 2D and

3D analyses with an aim to obtain the material constants for a viscoplastic damage model

(Saksala (2010)). A combined FEM/DEM hybrid method was used by Mahabadi et al.,

(2010) to simulate dynamic behaviour of rocks. Their simulation confirmed the diametrical

splitting of the sample, although the initiation and the propagation pattern of the

macrocrack were not discussed in the study. The objective of the work (Mahabadi et al.,

2010) was to validate the hybrid FEM/DEM method (Munjiza et al., 1995) in prediction of

fracture strength and fracture pattern with the same observed from the experiment.

A 3D-elastic numerical simulation (Li and Wong, 2013) with brittle materials showed that

the stress concentration occurred near the sample bar interfaces (5 mm away), and based

28
on such observation they predicted that first macrocrack in the Brazilian disc test might be

initiated from those stress concentration areas (near sample corners). Also, a numerical

study with hybrid FEM/DEM model (Gui et al., 2016) showed that the macrocrack

initiated near sample-incident bar face and propagated towards the sample-output bar face

of the sample. This prediction (Li and Wong, 2013) and observation (Gui et al., 2016) are

a contrast to the traditional valid Brazilian disc test assumption (Rocco et al., 2001; Wang

et al., 2009) where the crack is supposed to be initiated from the sample centre. In

addition, several numerical studies (Chen et al., 2014a; Hughes et al., 1993; Zhou and Hao,

2008) confirmed the initiation of macrocrack were observed near the sample centre.

2.8 DRUCKER PRAGER MODEL

The Drucker-Prager (D-P) yield criterion is a pressure dependent model to determine

whether the material undergone plastic yielding or failed. A linear strain hardening D-P

model can be described by the yield function,

Where is the pseudo effective stress defined by [ (

) ( ) ], the effective stress, √ , The hydrostatic pressure , d is

cohesion of material defined as * + ; The ; is slope of the

linear yield surface in the P- μDP stress plane, I1 is the first stress invariant, J2 is the second

invariant of deviatory stress, J3 is third invariant of deviatory stress kDP is the ratio between

tensile and compressive triaxial strengths and σc is the yield strength of material in

compression. The value of kDP has a typical range between 0.778 and 1 and the Drucker-

Prager model is not very sensitive to this value (Park et al., 2001). The flow potential

29
( and flow rule for liner Drucker-Prager model can be defined by Eq. (2.30) and Eq.

(2.31), respectively.

Where, , Ψ is the dilation angle at P- stress plane, ̅ is the

equivalent plastic strain and σc is the compressive stress. Figure 2.9shows the yield surface

of P- stress plane.

𝜇𝐷𝑃 𝑝
d

Hardening

Figure 2.9. Yield surface of P- stress plane.

30
2.9 SUMMARY

Dynamic tensile strength of brittle material has been observed to increase significantly (in

contrast to quasi static and intermediate strain rate) at high strain rates. Thus to simulate

such high strain rate fracture in brittle solid, the constitutive models should consider the

strain rate effect explicitly. Several fracture models have postulated the unconventional

experiments which have been difficult to calibrate, whereas some of those available

models have required to undergo several iterations in the way to parameter calibration. The

lack of such special experimental facility may restrict users to utilize the model in practical

application. The Chapter has provided a short review of fragment size distribution models

under high strain rate loading for brittle solids. It has been found that several fragment size

models have utilised the energy balance principle where strain energy contribution to post

strength-peak have been neglected and fragment size has been predicted assuming the

shape to be spherical (Glenn and Chudnovsky, 1986; Grady, 1988, 1982; Zhang et al.,

2006). Eventually, the energy calculation has not been accurate in those fragmentation

models.

Experimental characterisation of brittle solid has been widely used by the split Hopkinson

pressure bar testing facility and using this method strain rate vs fracture strength data can

be obtained. Among the three methods in determining the strain rate as discussed in

section 2.4.1, the use of strain gauge (at the sample centre) has allowed to obtain the

strain/strain rate history of the disc centre. In fact, a complete strain rate history (using

strain gauges) from the sample centre has allowed in obtaining the strain rate at the

fracture of the material. Thus, this method (strain gauge) has not assumed a constant strain

rate deformation and strain rate at the fracture can be explicitly determined.

31
Numerical study with dynamic Brazilian disc test has been used to characterize dynamic

mechanical property of brittle materials. The other scope has become to validate several

constitutive models by comparing the numerical data with the experimental observations.

Some numerical study has found that the crack initiation in Brazilian disc sample occurred

near the sample ends. This observation has contradicted the theoretical aspect of a valid

Brazilian disc test. The numerical simulation of this thesis would explain the macrocrack

initiation in Brazilian disc test under the dynamic loading.

32
CHAPTER 3

MECHANICAL CHARACTERIZATION OF SANDSTONE

3.1 INTRODUCTION

Sandstones are generally quasi-brittle porous materials that are found to be in different

colours depending on chemical composition in geological formation. Great sources of energy

and fresh water have been found to be trapped inside the sandstones which are sometimes

termed as reservoir sandstones. In fact, 60% of energy resources were found within sandstone

formation that are situated below ocean and land (Perrodon and Marshall, 1983). Sometimes,

it is necessary to perform core drilling to reach the hydrocarbon resources or clean source of

water which inevitably requires mechanical characterization of this material.

In this chapter, the mechanical behaviour of sandstone is studied at both static and dynamic

loading conditions. Split Hopkinson Pressure Bar facility was used to study the dynamic

fracture strength characteristic of Brazilian disc specimens at high strain rate loading. Details

of experiment, mechanical behaviour and fracture pattern (observed from experiment) of

sandstone are provided in this chapter. Along with mechanical characterisation, this chapter

has provided a direct observation of a dynamic problem which is the basis of FEM study by

SHPB setup in this thesis.

33
3.2 MATERIAL AND SAMPLE PREPARATION

The sandstone that has been studied here is commercially named was Piles Creek which is

originated from Gosford, Central coast, NSW, Australia. Based on the geological location

and chemical composition, the material (Piles Creek sandstone) is found to be in two colours

namely, orange and cream. A survey on different available types of sandstone reveals that the

Piles Creek sandstone with cream colour has no visible foliation and layer in rock. Since the

effect of rock layers is beyond the scope of the study, a homogeneous material assumption

can be adopted for the Piles creek sandstone with the cream colour appearance. Moreover, a

great reserve of natural gases is found in the central coast regions and both dynamic and static

material properties would help design the suitable drilling operation through the stone if

required. For these reasons, cream Piles Creek is selected for mechanical testing.

Sandstone plates of dimension 400 mm × 400 mm × 50 mm are collected to prepare samples

for this study. High speed water jet cutting facility is used to cut samples with cylinder Ø25

mm × 50 mm from the plates which are obtained from the supplier. The sample having Ø25

mm × 50 mm is used to obtain quasi static compressive strength and Young’s modulus of the

material. The dynamic tensile experiment and quasi static tensile experiments were conducted

using Brazilian disc having approximate dimension Ø25 mm × 10 mm. The discs of specified

dimension are prepared from the cylinders by cutting with a circular rotary saw.

34
3.3 EXPERIMENTAL TECHNIQUE

3.3.1 Quasi-static Testing

A displacement controlled Instron (MTS Criterion, model 43) was used to obtain quasi-static

compressive strength and Young’s modulus from the cylindrical samples. The sample was

placed between the loading plates and a compressive load was applied to the sample. Figure

3.1 shows the test setup for quasi static compression test with a cylinder specimen. Applying

a constant displacement rate to the loading plate, a constant strain rate of 1 × 10-5/s was

provided to the samples. During the tests, the interfaces between loading platens and sample

were lubricated by grease so that friction force could be minimised. Strain of the sample was

calculated from the displacement measured from the LVDT attached to the loading platen.

Figure 3.1. Experimental setup for quasi static compressive test using the Instron machine.

35
Figure 3.2 shows the experimental setup for Brazilian disc test at quasi static loading

condition. The objective of this Brazilian disc test was to obtain the quasistatic tensile

strength of the material. The sample was placed between the loading plates and a

compressive load was applied along the diameter of the sample. A constant strain rate of 1

×10-5/s (applying a constant displacement rate of loading plate) was applied diametrically

until the material fails. In fact, the mentioned strain rate did not present the strain rate in the

direction of tensile loading, and the strain rate that addressed the quasi-static failure in

tension was not determined. Nevertheless, it is plausible that the strain rate at the tensile

direction of the sample would be within the range of quasi-static loading by the mentioned

strain rate as 1 ×10-5/s.

Figure 3.2. Experimental setup for quasi static tension test for Brazilian disc sample.

36
3.3.2 Dynamic Brazilian Disc test with SHPB

Split Hopkinson pressure bar (SHPB) experiment facility has been widely used to study

dynamic tensile behaviour of rocks through Brazilian disc tests. Smaller specimens were

always preferred in Brazilian disc test to achieve dynamic force equilibrium in high loading

rate (Zhou et al., 2012). In addition, smaller diameter Hopkinson bars are sought to minimise

wave dispersion effect (Gama et al., 2001). For these purposes, the sample having a diameter

of 25 mm and an approximate height 10 mm was used for this experiment and, the

Hopkinson bars of this experiment had a diameter of 15 mm. The bars were made from

margarine steel having Young’s modulus, density and Poisson’s ratio as 210 GPa, 7800

kg/m3 and 0.29 respectively. The length of input and output bars are 1.5 m each and strain

gauges were attached at the mid length locations. The schematic of the experimental setup for

dynamic Brazilian disc test is shown in Figure 3.3.

Pulse shaper
Strain gauges
gauges
Output Bar Input Bar Striker Bar
gauges gauges gauges

Sample
1500 mm gauges 1500 mm
160 mm
750 mm 25 mm 750 mm or
120 mm

Figure 3.3. Schematic view of Brazilian Disc test with the split Hopkinson pressure bar

testing facility.

37
The methods to determine the strain/strain rate with Brazilian disc samples were discussed in

section 2.4.1. In this experiment, strain gauges were used to obtain the strain/strain rate for

the sandstone samples. Figure 3.4 is a photo showing the sample orientation with strain

gauges during the Brazilian disc test by the split Hopkinson pressure bars test. During

experiment, the sample was placed between input and output bars keeping the measurement

axis of strain gauges perpendicular to loading axis of Hopkinson bars (see Figure 3.4). The

strain gauges were attached at the centre of two surfaces of the Brazilian disc. The strain of

the sample is determined by the numerical average of two strain gauges data at the end of the

experiment. Before attaching the strain gauges, it was ensured that the disc surface was

smooth, flat and clean enough so that the gauges could be glued properly on the sample.

A pulse shaper of a relative soft material (Copper C11000) was placed between the input and

striker bars to gradually increase the stress pulse, thereby, providing better dynamic

equilibrium for the sample. The pulse shapers were disc in shape with a dimension of Ø6.35

mm × 1.6 mm. A 160-mm striker bar was used in the experiment with two impact velocities

as 11.1 m/s and 15.2 m/s. However, a higher impact velocity revealed that the sample was

fragmented to several pieces and tensile splitting could not be confirmed by this way. In order

to capture a wider range of strain rate, a smaller length of striker as 120 mm was used. The

reason of selecting a smaller length striker in higher strain rate study was to ensure

diametrical splitting without the developing of other major cracks in each of the two

segments (separated segments of the sample due to tensile cracking as discussed in section

2.4.1, Chapter 2) of the samples. Several trial shots with different bars revealed that the 120-

mm striker with 20.4 m/s impact had a fracture plane passing through the centre and a higher

strain rate can be obtained.

38
Figure 3.4. A Brazilian disc sample is placed between the input and output bars keeping the

axis of gauges normal to loading axis of the bars.

3.4 RESULTS AND DISCUSSION

3.4.1 Quasi-static behaviour of sandstone

Figures 3.5 and 3.6 show the stress strain plot and post failure fracture mode of a cylinder

specimen under the uniaxial compressive loading (a strain rate of 1×10-5/s), respectively.

Figure 3.5 shows that the early stage of stress strain response has a flat slope (until strain

0.003) followed by a steeper slope until the ultimate strength of the material. A similarity of

such response in stress strain relation could be found for sandstone (Bagde and Petros, 2005)

where the latter slope was used to determine the Young’s modulus of the sandstone. In the

current study, the latter slopes were also used to determine the Young’s modulus for

39
sandstone. A typical 45˚ shear failure plane was obtained from the quasi static compressive

test with cylinder specimen as shown in Figure 3.6. Table 3.1 summarises the results obtained

from quasi static compression test for cylinder specimens. The experiment revealed that the

average compressive strength and Young’s modulus of the material were 46 MPa and 9.1

GPa, respectively. The density of the material was obtained as 2380 kg/m3.

60

Test 1
50
Test 2

Test 3
40
Axial Stress (MPa)

Test 4

30 Test 5

Test 6

20

10

0
0 0.003 0.006 0.009 0.012 0.015 0.018 0.021
Axial Strain

Figure 3.5. Stress strain curve obtained from sandstone at the strain rate of 10-5 s-1.

40
Figure 3.6. Recovered sample after the quasi static compression test.

41
Table 3.1 Result obtained from quasi static compression experiment of sandstone.

Test No Young’s Failure Average Average


Modulus Strength Young’s Failure
Modulus Strength
(GPa) (MPa)
(GPa) (MPa)

1 9.25 50

2 10.3 45

3 8.4 43
9.1±1 46±4.5
4 7.4 40

5 9.5 48

6 10 52

1600

1400

1200
Test 1
1000 Test 2
Load (N)

800 Test 3
Test 4
600
Test 5
400

200

0
0 200 400 600 800 1000
Time (s)

Figure 3.7. Time load history of Brazilian disc test at quasi static loading condition.

42
Figure 3.8. Recovered sample after the quasi static tension test.

The loading histories of the quasi static Brazilian disc test are shown in Figure 3.7. The peak

load for each experiment corresponds to the failure load for the Brazilian disc. Using the

failure load and Eq. (2.17), the facture stress of the Brazilian disc samples were calculated.

Figure 3.8 shows the recovered sample after this experiment. As expected, a tensile splitting

along the diameter of the sample is observed in Figure 3.8. The summary of the Brazilian disc

test is reported in Table 3.2 which reveals the average fracture strength (quasi static) of the

material to be 2.65 MPa.

43
Table 3.2 Results obtained from quasistatic tension test with Brazilian disc samples.

Test No Sample Sample Failure load Fracture strength Average


diameter height fracture
(N) (MPa)
strength
(mm) (mm)
(MPa)

1 24 9 860 2.5

2 25 9 1110 3.1

3 25 11 1050 2.4
2.64 0.34
4 23 12 1065 2.3

5 25 12 1365 2.9

3.4.2 Dynamic behaviour of sandstone

The signals obtained for three impact velocities of 11.1 m/s, 15.2 m/s and 20.4 m/s were

analysed to obtain the dynamic fracture strength of the material for three strain rates. The

incident (ɛi) and reflected (ɛr) strain wave were obtained from the strain gauge attached to the

input bar, where the strain wave that was transmitted through the sample (ɛt) was captured by

the output bar gauge. Strain histories of these three signals are shown in Figure 3.9 with a

projectile impact of 15.2 m/s. It is observed from the Figure 3.9 that the designed dimensions

of bars, gauge locations and pulse shaper of the experiment were good enough to avoid any

undesirable superimposition of waves. Figures 3.10-3.12 show the fracture mode of samples

under impact velocities of 11.1 m/s, 15.2 m/s and 20.4 m/s, respectively. As expected from

this dynamic test, the discs split diametrically with crushed zones of samples at sample-bar

interfaces. The crushed zone (shear failure) with impact velocity 20.4 m/s (see Figure 3.12) is

observed with higher volume than other two impact velocities (see Figures 3.10 and 3.11). In

44
fact, it was a common trend in dynamic Brazilian disc test that the size of the crushed zones

increases with the increase of strain rate (Chen et al., 2014; Saksala et al., 2013). In spite of

having higher crushing volume (compared to 11.1 m/s and 15.2 m/s impact) for impact

velocity of 20.4 m/s, a tensile splitting was observed nearly one third of sample diameter at

the centre (Figure 3.12). Thus tensile strain and strength at the centre of the disc can be

effectively obtained under 20.4 m/s impact.

0.0015

0.001 Input bar Signal Output bar signal

0.0005
Strain

0
0 0.0002 0.0004 0.0006

Time (s)
-0.0005

-0.001

-0.0015

Figure 3.9. Strain history of steel bars at the impact velocity of 15.2 m/s.

45
Figure 3.10. Recovered Brazilian disc sample for the impact velocity of 11.1 m/s.

Figure 3.11. Recovered Brazilian disc sample for the impact velocity of 15.2 m/s.

46
Figure 3.12. Recovered Brazilian disc sample for the impact velocity of 20.4 m/s.

One of the major issues in SHPB experiment was to ensure the dynamic equilibrium of the

sample before it failed. Figure 3.13 shows the force time history between Sample-Input (S-I)

and Sample-Output (S-O) interfaces under impact velocity of 15.2 m/s. The forces at sample-

input and sample-output interface were determined from the Eq. (2.10) and (2.11) (see

Chapter 2, Section 2.3), respectively. A reasonable match of forces at two ends of the sample

is observed in Figure 3.13 which indicates the dynamic equilibrium is reached in the sample

before it fails.

47
6

5 Sample-Outputbar end
Force (kN)

4 Sample-Inputbar end

0
0 0.03 0.06 0.09 0.12 0.15 0.18
Time (ms)

Figure 3.13. Force history at the two ends of the sample at impact velocity 15.2 m/s.

Figure 3.14(a) shows the strain at the centre of the sample for impact velocity 15.2 m/s. A

numerical average of two strain gauges is considered as a representative strain of the sample.

Strain rate of the sample is obtained from Eq. (3.1) and the plot between time and strain rate

for impact velocity 15.2 m/s is shown in Figure 3.14(b). The strain rate history at the sample

centre for impact velocity 11.1 m/s and 20.4 m/s are shown in Figure 3.15. It is observed

from Figure 3.14(b) that the strain rate shoots up at the transition point as indicated in this

figure. At this transition point, the fracture of the material happened which eventually

raptured the strain gauges and an abnormal increase in strain/strain rate with respect to time

was observed. The material was assumed to be failed at this transition point and the

corresponding strain rate is defined as the fracture strain rate.

48
(a) (b)

Figure 3.14. Strain and strain rate history at the sample centre for impact velocity 15.2 m/s.

(b)

Figure 3.15. Strain rate history at the centre of BD sample at projectile impact (a) 11.1 m/s

(b) 20.4 m/s.

The force histories of the sample-output bar ends (1-wave method as explained in Chapter 2,

section 2.3) were plotted in Figures 3.16-3.18 for impact velocities 11.1 m/s, 15.2 m/s and

20.4 m/s, respectively.

49
The peak force of each test in Figures 3.16-3.18 is the failure load (Fd) which can be used

with Eq. (2.17) to determine the dynamic tensile strength (one-wave method) of the disc.

Table 3.3 shows the strain rate and fracture strength data obtained from dynamic Brazilian

disc test. The fracture strength reported in Table 3.3 is obtained using 3-wave method (as

explained in Chapter 2, section 2.3) of SHPB test theory. It is observed (Table 3.3) that three

strain rates, namely, 24/s, 50/s and 96/s, are achieved from these tests and the average

fracture strengths are 8.9 MPa, 10.5 MPa and 14.4 MPa, respectively.

4
Test 1
Test 2
3 Test 3
Force (kN)

Test 4
Test 5
2

0
0 0.03 0.06 0.09 0.12 0.15
Time (ms)

Figure 3.16. Force history of the sample-output bar end at impact velocity 11.1 m/s.

50
6

5
Test 6
4 Test 7
Force (kN)

Test 8
3 Test 9
Test 10
2

0
0 0.03 0.06 0.09 0.12 0.15
Time (ms)

Figure 3.17. Force history of the sample-output bar end at impact velocity 15.2 m/s.

6 Test 11
Test 12
Force (kN)

Test 13
4
Test 14

0
0 0.02 0.04 0.06 0.08 0.1

Time (ms)

Figure 3.18. Force history of the sample-output bar end at impact velocity 20.4 m/s.

51
A plot between strain rate and fracture strength in log10 σ-log10 ̇ space is shown in Figure

3.19. It was suggested that brittle solids show a liner relation between strain rate and fracture

strength when it is plotted in log-log space (Grady and Kipp, 1980; Taylor et al., 1986). The

best fit linear regression for the sandstone (see Figure 3.19) gives the following expression:

In Eq. (3.2), the unit of stress and strain rate are MPa and s-1, respectively. It can be seen from

the Eq. (3.2) that the dynamic fracture strength is proportional to 1/3 power of strain rate.

Similar characteristic to fracture strength and strain rate relation has been observed for

different types of rocks (Cho et al., 2003; Kubota et al., 2008). Therefore, the strength and

strain rate relationship obtained from the dynamic Brazilian disc test agreed well with those

in literature. On the other hand, the tensile dynamic increase factor (TDIF) was reported with

three different strain rate regimes namely low ( ̇ 10-4 s-1), intermediate (10-4 s-1 ̇ 1 s-1)

and high strain rates ( ̇ s-1) (Zhou and Hao, 2008). In this study, the dynamic tensile

experiment is concentrated in the high strain rate regime. Therefore, the plot between TDIF

(Figure 3.20) and strain rate shows the following general expression in high strain rate regime

for Piles Creek sandstone.

̇ ̇

52
Table 3.3. The summary of dynamic Brazilian disc test with different strain rates.

Sample Average Average


Impact Strain Fracture
Test diameter Sample Strain fracture
velocity rate strength
No height rate Strength
(mm)
(ms-1) (s-1) (MPa)
(mm) (s-1) (MPa)

1 24 11 15 8.0

2 24 10 17 8.0

3 24 11 11.1 25 8.25 24.4 8.5 8.9 1.35

4 25 11 35 9.0

5 23 9 30 11.2

6 25 10 45 11.5

7 25 10 60 10.25

8 23 10 15.2 65 10.25 50.4 16 10.5 1

9 24 10 25 10.0

10 23 11 57 12.0

11 25 11 115 13.75

12 24 10 90 14.25
20.4 96.3 13 14.4 2
13 23 11 85 11.5

14 24 10 95 17.8

53
1.3

𝑙𝑜𝑔 𝜎𝑡𝑑 𝑙𝑜𝑔 ̇


1.2
log10 σtd (MPa)

1.1

1
Experiment data

Average data
0.9
Best Fit

0.8
0.9 1.1 1.3 1.5 1.7 1.9 2.1 2.3
log10 ̇ (s-1)

Figure 3.19. A plot between dynamic fracture strength and strain rate in log-log space.

4
TDIF

3
TDIF=3.35 log10 ̇ -1.482
2

0
1 1.2 1.4 1.6 1.8 2 2.2

log10 ̇ (s-1)

Figure 3.20. A plot between tensile dynamic increase factor (TDIF) with strain rate.

54
3.5 SUMMARY

In this Chapter, mechanical properties of Piles Creek cream sandstone at quasi static and

dynamic condition have been studied. The quasi static experiment has been conducted to

obtain the strength of material at compressive and tensile loading. The Young’s modulus,

compressive strength and tensile strength of sandstone have been obtained as 9.1 GPa, 46

MPa and 2.65 MPa, respectively. The dynamic Brazilian disc test has been conducted using

split Hopkinson Pressure bars. A diametrical splitting of sample has been observed from the

experiment which confirms the tensile failure at fracture processing zone of the sample. The

samples have been observed to reach dynamic force equilibrium before fracture has occurred.

The obtained dynamic tensile strengths at three different strain rates of 24/s, 50/s and 96/s

have been 8.9 MPa, 10.5 MPa and 14.4 MPa, respectively. Strain rate at fracture strength has

been obtained directly from the strain gauges attached to the samples. The dynamic tensile

experiment has showed that the fracture strength of sandstone is proportional to one third

power of strain rate. One of the purposes of these experiments has been to study the strain

rate dependent fracture behaviour of sandstone. Later, these experimental observations will

be used to validate the simulations of dynamic behaviour of sandstone using the proposed

tensile damage model.

55
CHAPTER 4

CONSTITUTIVE MODELLING: DAMAGE MODEL AND

FRAGMENT SIZE MODEL

4.1 INTRODUCTION

A three-parameter tensile damage model and a fragment size prediction model are developed

in this chapter to study the dynamic behaviour of brittle materials at high strain rate impact

loading. The model is developed in continuum mechanics approach where the growth of

microcracks (microcrack density) under the applied loading has been a function of tensile

volumetric strain. The damage parameter used here is dependent on microrack density of the

brittle material which was developed by Budiansky & O’Connell, (1975). Taking microcrack

induced damage and energy into consideration, the size of a fragment is formulated as a

function of the strain rate. The conventional energy balance approach understates the surface

energy and elastic energy (as discussed in Chapter 1) to obtain the size of a fragment. The

proposed model addresses those energy issues to better predict the size of a fragment. The

scope of this chapter is to develop the constitutive relations for the models and to show the

way the model parameters are determined.

4.2. CONSTITUTIVE MODELING FOR DYNAMIC TENSION

A microcrack induced damage model is proposed in this thesis, which can be implemented to

study the dynamic behaviour of brittle solid under high strain rate tension. In this proposed

56
model, the initial brittle material is assumed to be isotropic, continuous and homogeneous

with the activation of microcrack growth being solely determined by the extensional (tensile)

volumetric strain of the material. It is also assumed that the pre-existing microcracks are

activated to grow only when a critical extensional volumetric strain is exceeded. The model

does not explicitly treat individual microcracks or their orientations. It rather considers the

effect of microcrack growth on the degradation of material stiffness. Once the tensile damage

is activated, it cannot heal back. In other words, to further increase the damage after

unloading of partially damaged material, the current extensional volumetric strain has to be

higher than the historical maximum. Following the work of Yang (1996), the microcrack

density growth rate ( ̇ ) at time t is correlated to the tensile volumetric strain as follows:

( ̇) 〈 〉

where α refers to the number of microcracks that increases with time, β the strain-rate

sensitivity parameter, the volumetric strain at time t, and the volumetric strain of the

material at static tensile strength (σst). In order to ascertain the dynamic behaviour, the

parameters (α and β) are calibrated using a dynamic tensile experiment. The bracket term

〈 〉 in Eq. (4.1) means that the microcrack density will grow only when

. Eq. (4.1) has a similar form to Paris Law in metal damage (Pugno et al., 2006). If the static

tensile strength σst and bulk modulus K of the material are known, then can be calculated

as,

Considering constant strain rate deformation, the microcrack density growth rate can be

calculated by,

( ̇) ̇

57
where ̇ and are the volumetric strain rate and the time required to reach the critical

strain under a given constant strain rate, respectively. Integrating Eq. (4.3) yields,

where is the initial microcrack density of intact material. The scalar damage parameter is

scaled to zero with this initial microcrack density in intact material and thus it can be

neglected. As a result, Eq. (4.4) can be simplified as

Rearranging Eq. (4.5), we obtain,

Eqs. (4.5) and Eq. (4.6) give the microcrack density as a function of time and volumetric

strain, respectively, at a constant strain rate of ̇ .

In our model, material damage is defined through a scalar damage parameter D, which can

also be obtained from a constitutive relation of degraded bulk modulus under external

loading. Following the work of Budiansky and O’Connell (1975), the effective bulk modulus

of a damaged solid with a random distribution of microcrack density, Cd, can be expressed in

Eq. (4.7). A crack density expression as a function of Poisson’s ratio is given in Eq. (4.8)

(Budiansky and O’Connell, 1975)

̅
̅ ( )
̅

̅ ̅
̅ ̅

where K and v are the original bulk modulus and Poisson’s ratio of the undamaged material,

respectively. ̅ and ̅ are the effective bulk modulus and effective Poisson’s ratio of the

58
damaged material, respectively. Eq. (4.5) and Eq. (4.8) are two independent expressions to

obtain the microcrack density of the damaged material.

For more efficient numerical implementation, the cubic function of Eq. (4.8) can be further

simplified into a linear analytic function, with acceptable error, as follows (Taylor et al.

1986),

̅ (4.9)

Using the scalar damage parameter D, Eq. (4.7) can be simplified as,

̅ (4.10)

Inspection of Eqs. (4.7) and (4.10) reveals that the scalar damage parameter D can be

expressed by,

̅ (4.11)

̅
̅
̅

This degradation of material elastic properties is reflected by the increase of the scalar

damage parameter D, which can be calculated from the effective crack density of material Cd.

The range of the damage parameter D is between 0 (intact material) and 1 (completely failed
̅
material). It can be observed from Eqs. (4.9)-(4.12) that when . This critical

value of microcrack density ( = 9/16) represents the fracture (D = 1) when the tension

stiffness of material vanishes completely.

59
4.3. FRAGMENT SIZE MODEL

A strain-rate dependent fragment size model is developed based on energy balance for a

brittle material under dynamic tensile loading. The conventional energy balance neglects the

energy contribution during the post strength peak softening regime and the surface area is

underestimated by assuming sphere fragment shape (Grady, 1988, 1982; Rouabhi et al., 2005;

Yew and Taylor, 1994; and Zhang et al., 2006). In order to improve the energy balance

between elastic strain energy and surface energy, the fragment is assumed to have a prolate

spheroid shape in this model. Although a prolate spheroid shape is not an actual shape when

compared to irregular fragment shape observed in practical scenarios, this type of shape will

provide a more realistic surface area of a fragment under a given strain rate. The surface area

of a fragment is formulated by a prolate spheroid fragment assumption, in which the surface

area is a function of strain rate. In this way, the model not only considers the complete strain

energy during fragmentation, but also rules out the underestimated surface energy from the

spherical fragment assumption.

The constitutive relation for stress and strain, which is obtained using the rate-dependent

tensile damage model, will be utilised to calculate the total strain energy in the proposed

fragment size distribution model. Figure 4.1 shows a typical mean tension (P) vs volumetric

strain curve obtained from the rate-dependent tensile damage model derived in section 4.2.

Here, the volumetric strain at peak pressure and volumetric strain at complete separation of a

body due to fracture are termed as ɛP and ɛf, respectively. As can be seen from Figure 4.1, the

material loads elastically until the critical volumetric strain (ɛcr) is reached. Next, the damage

is activated and gradual hardening of the material leads to the dynamic fracture strength (Pd)

at volumetric strain ɛP. Eventually, a post-peak softening leads the material to create a

60
complete fracture at the volumetric strain of ɛf, where the damage parameter, D, reaches value

of 1.

𝑃𝑑 Linear softening modulus


Mean Tension (P)

assumption between ɛP and ɛf

𝑃𝑠𝑡

𝑐𝑟 𝑃 𝑓

Volumetric strain 𝑣

Figure 4.1. Typical mean tension vs volumetric strain curve obtained using the proposed

damage model.

4.3.1 Determination of strain energy

The strain energy determination of this fragment size model considers the complete strain

energy (until fracture) of the material under a constant strain rate. However, the strain energy

determination is different from the other energy models (Glenn and Chudnovsky, 1986;

Grady, 1988, 1982; Zhang et al., 2006) where energy until the strength peak is considered in

those models. The total strain energy density (ΠT) is the area under the pressure-volumetric

strain curve as shown in Figure 4.1. Total strain energy density can be written as,

61
Here, the elastic strain energy density between zero strain and ɛcr is Πe, the strain energy

density between strain ɛcr and ɛP is Πh, and Πs denotes the strain energy density from strain ɛP

to ɛf. Although the pressure-volumetric strain relation between ɛP and ɛf is nonlinear, a linear

softening modulus is assumed to calculate Πs, as shown in Figure 4.1. The strain energy Πe is

as follows,

An expression of Πh is as follows,

∫ ∫

Substituting Eq. (4.11) into Eq. (4.15) we have:

̅
∫ ( )

Here, Pst denotes the mean tension at static tensile strength of the material. Several studies

have indicated that the material reaches its peak tensile stress at a damage value of 0.2 (Grady

and Kipp, 1980; Hamdi et al., 2011; Taylor et al., 1986; and Zhang et al., 2006). The model is

insignificantly sensitive to the Poisson’s ratio function, ̅ , and this effect will be

discussed numerically in Section 4.4.1. Thus, the ̅ shown is Eq. (4.16) can be defined as

an effective Poisson’s ratio function at a damage value of 0.2 and it is denoted as { ̅ } .

However, the value of { ̅ } can be obtained from Eqs. (4.9), (4.11) and (4.12). Thus

replacing ̅ with { ̅ } in Eq. (4.16) and using Eq. (4.6) and Eq. (4.16) we obtain

{ ̅ }
* +
̇

Finally, the strain energy under post-strength peak softening (Πs) can be obtained as:

62
( )

Here Pd represents the mean tension (pressure) at a given strain rate. Since, the stress-strain

relation changes with the change of strain rate, the parameters Pd, ɛp and ɛf also change

according to change of strain rate.

4.3.2 Determination of surface energy

If the energy required to create a new surface per unit volume is labelled by 𝛤 then,

where is the specific surface area per unit volume of a nominal fragment. The fracture

surface energy per unit area ( ) in a brittle fragment is obtained as,

where c is the elastic wave speed, ρ the density and KIc the fracture toughness (mode I) of the

brittle material. Now, for irregular shaped fragments, the specific surface area per unit

volume can be defined as,

Where, Sf and Vf are the surface area and volume obtained from a fragment. The fragment is

assumed to have a prolate spheroid shape in this model. It should be noted that the prolate

spheroid assumption does not attempt to obtain the actual shape of the fragment, rather the

objective is to improve the surface area/energy by using more realistic shape assumption in

the modelling. When the fragment has a spherical shape with diameter a1, As can be obtained

63
as 6/a1 from Eq. (4.21). Considering a prolate fragment shape, the specific surface area

obtained using Eq. (4.21) is as follows,

where the expression of n is as follows,



[ ]

Here a and b are the major axis length and minor axis length of a prolate spheroid fragment. n

is a constant and is a function of fragment aspect ratio b/a.

Combining Eqs. (4.19), (4.20) and (4.22), we can have the surface energy density as follows:

4.3.3 Determination of fragment size

To obtain the average fragment size, the concept of communication horizon which provides a

relationship between fragment size and fracture time is frequently used (Grady, 1988; Levy

and Molinari, 2010). This concept suggests that when a body undergoes a failure due to an

internal spall at time , it is necessary for every fragment with a radius c to fail

independently. Figure 4.2(a) and Figure 4.2(b) demonstrate the hypothesis of a spherical

fragment by Grady (1988) and the corresponding time-mean tension curve in the

conventional communication horizon concept, respectively. Here, the fragment is assumed

with a constant shape as a sphere in this conventional approach and a1 represents the diameter

64
of a fragment. represents the time required to load the material until the fracture strength.

By neglecting the post strength peak softening energy, Grady (1988) assumed that, both peak

stress and fracture occur at time as shown in Figure 4.2. Such assumption is suitable for

materials that shows extreme brittle characteristics, and less suitable in quasi-brittle materials

like rocks and concrete. Thus, a robust model is necessary to capture such post- peak stress

characteristic in quasi-brittle and brittle material, to improve the prediction of fragment size.

a1= 2c𝑡𝑓
(a)

P
(b)

t
𝑡𝑓

Figure 4.2. Fragment formation hypothesis by Grady (1988). (a) Spherical fragment, (b)

Time-mean tension plot.

It should be noted that in a real scenario, the fragments follow irregular shapes as compared

to the standard spherical shape suggested by the communication horizon concept. For the

purpose of simplicity, the complicated irregular fragment shape is not explicitly considered as

the underlying mechanisms of this irregularity are still unknown.

65
Instead two variables are used in this model which are the length, a, and the aspect ratio (b/a)

of a fragment in order to identify a more realistic surface area of a fragment. The first

variable in this model is the length of the fragment which is determined from the

communication horizon concept as shown in Eq. (4.25). In this model, the time that is needed

for the complete separation of a body (tf) is implemented to determine the maximum

dimension (length of a prolate spheroid) of the fragment. The maximum dimension of the

fragment is assumed in this determination (Eq. (4.25)) as it provides the maximum viable

dimension of a fragment in the communication horizon concept. In other words, the

dimension of a fragment under a constant strain rate cannot be greater than 2ctf . To better

explain the fragment formation assumption of this model, the evolution of a typical fragment

and its corresponding time-mean tension curves are presented in Figure 3(a) and Figure 3(b),

respectively.

The other variable is the aspect ratio (b/a) of the prolate spheroid fragment which is a

function of the strain rate. The physical meaning of this co-relation is that, the strain rate

defines the surface area of the fragment once the length of the fragment is determined from

the communication horizon concept. The co-relation between the aspect ratio and strain rate

is monitored with a constitutive relation in damage model.

Balancing the strain energy ( ) with surface energy (𝛤) we obtain a as:

Where, the kinetic energy is neglected since it has an insignificant contribution to

fragmentation (Grady, 1988; Rouabhi et al., 2005; Zhang et al., 2006).

66
Using Eqs. (4.25) and (4.26) the expression for fracture time can be obtained as,

(a) a=2c tf
Fracture

P
(b)

t
tf

Figure 4.3. Fragment formation hypothesis in current model (a) Prolate spheroid fragment,

(b) Time-mean tension response.

Eq. (4.23) shows an expression for the constant n as a function of the fragment aspect ratio.

Now it is required to find another independent expression for n as a function of the strain rate

to correlate fragment aspect ratio with the strain rate. Another expression of n is derived from

the condition when fracture (Damage parameter, D=1) under a constant strain rate is

predicted in the damage model. Eq. (4.5) has been used for this purpose. When the material is

loaded until the fracture time tf, the microcrack density has a threshold value of 9/16.

Substituting Cd = 9/16 and t = tf from Eq. (4.27) into Eq. (4.5) we have the following

expression for n as a function of the strain rate,

67
[, - ]
̇

Once the value of n is obtained from Eq. (4.28) for a constant strain rate, the aspect ratio is

obtained using Eq. (4.23). In this way, the aspect ratio is modelled for a constant strain rate.

When the aspect ratio of the fragment is 1, Eq. (4.23) shows that n tends to a value of 6.

Thus, n has no physical meaning if its value is less than 6.

Once the material parameters are determined and ΠT is calculated, Eq. (4.28) is used to

determine the value of n under a given constant strain rate. However, it can be observed from

Eq. (4.28) that the value of n reduces with the increase of the strain rate. The length of a

fragment, a, can be calculated from Eq. (4.26). Next, the aspect ratio and eventually the

fragment width, b, of a fragment can be determined by using Eq. (4.23). Finally, the

equivalent diameter ( ) of the irregularly shaped fragment can be obtained as follows,

It should be noted that the equivalent diameter is obtained by equating the fragment volume

of a prolate spheroid with a same volume of sphere having a diameter .

4.4. DETERMINATION OF MODEL PARAMETERS

4.4.1 Model parameter α and β

To obtain the values for the model parameters α and β, we first use the following relations of

the mean tensile stress P as a function of the volumetric strain,

68
In order to determine the model parameters, a straight forward approach, which is replacing

̅ with the constant (See Eq. (4.11), is considered here. The reason behind this

assumption is to find a simplified form in the formula with an acceptable error which can be

co-related by a dynamic experiment for the purpose of calibration. In fact, several studies

have indicated that the material reaches its peak tensile stress at a damage value of 0.2 (Grady

and Kipp, 1980; Hamdi et al., 2011; Taylor et al., 1986; Zhang et al., 2006). It can be

obtained from Eq. (4.12) that has a numeric value of 1.6 with original Poisson’s ratio of

0.2. In contrast, ̅ has a numeric value of 1.48 at D = 0.2, which can be easily obtained by

using Eqs. (4.9), (4.11) and (4.12). Thus it can be observed that the function ̅ will vary

between 1.6 and 1.48 to reach the fracture strength for a typical value of Poisson’s ratio 0.2.

However, the calibration process is designed such a way that the assumption of keeping

constant influences to the model parameter α only. In fact, it was observed that the model

parameter which defines the number of microcrack or microcrack growth rate (in current

model it is termed as α) is not very sensitive to fracture strength, although the other model

constant (for current model, β) is highly influenced to strength in term of sensitivity of its

value (Grady and Kipp, 1980; Liu and Katsabanis, 1997; Taylor et al., 1986). Taylor et al.

(1986) found that the fracture strength was affected by 5 % with the change of this parameter

(the parameter that defined the number of microcracks in unit volume) by 25 %. The

numerical implementation of the model in Chapter 5 shows the fracture strength is changed

by only 6 % with change in α by 25 %. By using the expression of damage from Eq. (4.11)

and replacing ̅ with the constant , we obtain:

( )

69
Here, Pst represents the mean tension of the material at quasi-static loading condition. By

using the crack density expression in Eq. (4.5) for a material under constant strain rate

loading, we can re-write Eq. (4.32) as:

* ̇ + ̇

Since is a constant for any given strain rate ̇ , by differentiating Eq. (4.33) with respect

to time we obtain,

̇ ̇

The time required to reach dynamic fracture strength can be obtained by setting Eq. (4.34) to

zero and can be expressed as follows,

[ ] [ ]
̇

By substituting Eq. (4.35) into Eq. (4.33) we obtain:

̇
( ) ( )

Rearranging Eq. (4.36) we obtain:

̇
[ ]

Applying logarithm in Eq. (4.36) we have:

̇ [ ( ) ( ) ]

70
Eq. (4.38) represents a straight line in log ̇ -log(Pd-Pst) space. This suggests that the value of

model parameter β could be determined by the slope from a linear line between logarithm of

strain rate ( ̇ ) and logarithm of tensile strength ( ) plot. Here, and ̇ are the

uniaxial dynamic tensile strength and uniaxial strain rate of the material, respectively.

Conventional experimental facilities such as split Hopkinson pressure bar, plate impact test or

direct tension test can be performed to extract such data for a brittle material. A quasi static

tensile experiment is required to determine σst. Once β is determined, the other parameter, α,

can be determined from a data point on log-log space and putting those values in Eq. (4.37).

4.4.2 Determination of and

The two parameters and are required to determine the strain energy associated under a

constant strain rate loading. The volumetric strain at peak stress can be obtained by

multiplying the time obtained from Eq. (4.35) with strain rate ̇ . Thus the expression for

is as follows,

̇
* +

Volumetric strain at fracture (D = 1) can be obtained from Eq. (4.6) by setting the boundary

condition: when Cd = 9/16. However, it is shown in Section 4.2 that the microcrack

density has a critical value as 9/16 at the damage value of 1. Thus the expression for is

given below:

̇
[ ]

71
Once the model parameters α, β and are determined, the volumetric strain at peak stress

and volumetric strain at facture can be easily determined from Eq. (4.39) and Eq. (4.40),

respectively. It should be noted that Pd at different strain rate can be obtained from Eq. (4.36)

which is required for determining in Eq. (4.18).

4.5 SUMMARY

In the frame of continuum damage mechanics, a rate dependent tensile damage model has

been developed to predict fracture behaviour of brittle solid at high strain rate loading.

Considering the effect of damage into account, a method to determine the size of a fragment

under a constant strain rate has been developed. Both the strain energy and surface energy of

the fragment has been improved in fragment size prediction model, and thus a more accurate

energy balance have yielded a more accurate determination in fragment size. The model has

been easy and straight forward in calibration which requires conventional experimental

facility to determine the model parameters. The calibration of the models can be performed

by dynamic tensile experiments such as spalling test, Brazilian disc and direct tension

experiments. The quasistatic tensile strength can be obtained by Brazilian disc test and direct

tension test at quasi static loading condition.

72
CHAPTER 5

CALIBRATION AND UTILISATION OF THE MODEL

5.1 INTRODUCTION

This chapter is concerned with the numerical implementation of the proposed model to

predict the rate dependent fracture and fragment size of the brittle solid under impact loading.

The damage model was implemented in Abaqus/explicit via a user defined material

subroutine. A spalling experiment with concrete was conducted by Schuler et al. (2006)

providing strain rate, fracture strength and fragment sizes data that were observed from the

experiment. This experiment is numerically studied in this chapter with the proposed damage

model. The numerical study of this chapter shows the damage contour, fracture stress and

strain rate prediction by the damage model and compares the same with the experiment.

Thus, the comparison of experimental observations with numerical predictions would show

the applicability of the model in real case scenarios.

5.2. IMPLEMENTATION OF DAMAGE MODEL TO SUBROUTINE

The tensile damage model is activated only when the volumetric strain is greater than zero. In

other loading conditions, the pressure dependent Drucker-Prager model is used for dynamic

simulation. The volumetric strain are calculated from the sum of true strains at three principal

axes. Once the material parameters α, β and ɛcr are determined from experimental data,

73
microcrack density rate is determined using Eq. (4.1). Eventually, the effective micro crack

density is calculated to get effective Poisson’s ratio and damage using Eq. (4.9) and Eq.

(4.11), respectively. Then, the stress tensor for tensile loading is calculated using,

̅ ̅

where is the kronecker delta, tr(ɛ) is the trace function of strain tensor ɛ, ̅ is the effective

shear modulus of the damaged material, which can be calculated from the effective bulk

modulus ( ̅ ) (see Eq. (4.10)) and effective Poisson’s ratio ( ̅ ) as follows,

̅ ̅
̅
̅

̅ is the effective lame parameter, defined by,

̅ ̅ ̅

Damage evolution is an irreversible process and unloading process is assumed to be elastic.

Damage can only accumulate further when the historical maximum volumetric strain for that

point has been exceeded during reloading. In this simulation, an elasto-plastic Drucker-Prager

(DP) model is used to describe the material response in the compression regime. A brief

review of the Drucker-Prager model is shown in Chapter 2, Section 2.8. A value of 40º for

the slope of linear yield surface, , was used in this study as it is a typical value observed for

different brittle materials (Li et al., 2009; Park et al., 2001).

The rate dependent tensile damage model and pressure dependent Drucker-Prager model

were implemented in Abaqus/Explicit via a user-defined material subroutine (ABAQUS 6.11,

2011). The combined Drucker-prager model and tensile damage model were then used to

simulate the dynamic behavior of brittle solid at high strain-rate loading. Figure 5.1 shows the

flow chart of the damage model used in the nuumerical study.

74
START

Y
ɛv > 0 & Δɛv < 0

N
Y
ɛv < 0 & Δɛv > 0

N
N
Drucker ɛv > 0
Prager
Model Y
N
ɛv > (ɛv)max 𝐶𝑑 =0

Y
Y N
(ɛv)i-1> (ɛv)cr & (ɛv)i < (ɛv)cr

(ɛv)eff = [(ɛv)i-1+ ɛcr]/2 (ɛv)eff = [(ɛv)i-1+ (ɛv)i]/2

𝛽
𝐶𝑑 𝛼 𝑣 𝑒𝑓𝑓 𝑐𝑟

𝐶𝑑 𝑖 𝐶𝑑 𝑖−1 𝐶𝑑 × 𝛥𝑡

9 N
𝐶𝑑 𝑖 <16

D=1
Y
6
𝑣̅ 𝑣 𝐶𝑑 𝑖
9

2 3 4
1

75
1 2 3 4

𝑣̅ 2
𝑓
𝑣̅

6
𝐷 𝑓 𝐶𝑑 𝑖
9

̅
𝐾 𝐷 𝐾

̅
𝐾 𝑣̅
𝐺̅
𝑣̅ 𝜆 𝐾 𝐺

𝜆̅ ̅
𝐾 𝐺̅ 𝜎𝑖𝑗 𝛿𝑖𝑗 𝜆 𝑡𝑟

𝐺 𝑖𝑗

𝜎𝑖𝑗 𝛿𝑖𝑗 𝜆̅ 𝑡𝑟 𝐺̅ 𝑖𝑗

END

Figure 5.1. Flowchart for the damage model for the numerical implementation.

5.3 SIMULATION OF DYNAMIC RESPONSE FOR OIL SHALE

A dynamic tensile experiment with 30 GPT oil shale was conducted by Grady and Kipp

(1980) and the uniaxial strain rate and uniaxial fracture strength (σtd) of the material data

were reported. From Taylor et al. (1986), the mechanical properties of 30 GPT oil shale was

be obtained as: density  = 2270 kg/m3, Young’s modulus E = 10.8 GPa and Poisson’s ratio 

= 0.2. The static tensile strength (σst) of oil shale was 5 MPa (Zhang et al., 2004). Figure 5.2

76
shows this plot in the log10 - log10 (σtd - σst) space (Grady and Kipp (1980)). Using these

material properties, the model parameters α and β were determined (Chapter 4, Section 4.4.1)

as α = 1.1 × 1011 s-1m-3 and β = 2.5, respectively.

In order to observe the response of the model at different strain rates, a numerical simulation

with cylinder was simulated and the results are discussed in this section. In the simulation,

one end of the oil shale cylinder was subjected to a unixial tensile loading while the other end

is on roller support. The diameter and height of the cylinder were 10 mm and 20 mm,

respectively. Eight nodes brick elements (C3D8R) were used throughout the simulation. The

schematic drawing of the loading and boundary condition is shown in Figure 5.3 (a). The

FEM meshing of the cylinder can be found in Figure 5.3 (b). Figure 5.4 shows the mesh

convergence study with element size 1.5 × 10-3 mm, 1 × 10-3 mm and 5 × 10-4 mm for the

simulation with 10 s-1 strain rate. It is observed from the Figure 5.4 that the element having

mesh smaller than 1 × 10-3 mm would incur very negligible difference to the fracture

strength. Therefore, the elements with 1 × 10-3 mm were used in this cylinder simulation

providing strain rates of 10 s-1, 100s-1 and 1000 s-1. Figure 5.5 shows the stress strain relations

of oilshale at strain rates of 10 s-1, 100 s-1 and 1000 s-1 obtained using the proposed rate-

dependent tensile damage model. It is observed from the Figure 5.5 that the predicted fracture

strengths for strain rates 10 s-1, 100 s-1 and 1000 s-1 are 16 MPa, 27 MPa and 57 MPa,

respectively. In contrast, Figure 5.2 shows the experimental values of fracture strengths are

18 MPa, 29 MPa and 55 MPa for those strrain rates, respectively. Hence, the model

prediction to fracture strength matched well with the experiment.

77
2.2

Uniaxial stress log10 (σtd-σst) (MPa) 2

1.8

1.6

1.4

1.2 log10 (σtd-σst) = 0.285 log10 ɛ + 0.8324

0.8
0 1 2 3 4 5

log10 (s-1)
Figure 5.2. Strain rate vs fracture stress for oil shale (Grady and Kipp, 1980).

20 mm
Loading

 10 mm

(a)

78
(b)

Figure 5.3. Numerical simulation of oil shale under uniaxial loading (a) Schematic with

loading and boundary condition (b) FEM meshing of the cylinder

18
16
14
Tensile stress σtd (MPa)

12
10
8 1.5E-3 mm
6 1E-3 mm
4 5E-4 mm

2
0
0 0.001 0.002 0.003 0.004
Axial Strain ɛ

Figure 5.4 Mesh convergence study for oil shale cylinder for strain rate of 10 s-1.

The evolution of damage and axial stress are shown in Figure 5.6. The effects of the model

parameters α and β on the stress-strain curves of oil shale under tensile strain rate of 10 s-1 are

shown in Figure 5.7 and Figure 5.8, respectively. It can be seen from Figure 5.7 that fracture

79
strength increases with the decrease of α. On the other hand, the fracture strength increases

with the increase of the rate sensitive parameter β (Figure 5.8).

60

50

1000/s
Axial Stress (MPa)

40

30
10/s
20 100/s
100/s
1000/s
10 10/s

0
0 0.002 0.004 0.006 0.008
Axial Strain

Figure 5.5. Axial stress-strain curves of oil shale under different uniaxial tensile strain rates.

80
1
16
Axial stress σtd (MPa) 14 0.8
12

10 0.6

Damage
8
0.4
6

4
0.2
2

0 0
0 0.00005 0.0001 0.00015 0.0002 0.00025

Time (s)

Figure 5.6. Uniaxial stress and damage histories for oil shale at strain rate of 10 s-1.

30

25
α = 1.1 ×1010 m-3s-1
Axial stress σtd (MPa)

20

15
α = 1.1 ×1011 m-3s-1
10

α = 1.1 ×1012 m-3s-1


5

0
0 0.001 0.002 0.003 0.004 0.005 0.006
Axial strain ɛ

Figure 5.7. The effect of microcrack growth rate (α) on the stress-strain curves of oil shale

under tensile strain rate 10 s-1.

81
25

β = 2.75
Axial stress σtd (MPa) 20

15
β = 2.5

10
β = 2.25
5

0
0 0.001 0.002 0.003 0.004 0.005

Axial strain ɛ

Figure 5.8 The effect of strain rate sensitivity parameter β on the stress-strain curves of oil

shale under tensile strain rate 10 s-1.

5.4 SIMULATION OF DYNAMIC RESPONSE FOR CONCRETE

5.4.1 Spall experiment of concrete

To demonstrate the capability of the model for predicting fracture location and fracture

strength at different strain rates, a spall experiment of concrete using a split Hopkinson

pressure bar (Schuler et al., 2006) was simulated with the proposed tensile damage model. A

brief review of spall experiment is provided in Section 2.4.2, Chapter 2. The experimental

setup of Schuler et al. (2006) consisted of aluminium bars with Young’s modulus, Poisson’s

ratio and density of 72.7 GPa, 0.34 and 2720 kg/m3, respectively (Schuler, 2007). Bars with a

diameter of 74.2 mm and specimens with the same diameter were used throughout their test.

A schematic of the spall test is given in Figure 5.9. Three impact velocities: 4.1 m/s, 7.6 m/s

82
and 11.1 m/s were used for the striker in spall experiment. Mechanical properties of concrete

used in the experiment were E = 38.9 GPa,  = 2320 kg/m3 and σst = 3.24 MPa. In this

simulation, Poissoin’s ratio was assumed as 0.2.

Striker Input bar Specimen

60 mm 5500 mm 250 mm

Figure 5.9. Schematic of spall experiment using Hopkinson bars.

The experimental strain rates and tensile strengths (σtd- σst) are plotted in logarithm format in

Figure 5.10. The model parameters α and β are determined as 1.45 ×109 s-1m-3 and 1.25,

respectively, using the procedure described in Chapter 4, Section 4.4.1. The critical strain

( ) was determined from Eq. (4.2) as 2.8 × 10-5. The geometry of the 3D model was based

on the experimental setup shown in Figure 5.9. In this FEM simulation, the eight node brick

elements (C3D8R) were used. Three simulations were performed with striker velocities as 4.1

m/s, 7.6 m/s and 11.1 m/s, following the experimental setup used by Schuler et al. (2006). A

constant velocity was applied to the the nodes of the striker bar to generate an impact loading

in Abaqus. The interface between sample and bars were assumed frictionless in all

simulations. No constraints are applied to nodes of the bars and sample .e.g. all the nodes had

rotational and translation degree of freedom.

83
1.2
log10 (σtd-σst) = 0.45 log10 ɛ + 0.29

Uniaxial stress log10 (σtd-σst) (MPa)


1.15

1.1

1.05

0.95
1.4 1.5 1.6 1.7 1.8 1.9 2

log10 (s-1)

Figure 5.10. Logarithmic plot of fracture stress versus strain rate obtained from spall

experiment (Schuler et al., 2006).

5.4.2 Mesh convergence study

Poor mesh size causes a wrong prediction of strength and fracture in finite element

simulation. For this purpose, a mesh convergence study was conducted with the proposed

damage model to find out the optimum mesh that can predict the actual scenario being most

efficient in computational cost. Continuum elements, C3D8R (eight node brick elements),

having mesh lengths 8 mm, 4 mm, 2 mm and 1 mm are studied and corresponding damage

contours are shown in Figure 5.11 for the impact velocity of 7.6 m/s. The stress histories of

the right marocrack for these four mesh sizes are shown in Figure 5.12. Damage contours

have better comparability (observed from Figure 5.11) between mesh 2 mm and 1 mm in

term of thicknesses in two macrocracks on the samples.

84
(a)

(b)

(c)

(d)

Figure 5.11. Mesh convergence study for impact velocity 7.6 m/s, with mesh size (a) 8 mm

(b) 4 mm (c) 2 mm and (d) 1 mm

85
20
15
10
Axial Stress (MPa)

5
0
1.08 1.12 1.16 1.2
-5
-10
-15 8 mm
4 mm
-20
2 mm
-25
1 mm
-30
Time (ms)

Figure 5.12. Axial Stress history to the first macrocrack (right hand side of the sample) for
different mesh sizes under impact 7.6 m/s.

It is observed from the Figure 5.10 that the axial stresses nearly overlap to each other with the

mesh sizes 2 mm and 1 mm. Therefore, it can be concluded that mesh with dimension 2 mm

is adequate enough to obtain a good prediction and a smaller mesh would show a negligible

sensitivity in results. All the simulations for the spall experiment were conducted with 2 mm

mesh in this study.

5.4.3 Spall simulation using the proposed damage model

The spall experiment by Schuler et al. (2006) was simulated by the proposed damage model

with three projectile impact velocities as described earlier. Figure 5.13 depicts the history of

the damage evolution when a striker with an impact velocity of 7.6 m/s is used. It can be seen

from Figures 5.13(a) and 5.13(b) that the first and second macro cracks appear approximately

86
105 s and 112 s after the impact. After this time, no additional macrocracks are developed

throughout the experiment. The distinctive red contours on the sample indicate the material

elements with damage value D of 1, which in turn show the fracture locations in the sample.

Figures 5.14(a), 5.14(b) and 5.14(c) compare the simulated damage contours of the

specimens with experimentally observed fractured sample at projectile velocities of 4.1 m/s,

7.6 m/s and 11.1 m/s, respectively. The damage legend is provided at the left side of each

impact simulation.

(a)

(b)

(c)

Figure 5.13. Damage evolution of spall test by projectile impact 7.6 m/s at time (a) 105μs (b)

112 μs (d) 150 μs

87
Impacted end Free end

(a)

Impacted end Free end

(b)

Impacted end(b) Free end

(c)

Figure 5.14. Comparison of experimental fracture locations (Schuler et al., 2006) and

simulated damage contour under projectile impact velocities of (a) 4.1 m/s, (b) 7.6 m/s and

(c) 11.1 m/s.

88
It can be seen in Figure 5.14 that the size of fragments decreases when the strain rate

increases. This was shown in the concrete spall experiment of Schuler et al. (2006) where two

major cracks were observed under projectile velocities of 7.6 m/s and 11.1 m/s as compared

to one major crack in the case of 4.1 m/s projectile impact. It can also be seen from Figure

5.14 that the proposed rate-depedent tensile damage model could effectively predict the

numbers of major cracks and their locations as impact volecity varies.

Figure 5.15 shows the uniaxial stress-strain curves at the first fracture location (the right side

of the macrocrack) under different impact velocities. The peak stress in Figure5.15 represents

the corresponding fracture strength of the material under the given impact velocity. It is

observed from Figure 5.15 that the fracture strength increases with the increase of projectile

impact velocity. The tensile strength and strain rate history at the right side of the macrocrack

in Figure 5.14(b) (for impact velocity 7.6 m/s) is plotted in Figure 5.16. It is observed in

Figure 5.16 that the strain rate increases rapidly once the fracture strength is exceeded by the

material. This observation can be explained by the model response to damage history in

Figure 5.6 where it is evident that the damage increases abruptly after the peak stress. Thus

the fragmentation process may start after the peak stress, which could lead to the abrupt

increase in deformation, consequently, showing a rapid increase of strain rate as observed in

Figure 5.16. Hence, the strain rate at this transition point (from where it starts to increase

rapidly after the peak stress) is considered as the strain rate at fracture. Simlarly, strain rate at

fracture strength for the other two impact velocities are plotted at the macrocrack locations in

Figure 5.17. In Figures 5.16 and 5.17, the points where the strain rate at fracture is observed

are marked with a triangle pointer.

89
Table 5.1 summarises the fracture strengths, strain rates and fracture locations obtained from

both the simulations and experimental data (Schuler et al., 2006). In the Table 5.1, the

distances 1 and 2 denote the first and second macrocrack locations from free end of the

sample, respectively. It can be seen from Table 5.1 that the predicted fracture strengths agree

very well with the experimentally-observed average fracture strength with an error of less

than 5% for all three impact velocities. Moreover, the fracture locations predicted by the

numerical simulations using the proposed tensile demage model matched the experimental

observations very well.

20

18

16
Axial Fracture Stress (MPa)

14

12
11.1 m/s
10

8
7.6 m/s
6

4 4.1 m/s

0
0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012
Axial Strain

Figure 5.15. Uniaxial stress-strain curves at the fracture planes under projectile impact

velocities.

90
18 140
16
Fracture strength σst (MPa) 120
14
100
12

Strain rate (s-1)


10 80

8 60
6
40
4 Tensile stress
20
2
Strain rate
0 0
1.165 1.17 1.175 1.18 1.185 1.19
Time (ms)

Figure 5.16. Tensile strength and strain rate evolution for an impact velocity of 7.6 m/s

180
160
140
11.1 m/s
120
Strain rate (s-1)

100
80
v 7.6 m/s
60
40 4.1 m/s
20
0
1.165 1.17 1.175 1.18 1.185 1.19 1.195 1.2 1.205
Time (ms)

Figure 5.17. Histories of axial strain rate of concrete at three differen timpact velocities

91
Table 5.1 Comparison of fracture strength and fracture location obtained from experiments

(Schuler et al., 2006) and numerical simulations.

Impact velocity(m/s)
Item
4.1 7.6 11.1

Dynamic uniaxial fracture strength Experiment 12.9 1 16.2 2 18 0.9

σtd (MPa) Simulation 13.3 16.7 18.8

Uniaxial strain rate Experiment 33 67.8 79.5

Simulation 20 55 75

First macrocrack location x1 Experiment 103 79 76

(mm) Simulation 85 74 85

Second macrocrack location x2 Experiment N/A 112 114

(mm) Simulation N/A 124 130

5.5 VALIDATION OF FRAGMENT SIZE MODEL

The fragment size prediction model is also validated using the spall experiment conducted by

Schuler et al. (2006) described in section 5.4.1. Figure 5.14 shows the recovered samples

(Schuler et al., 2006) for experiments under three impact velocities. In order to obtain a

fragment size under a given impact velocity, at least two fracture planes are required. In the

case of an impact velocity of 4.1 m/s, only one fracture plane is developed and fragment size

92
for this setup cannot be obtained. However, from Figure 5.14(b) and 5.14(c), it is observed

that a fragment size under a given impact velocity can be obtained using the fragment defined

by the two fracture planes. Table 5.2 reports the size of two fragments that have been

observed from the experiment. In Schuler et al. (2006), fracture toughness was not reported;

however, it has been reported in the literature that the fracture toughness of plain concrete

varies between 0.5 MPa/m1/2 and 1.7 MPa/m1/2 (Alam et al., 2010; Chen et al., 2015;

Karihaloo, 1986). A typical value of 0.8 MPa/m1/2 is assumed for the plain concrete in this

fragment size distribution study. Experimentally observed strain rates of 67 /s and 80/s (From

Table 5.1) are used to calculate the fragment sizes for impact velocities of 7.6/s and 11.1/s,

respectively. The procedure described in Chapter 4 has been used to determine the fragment

size for this model. Table 5.2 shows a comparison of the fragment sizes obtained using the

proposed fragment size model and estimated based on the experimental data reported by

Schuler et al. (2006). It is observed that the predicted fragment sizes agree well with the

experimental results with errors of less than 15%.

Table 5.2 Fragment sizes predicted by the proposed model and estimated from experiments

by Schuler et al. (2006)

Impact Strain n Aspect Fragment Fragment Fragment Fragment


velocity rate ratio length width size size
experiment
(m/s) (s-1) (b/a) a b (aeq)
(Schuler et
(mm) (mm) (mm) al., 2006)
(mm)

7.6 67 20 0.25 82 20 32 38

11.1 80 22 0.24 74 17.8 28 33

93
5.6 SUMMARY

The damage model has been numerically implemented in this chapter to simulate the high

strain rate behaviour of concrete under impact loading. The fracture strength predicted by the

proposed damage model is observed to increase with the increase of the strain rate. Also the

size of a fragment, which has been obtained from the model, has reduced when a higher strain

rate load has been applied to the material. The comparison of the model prediction over the

experimental observations has been conducted, specially observing the fracture strength,

strain rate, fracture location and fragment size that are obtained from different strain rate

impact to the concrete sample. The strain rate, fracture strength, fracture locations and

fragment size prediction have been found to be in good agreement with experimental

observations. Therefore, the model can be effectively used to study dynamic behaviour of

concrete under high strain rate loading.

94
CHAPTER 6

APPLICATION: SIMULATION OF DYNAMIC BRAZILIAN


DISC TEST

6.1 INTRODUCTION

The validation of the proposed damage model is extended to this chapter by simulating the

dynamic behaviour of rocks at high strain rate loading. Laboratory experiment using a split

Hopkinson pressure bar apparatus was conducted with Brazilian disc specimens described in

Chapter 3. The measured experimental data concerning fracture strength, strain rate and

fracture pattern are going to be compared with the numerical prediction of the damage model.

In this way, the applicability of the damage model to rock is verified.

In addition to the validation, the history of tensile cracking in Brazilian disc is going to be

studied numerically. The conflicting statements regarding the macrocrack initiation and

propagation in dynamic Brazilan disc test have been discussed in Section 2.7, Chapter 2. The

dynamic Brazilian disc study of this chapter provides the insight to the macrocrack initiation

and propagation under high strain rate impact loading.

6.2 MODELLING GEOMETRY OF FEM STUDY

In order to study the dynamic behaviour of sandstone in finite element analysis, three-

dimensional models were constructed for split Hopkinson pressure bar setup. The dimension

95
of Hopkinson bars (Ø 0.015 m × 1.5 m), samples (Ø 0.025 m × 0.01 m) and loading

conditions were same as those in the experiment so that the experimental observation can be

reproduced by the simulation. Finite element solver Abaqus/Explicit, version 6.11 was used

to generate the 3D model and necessary computations by the proposed damage model via a

user defined material subroutine. Eight node brick elements (C3D8R) were used for bars and

sample in this finite element study. In the experiment (Chapter 3) a striker bar was used to

generate the stress pulse in the input bar in the split Hopkinson pressure bar study. However,

in this FEM study, the impact of striker was replaced by a time-stress boundary condition

applied to the input bar end, and it was explained in the following section. Figure 6.1 shows

the geometry of the model used in this simulation. Throughout this chapter the Cartesian co-

ordinate shown in Figure 6.1 is used in describing the stress/strain direction in the simulation.

Input bar Output bar

Y
Figure:

Sample
Z
X

Figure 6.1. The FEM model of Brazilian disc simulation with split Hopkinson pressure bar

setup.

96
6.3 DYNAMIC LOADING AND BOUNDARY CONDITION FOR SHPB

The Brazilian disc experiment (Chapter 3) was performed with striker lengths of 160 mm and

120 mm. In that experiment the incident pulse was shaped with the aid of copper (C11000)

pulse shaper. Figure 6.2 shows the experimentally observed uniaxial stress histories at the

mid length location of input bar for impact velocity 11.1 m/s, 15.2 m/s and 20.4 m/s,

respectively. In case of 20.4 m/s impact, a striker bar of 120 mm was used in the experiment.

In order to reduce the computation time, the simulations were carried out with the time-stress

boundary condition at the impact end of input bar as shown in Figure 6.2. The interface

between sample and bars were assumed frictionless in all simulations. No constraints were

applied to nodes of the bars and sample .e.g. all the nodes had rotational and translation

degree of freedom.

50

0
0 0.04 0.08 0.12 0.16 0.2
-50
Axial stress in bar σz (MPa)

-100

-150

-200

-250
11.1 m/s
-300 15.2 m/s
20.4 m/s
-350

-400
Time (ms)

Figure 6.2. Experimental obtained incident stress pulses at the mod length location of input

bar for three impact velocities.

97
6.4 MATERIAL MODEL

The proposed damage model and Drucker Prager model were used in dynamic tension and

compression simulation, respectively. Both the models were calibrated from the necessary

experiments that were carried out in Chapter 3. Using the bulk modulus and quasi-static

tensile strength, the critical strain of the material was calculated as 1.5 × 10-4 from Eq. (4.2).

The model parameter α and β were calculated as 1.1 × 108 s-1m-3 and 1.25, respectively,

which were obtained by the procedure demonstrated in Chapter 4 (Section 4.4) and

mechanical data that were reported in Tables 3.1-3.3 in Chapter 3. Table 3.1 was used to

determine the necessary calibration for the Drucker Prager model. An associated flow rule

was assumed for the Drucker-Prager model. The slope of linear yield surface at P-μDP plane

was considered as 40˚ and the value of kDP was taken as 1 in this study since those were the

typical values observed for brittle materials (Li et al., 2009; Park et al., 2001). The steel bars

(margarine steel) used in this simulation were assumed elastic. The material properties of bars

were: v = 0.29 and E = 210 GPa and ρ = 7800 kg/m3.

6.5 MESH CONVERGENCE STUDY

The eight node brick element, C3D8R, was selected for this finite element study. The mesh

convergence study was performed with mesh size 0.8 mm, 0.4 mm, 0.2 mm and 0.1 mm in

Brazilian disc sample. Figures 6.3 and 6.4 show the damage contour and tensile stress (σy)

history at the centre of the sample under the impact velocity of 11.1 m/s, respectively. The

distinctive red mark on the sample (see Figure 6.3) indicates the elements that have a damage

parameter of 1. It has been observed from Figures 6.3 and 6.4 that the mesh with element size

98
0.2 mm has well converged to predict both fracture strength and fracture plane for this

simulation. All the simulations were performed with mesh size 0.2 mm in this current study.

(a)

(b)

(c)

(d)

Figure 6.3. Damage contour of the disc at 150 ms by element size (a) 0.8 mm (b) 0.4 mm (c)

0.2 mm and (d) 0.1 mm.

99
12

10
Tensile stress (MPa)

6
0.8mm
0.4mm
4
0.2mm
0.1 mm
2

0
0.28 0.33 0.38 0.43
Time (ms)

Figure 6.4. Mesh convergence study for Brazilian disc test.

6.6 RESULTS AND DISCUSSION

6.6.1 Damage contour and fracture pattern

Figure 6.5 shows the tensile stress distribution of Brazilian disc sample when (at time 36 ms)

the fracture strength of the material is reached under the impact velocity of 11.1 m/s. The

dark colour on the sample indicates the location where the elements experience compression

(ɛy < 0) during the loading. The x-z plane shown in the Figure 6.5 passes through the centre of

the disc sample. It should be noted that the x-z plane shown in the Figure 6.5 is the fracture

plane of the sample. The concentration of tensile stress is observed near four corners of the x-

z plane as shown in Figure 6.5 where the stress is higher than anywhere else on that plane. A

100
similar tensile stress concentration (with different stress levels) near four corners as observed

in Figure 6.5 was also observed under the other two impact velocities.

In order to claim a valid split Hopkinson pressure bar experiment, it is essential that the

sample reach equilibrium before the failure occurs. The elements that were located 2 mm

apart from the two interface corners of the sample, and located on the fracture plane (x-z

plane passing through sample centre) were selected to output the stress histories. The tensile

stress history of these two locations for impact velocities 11.1 m/s, 15.2 m/s and 20.4 m/s are

plotted in Figures 6.6, 6.7 and 6.8, respectively. In case of impact velocity 11.1 m/s (Figure

6.6), the equilibrium in the sample is well reached 34 ms after the impact. In contrast, a time

lag is observed between stress waves in two ends of the sample for impact velocities 15.2 m/s

and 20.4 m/s that can be observed from Figure 6.7 and 6.8, respectively. However, a similar

trend in equilibrium was observed in low and high strain rate impact by Saksala et al. (2013),

where a numerical study of Brazilian disc was conducted with split Hopkinson pressure bar

test. Overall, it can be concluded that the equilibrium was approximately reached for impact

velocity 15.2 m/s and 20.4 m/s in this study.

101
Section x-z
σy

Figure 6.5. Tensile stress distribution of the sample at time 36 ms under impact velocity 11.1.

m/s.

12

10
Sample-Input end
Tensile stress σy (MPa)

8
Sample-Output end

0
0.28 0.31 0.34 0.37 0.4
Time (ms)

Figure 6.6. Tensile stress evolution at the two ends of the disc at impact velocity 11.1 m/s.

102
12

10
Tensile stress σy (Mpa) Sample-Input bar end

8
Sample-Output bar end
6

0
0.28 0.3 0.32 0.34 0.36 0.38
Time (ms)

Figure 6.7. Tensile stress evolution at the two ends of the disc at impact velocity 15.2 m/s.

16

14

12 Sample-Inout end
Tensile stress (MPa)

Sample-Output end
10

0
0.28 0.3 0.32 0.34 0.36
Time (ms)

Figure 6.8. Tensile stress evolution at the two ends of the disc at impact velocity 20.4 m/s.

103
D

(a)

(b)

(c)

Figure 6.9. Damage evolution of Brazilian disc under impact velocity of 11.1 m/s at time: (a)

0.35 ms (b) 0.38 ms and (c) 0.5 ms.

104
D

(a)

(b)

(c)

Figure 6.10. Damage evolution of Brazilian disc under impact velocity of 15.2 m/s at time:

(a) 0.34 ms (b) 0.36 ms and (c) 0.5 ms.

105
D

(a)

(b)

(c)

Figure 6.11. Damage evolution of Brazilian disc under impact velocity of 20.4 m/s at time:

(a) 32 ms, (b) 35 ms and (c) 0.50 ms.

106
(a)

(b)

(c)

Figure 6.12. Damage contour at the end of simulation (at time of 0.50 ms) under impact

velocity (a) 11.1 m/s (b) 15.2 m/s and (c) 20.4 m/s.

107
Figures 6.9-6.11 show the damage evolution in the sample for projectile impact velocity of

11.1 m/s, 15.2 m/s and 20.4 m/s, respectively. The damage contours are shown in a similar

isometric view (one half of the sample) as mentioned for Figure 6.5. The specific attentions

of Figures 6.9-6.11 are to observe the macrocraking/damage pattern on the tensile fracture

plane under the studied impact loadings. It is observed from the Figures 6.9 (a), 6.10 (a) and

6.11(a) that the damage is concentrated to the four corners of the tensile fracture planes at the

early time of the loading. Thus, both stress (see Figure 6.5) and damage concentrations are

observed at the four corners of the sample.

Figures 6.9 (b), 6.10 (b) and 6.11(b) show the damage contour for the impact velocity 11.1

m/s, 15.2 m/s and 20.4 m/s at the time of 38 ms, 36 ms and 35 ms, respectively. The time

mentioned in those figures (Figures 6.9 (b), 6.10 (b) and 6.11(b)) are the time when the

damage is observed to form a plane spreading along the thickness of the sample.

Although damage concentrations are observed at the four corners of the sample (see Figures

6.9(a), 6.10(a) and 6.11(a)), the fracture plane is not initiated from those sample ends. In case

of lower impact velocities such as 11.1 m/s and 15.2 m/s, the fracture plane is observed to

start very close to the sample centre (see Figure 6.9 (b) and Figure 6.10 (b)). In contrast, in

one case of 20.4 m/s impact velocity, the fracture plane initiates between sample centre and

sample-input bar face (see Figure 6.11 (b)). Two-dimensional numerical simulation (Chen et

al., 2014(a); Hughes et al., 1993; Zhou and Hao, 2008) and experimental observations

(Saksala et al., 2013; Wang et al., 2009) also confirmed that the fracture plane was initiated

between disc centre and sample-input bar face in a high strain rate study with split Hopkinson

bar test. Eventually, with the increase in loading time, it is observed that the damage spreads

from the sample centre towards the sample ends on the fracture planes as shown in Figures

108
6.9(c), 6.10(c) and 6.11(c). The damage contours at the end of the simulation (0.50 ms) under

three different impact velocities are shown in Figure 6.12.

However, it should be noted that the initiation of fracture plane had a conflicting findings in

Li and Wong (2013) where it was stated that that the initiation of macrocrack could occur

from the sample ends. They observed that the stress and strain concentrated near the sample

ends, and based on this observation they predicted that the fracture plane might be initiated

near the sample-bar interface which is a contrast to well observed macrocrack initiation near

the sample centre (Chen et al., 2014; Saksala et al., 2013; Wang et al., 2009). The current

simulation of Brazilian disc showed that, despite of having stress and damage concentrations

near sample edges during the early period of loading, the fracture plane always initiated near

disc centre and then propagate towards sample ends to complete the diametrical splitting of

the disc.

6.6.2 Fracture Strength and strain rate

Figures 6.13 and 6.14 show the evolution of tensile stress (σy) and strain rate at the centre of

the disc under three different impact velocities, respectively. The peak stresses in Figure 6.13

are the fracture strength of sandstone predicted by the damage model. In order to demonstrate

the strain rate at fracture, the evolution of fracture stress and strain rate are plotted in Figures

6.15, 6.16 and 6.17 for impact velocities of 11.1 m/s, 15.2 m/s and 20.4 m/s, respectively. In

order to better visualise the strain rate at strength peak, the strain rate scale was adjusted to

the maximum of 500/s in Figure 6.17. Similar to the experiment, it is observed that the strain

rate increases sharply at the transition point where the triangular mark is positioned in Figures

6.15-6.17. The simulation revealed that the point is located very close to the time when the

109
fracture strength is reached. Upon reaching the peak stress, the stress plunges to zero within a

very small period of time. When the fracture happens to the sample, the strain becomes

infinity with respect to loading. As a consequence, the strain rate shoots up at this fracture

point. The triangular marks in the Figures 6.15-6.17 represent the strain rates at the time of

the fracture of the sample. In the dynamic Brazilian disc experiment (Chapter 3), it was

assumed that the point beyond which strain rate increased abruptly was the point of fracture

in the disc. Thus, this numerical study captured the material response and confirmed this

phenomenon.

18
16
14
12
Tensile stress (MPa)

10
11.1 m/s
8
15.2 m/s
6
20.4 m/s
4
2
0
0.28 0.3 0.32 0.34 0.36 0.38 0.4

Time (ms)

Figure 6.13. Tensile stress (σy) evolution at the centre of Brazilian disc at three impact

velocities.

110
1400

1200

1000
Strain rate (s-1)

800
11.1 m/s
600 15.4 m/s
20.4 m/s
400

200

0
0.28 0.3 0.32 0.34 0.36 0.38

Time (ms)

Figure 6.14. The strain rate history at the centre of the Brazilian disc for three impact

velocities.

12
200
10
Fracture stress
Tensile Stress (σy) (MPa)

8 Strain rate 150


Strain rate (s-1)

6
100

4
50
2

0 0
0.28 0.3 0.32 0.34 0.36 0.38 0.4
Time (ms)

Figure 6.15. Evolution of tensile stress and strain rate at the centre of the disc at the impact of

11.1 m/s.

111
14 500
450
12
Tensile stress (σy) (MPa) 400
Tensile stress
10 350
Strain rate

Strain rate (s-1)


8 300
250
6 200
4 150
100
2
50
0 0
0.28 0.3 0.32 0.34 0.36 0.38
Time (ms)

Figure 6.16. Evolution of tensile stress and strain rate at the centre of the disc at the impact of

15.2 m/s.

18 500
16 450
Tensile stress (σy) (MPa)

14 400
Tensile stress 350
12
Strain rate
Strain rate (s-1)

300
10
250
8
200
6
150
4 100
2 50
0 0
0.28 0.3 0.32 0.34 0.36
Time (ms)

Figure 6.17. Evolution of tensile stress and strain rate at the centre of the disc at the impact of

20.4 m/s.

112
Table 6.1 shows the comparison of experimental data and simulated results using the

proposed damage model. A plot between strain rate and fracture strength is shown in Figure

6.18, providing both experimental and numerical observations. As expected, both the strain

rate and fracture strength are observed to increase with the increase of impact velocity of

striker. It is observed from the Table 6.1 and Figure 6.18 that the predicted strain rate and

fracture strength are in a good agreement with the experiment.

Table 6.1 Comparison of fracture strength and strain rate obtained from experiments and
numerical simulations.

Impact velocity
Item Result obtained
from 11.1 15.2 20.4
(m/s) (m/s) (m/s)

Experiment 8.9 1.4 10.5 1 14.4 2


Fracture Strength (σtd)

(MPa) Simulation 10.9 12.8 16.75

Strain rate ( ̇ Experiment 24.4 8.5 50 16 96 13


(s-1)
Simulation 35 72 120

113
1.3

1.2

1.1
log10 𝜎𝑦 (MPa)

0.9
Experiment
0.8 Model prediction

0.7

0.6
0.8 1.2 1.6 2 2.4

log10 ̇ 𝑦 (s-1)

Figure 6.18. Strain rate and fracture strength data for dynamic Brazilian disc test which are

obtained from experiment and simulation.

6.7 SUMMARY

Laboratory experiments using a split Hopkinson pressure bar apparatus has been conducted

and the experimental observations have been used to validate the proposed damage model.

The simulated results including tensile strength, strain rates and fracture pattern at three

different impact velocities have been observed in good agreement with the experimental

findings. It has been observed from the 3D simulations that the stress and damage concentrate

near the four corners of the fracture planes of the disc sample, but the fracture plane has not

initiated from the sample ends. The damage that has contributed to the diametrical splitting

has initiated near the sample centre and propagates towards the sample ends to complete the

fracture.

114
CHAPTER 7

CONCLUSION AND FUTURE WORK

7.1 CONCLUSION

In this work, the constitutive response of brittle solid under dynamic tensile loading has

studied by means of analytical modelling, experimental observations and numerical

simulations. A rate-dependent tensile damage model was developed to predict dynamic

behaviour of brittle solid at high strain rate loading. Taking microcrack-induced damage and

energy into consideration, the model has able to predict the fragment size under a constant

strain rate loading. The scope of the work was to ensure a sound formulation in modelling

and to show the validity of the proposed model by comparing the experimental observations

with numerical simulations. In order to claim the model applicability in different types of

brittle materials, validation of the damage model was studied for concrete and rocks under

dynamic loading. The following conclusions have been drawn from this study

1) The elasto-damage behaviour of brittle solid has been monitored by the reduction of bulk

modulus as a consequence of the growth of microcrack density. This reduction of bulk

modulus was obtained by the effective moduli of the cracked solid under applied loading

which was developed by Budiansky and O’Connell (1975). A power law relation of

volumetric strain with microcrack density, which had a similar form as Paris law for metal

damage evolution, has been implemented in proposed damage model to calculate the

115
microcrack density as a function of volumetric strain. Calibration of the model has been

straightforward as it requires neither iterations nor sophisticated experiments to obtain the

mechanical properties of the material. The model can be implemented to the macro level

analysis; therefore, typical large volume of study is practical with this model and it is more

efficient in term of computation cost. The macrocrck/fracture plane can be predicted by this

damage model with the details of damage contour. This model can also be applicable to study

fracture strength, strain rate and the duration of tensile loading at a dynamic condition of

brittle materials.

2) Using the energy balance, a fragment size prediction method has been developed. The

fragment size prediction model has been more practical than the conventional approach in

energy balance (Glenn and Chudnovsky, 1986; Grady, 1988, 1982; Zhang et al., 2006) as the

model has considered the aspects: a) Complete strain energy until the fracture of the material,

b) The surface energy of the fragment has been formulated with a prolate spheroid fragment

assumption that has been able to capture more accurate surface area (compared to

conventional sphere assumption) obtained from an irregular shape and size of the fragment.

As a consequence, the global energy balance has been more accurate in term of energy

calculations and a sound prediction of fragment size under a high strain rate loading can be

ensured.

3) Quasi-static tests have been performed on Gosford sandstone (commercially named as

Piles Creek) to obtain quasi static compression, tension and Young’s modulus as 46 MPa,

2.65 MPa and 9.1 GPa, respectively. The dynamic tensile strengths of the material have been

observed to be 8.9 MPa, 10.5 MPa and 14.4 MPa for strain rate of 24/s, 50/s and 96/s,

respectively.

116
4) The constitutive model has been numerically implemented to study 3D dynamic problem

for brittle material. A spall experiment with concrete and a dynamic Brazilian disc

experiment with sandstone have been numerically studied by the propose model. Fracture

strengths, strain rates, fracture location and fragment sizes that were obtained from the

experiments under different strain rates have been compared with numerical predictions. The

numerical study has revealed that the fracture strength, strain rate, fracture locations and

fragment size prediction have been found to be in good agreement with experimental

observations.

4) The fracture propagation on the tensile splitting plane of Brazilian disc has been shown in

this study using the proposed damage model. It has been observed from the 3D simulation

that the stress/damage concentration has occurred near four corners of the fracture plane

during the early period of loading, but the fracture, which has dominated the diametrical

splitting of the disc, has not initiated from those corners or even from the sample ends. The

damage evolution on the fracture plane has confirmed that the fracture plane initiated near the

disc centre. Eventually, with time the crack propagated towards the sample ends to

accomplish the diametrical splitting. In case of higher strain rate (96 s-1 from experiment) the

initiation point of fracture has been observed to shift between sample centre and sample-input

bar end, being the starting point closer to disc centre. The fracture evolution found in this

current study agreed with the classical observation of crack propagation pattern in Brazilian

disc.

117
7.2 FUTURE WORK

The response of stress strain relation in post-strength peak region (e.g. softening modulus)

has not been studied experimentally in this work. As the model was able to predict the whole

stress strain curve, it was necessary to know the model’s capability in determining the

softening modulus under a dynamic condition. For example, how the model responses to

extreme brittle and quasi-brittle material in post strength-peak regime has to be studied and

verified. In fact, obtaining a complete stress strain curve in dynamic tensile experiment is

quite hard to perform and most of the experiment methods such as spall test, dynamic

Brazilian disc test and plate impact tests are unable to provide the stress strain response. The

dynamic Brazilian disc test, which has been performed in this study, could only provide an

indirect tensile strength under a strain rate. Thus, direct tension test with the split Hopkinson

bar facility has to be performed to obtain a post strength peak response between stress and

strain. Next, a numerical study with this current model that simulates the experiment would

be able to justify this softening modulus predicted by this model. In case, the model

prediction differs with experiment in softening modulus issue, it can be improved using the

de-cohesion method to model the post strength peak stress strain response under dynamic

tension.

In this current work, a damage model and a fragment size distribution model under dynamic

tension was formulated. However, when the material experiences compression, a linear

elasto-plastic Drucker-Prager model was used along with this tensile damage model. Since

Drucker Prager is a pressure dependent model, it does not explicitly address the rate

sensitivity of material in compression. Moreover, the Drucker-Prager model that has been

used here did not have any internal variable that could visualise the shear cracking in

118
compressive loading. The future work will be to include/develop a rate dependent

compressive model which could visualise compressive failure.

119
REFERENCES

ABAQUS 6.11, 2011. Abaqus user subroutine reference manual, VUMAT, SIMULIA.
Alam, M.R., Azad, M.A.K., Kadir, M.A., 2010. Fracture toughness of plain concrete
specimens made with industry-burnt brick aggregates. J. Civ. Eng. 38, 81–94.
Albertini, C., Montagnani, M., 1994. Study of the True Tensile Stress-Strain Diagram of
Plain Concrete with Real Size Aggregate; Need for and Design of a Large Hopkinson
Bar Bundle. J. Phys. IV 4, 113–118.
Asprone, D., Cadoni, E., Prota, A., Manfredi, G., 2009. Dynamic behavior of a
Mediterranean natural stone under tensile loading. Int. J. Rock Mech. Min. Sci. 46, 514–
520. doi:10.1016/j.ijrmms.2008.09.010
ASTM, 1996. Standard test method for splitting tensile strength of cylindrical concrete
specimens. c45 West Conshohocken, PA.
Bagde, M., Petros, V., 2005. Fatigue properties of intact sandstone samples subjected to
dynamic uniaxial cyclical loading. Int. J. Rock Mech. Min. Sci. 42, 237–250.
doi:10.1016/j.ijrmms.2004.08.008
Berfield, T.A., Patel, J.K., Shimmin, R.G., Braun, P. V., Lambros, J., Sottos, N.R., 2007.
Micro- and Nanoscale Deformation Measurement of Surface and Internal Planes via
Digital Image Correlation. Exp. Mech. 47, 51–62. doi:10.1007/s11340-006-0531-2
Bieniawski, Z.T., 1967. Mechanics of Brittle Fracture of Rocks. Int. J. Rock Mech. Min. Sci.
4, 395–406.
Boade, R.R., Kipp, M.E., Grady, D.E., 1981. Blasting concept for preparing vertical modified
in-situ oil shale reports.
Brace, W.F., Jones, A.H., 1971. Comparision of Uniaxial Deformation of Shock and Static
Loading of Three Rocks. J. Geophys. Res. 76, 4913–4921.
Brara, A., Camborde, F., Klepaczko, J.R., Mariotti, C., 2001. Experimental and numerical
study of concrete at high strain rates in tension. Mech. Mater. 33, 33–45.
doi:10.1016/S0167-6636(00)00035-1
Budiansky, B., O’Connell, R.J., 1975. Elastic Moduli of a Cracked Solid. Int. J. Solids Struct.
12, 81–97.
Cadoni, E., 2010. Dynamic Characterization of Orthogneiss Rock Subjected to Intermediate
and High Strain Rates in Tension 667–676. doi:10.1007/s00603-010-0101-x
Chen, H.H., Su, R.K.L., Kwan, A.K.H., 2015. Crushed Granite Aggregate Fracture
Toughness of Plain Concrete Made of Crushed Granite Aggregate. HKIE Trans. 18, 6–
12.
Chen, R., Dai, F., Qin, J., Lu, F., 2013a. Flattened Brazilian Disc Method for Determining the
Dynamic Tensile Stress-Strain Curve of Low Strength Brittle Solids. Exp. Mech. 53,
1153–1159. doi:10.1007/s11340-013-9733-6

120
Chen, R., Dai, F., Qin, J., Lu, F., 2013b. Flattened Brazilian Disc Method for Determining the
Dynamic Tensile Stress-Strain Curve of Low Strength Brittle Solids. Exp. Mech. 53,
1153–1159. doi:10.1007/s11340-013-9733-6
Chen, W.W., Song, B., 2011. Split Hopkinson (Kolsky) bar: design, testing and applications.
Springer, New York.
Chen, X., Wu, S., Zhou, J., 2014a. Quantification of dynamic tensile behavior of cement-
based materials. Constr. Build. Mater. 51, 15–23.
doi:10.1016/j.conbuildmat.2013.10.039
Chen, X., Wu, S., Zhou, J., 2014b. Experimental Study on Dynamic Tensile Strength of
Cement Mortar Using Split Hopkinson Pressure Bar Technique. J. Mater. Civ. Eng. 26,
1–10. doi:10.1061/(ASCE)MT.1943-5533.0000926
Cho, S.H., Ogata, Y., Kaneko, K., 2003. Strain-rate dependency of the dynamic tensile
strength of rock. Int. J. Rock Mech. Min. Sci. 40, 763–777. doi:10.1016/S1365-
1609(03)00072-8
Clifton, R.J., 2000. Response of materials under dynamic loading. Int. J. Solids Struct. 37,
105–113. doi:10.1016/S0020-7683(99)00082-7
Cotsovos, D.M., Pavlović, M.N., 2008. Numerical investigation of concrete subjected to high
rates of uniaxial tensile loading. Int. J. Impact Eng. 35, 319–335.
doi:10.1016/j.ijimpeng.2007.03.006
Curran, D., Seaman, L., Shockey, D.A., 1987. Dynamic failure of solids. Phys. Rep. 147,
253–388. doi:10.1016/0370-1573(87)90049-4
Dai, F., Xia, K., 2010. Loading Rate Dependence of Tensile Strength Anisotropy of Barre
Granite. Pure Appl. Geophys. 167, 1419–1432. doi:10.1007/s00024-010-0103-3
Drugan, w. J., 2001. Dynamic Fragmentation of Brittle Materials : analytical mechanics-
based models. J. Mech. Phys. Solids 49, 1181–1208.
Esen, S., Onederra, I., Bilgin, H. a., 2003. Modelling the size of the crushed zone around a
blasthole. Int. J. Rock Mech. Min. Sci. 40, 485–495. doi:10.1016/S1365-
1609(03)00018-2
Fahrenthold, E.P., 1991. A continuum Damage for Facture of Brittle Solids Under Dynamic
Loading. J. Appl. Mech. 58, 904–909.
Galvez Diaz-Rubio, F., Rodriguez Perez, J., Sanchez Galvez, V., 2002. The spalling of long
bars as a reliable method of measuring the dynamic tensile strength of ceramics. Int. J.
Fract. 27, 161–177. doi:http://dx.doi.org/10.1016/S0734-743X(01)00039-2
Glenn, L.A., Chudnovsky, A., 1986. Strain-energy effects on dynamic fragmentation. J. Appl.
Phys. 59, 1379. doi:10.1063/1.336532
Gomez, J.T., Shukla, A., Sharma, A., 2001. Static and dynamic behavior of concrete and
granite in tension with damage. Theor. Appl. Fract. Mech. 36, 37–49.
doi:10.1016/S0167-8442(01)00054-4
Grady, D.E., 1982. Local inertial effects in dynamic fragmentation. J. Appl. Phys. 53, 322.

121
doi:10.1063/1.329934
Grady, D.E., 1988. The Spall Strength of Condenced Matter. J. Mech. Phys. Solids 36, 353–
384.
Grady, D.E., Kipp, M.E., 1979. The Micromechanics of Impact Fracture of Rock *. Int. J.
Rock Mech. Min. Sci. Geomech. Abstr. 16, 293–302.
Grady, D.E., Kipp, M.E., 1980. Continuum modelling of explosive fracture in oil shale. Int. J.
Rock Mech. Min. Sci. Geomech. Abstr. 17, 147–157. doi:10.1016/0148-9062(80)91361-
3
Grantham, S.G., Siviour, C.R., Proud, W.G., Field, J.E., 2004. High-strain rate Brazilian
testing of an explosive simulant using speckle metrology. Meas. Sci. Technol. 15, 1867–
1870. doi:10.1088/0957-0233/15/9/025
Gui, Y., Bui, H.H., Kodikara, J., Zhang, Q., Zhao, J., 2016. Modelling the dynamic failure of
brittle rocks using a hybrid continuum-discrete element method with a mixed-mode
cohesive fracture model. Int. J. Impact Eng. 87, 146–155.
Hamdi, E., Romdhane, N.B., Le Cléac’h, J.M., 2011. A tensile damage model for rocks:
Application to blast induced damage assessment. Comput. Geotech. 38, 133–141.
doi:10.1016/j.compgeo.2010.10.009
Hild, F., Denoual, C., Forquin, P., Brajer, X., 2003. On the probabilistic–deterministic
transition involved in a fragmentation process of brittle materials. Comput. Struct. 81,
1241–1253. doi:10.1016/S0045-7949(03)00039-7
Howe, S., Goldsmith, W., Sackman, J., 1974. Macroscopic Static and Dynamic Mechanical
Properties of Yule Marble. Exp. Mech. 14, 337–346. doi:10.1007/BF02323559
Huang, F., Wu, H., Jin, Q., Zhang, Q., 2006. A numerical simulation on the perforation of
reinforced concrete targets. Int. J. Impact Eng. 32, 173–187.
doi:10.1016/j.ijimpeng.2005.05.009
Hughes, M.L., Tedesco, J.W., Ross, C.A., 1993. Numerical Analysis of High Strain Rate
Splitting-TensileTests. Comput. Struct. 47, 653–671.
Kanchibolta, S.S., 2003. Optimum Blasting? Is it Minimum cost per Brocken Rock or
Maximum value per Brocken Rock? Int. J. Blasting Fragm. 7, 35–48.
doi:10.1076/frag.7.1.35.14059
Karihaloo, B.L., 1986. Fracture toughness of plain concrete from compression splitting tests.
Int. J. Cem. Compos. Light. Concr. 8, 251–259. doi:10.1016/0262-5075(86)90052-7
Kubota, S., Ogata, Y., Wada, Y., Simangunsong, G., Shimada, H., Matsui, K., 2008.
Estimation of dynamic tensile strength of sandstone. Int. J. Rock Mech. Min. Sci. 45,
397–406. doi:10.1016/j.ijrmms.2007.07.003
Levy, S., Molinari, J.F., 2010. Dynamic fragmentation of ceramics, signature of defects and
scaling of fragment sizes. J. Mech. Phys. Solids 58, 12–26.
doi:10.1016/j.jmps.2009.09.002

122
Li, X. B., Chen, S. R., & Gu, D. S. (1994). Dynamic Strength of Rock Under Impulse Loads
with Different Stress Waveforms and Durations. J. Central -South Institute of Mining
and Metallurgy, 25(3).

Li, D., Wong, L.N.Y., 2013. The Brazilian Disc Test for Rock Mechanics Applications:
Review and New Insights. Rock Mech. Rock Eng. 46, 269–287. doi:10.1007/s00603-
012-0257-7
Li, Q.M., Lu, Y.B., Meng, H., 2009. Further investigation on the dynamic compressive
strength enhancement of concrete-like materials based on split Hopkinson pressure bar
tests. Part II: Numerical simulations. Int. J. Impact Eng. 36, 1327–1334.
doi:10.1016/j.ijimpeng.2009.04.010
Li, R., Michael, A.H., Henson, M.A., Kurtz, M.J., 2004. Selection of model parameters for
off-line parameter estimation. IEEE Trans. Control Syst. Technol. 12, 402–412.
doi:10.1109/TCST.2004.824799
Li, X.B., Chen, S.R., Gu, D.S., 1994. Dynamic Strength of Rock Under Impulse Loads with
Different Stress Waveforms and Durations. J. Cent. -South Inst. Min. Metall. 25.
Liu, L., Katsabanis, P.D., 1997a. Development of a Continuum Damage Model for Blasting
Analysis. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 34, 217–231.
Liu, L., Katsabanis, P.D., 1997b. A numerical study of the effects of accurate timing on rock
fragmentation. Int. J. Rock Mech. Min. Sci. 34, 817–835. doi:10.1016/S1365-
1609(96)00067-8
Lu, Y., Xu, K., 2004. Modelling of dynamic behaviour of concrete materials under blast
loading. Int. J. Solids Struct. 41, 131–143. doi:10.1016/j.ijsolstr.2003.09.019
Lu, Y.B., Li, Q.M., 2011. About the dynamic uniaxial tensile strength of concrete-like
materials. Int. J. Impact Eng. 38, 171–180. doi:10.1016/j.ijimpeng.2010.10.028
Luccioni, B.M., Ambrosini, R.D., Danesi, R.F., 2004. Analysis of building collapse under
blast loads. Eng. Struct. 26, 63–71. doi:10.1016/j.engstruct.2003.08.011
Ma, G.W., An, X.M., 2008. Numerical simulation of blasting-induced rock fractures. Int. J.
Rock Mech. Min. Sci. 45, 966–975. doi:10.1016/j.ijrmms.2007.12.002
Ma, G.W., Hao, H., Zhou, Y.X., 1998. Modeling of wave propagation induced by
underground explosion. Comput. Geotech. 22, 283–303. doi:10.1016/S0266-
352X(98)00011-1
Mahabadi, O.K., Cottrell, B.E., Grasselli, G., 2010. An Example of Realistic Modelling of
Rock Dynamics Problems: FEM/DEM Simulation of Dynamic Brazilian Test on Barre
Granite. Rock Mech. Rock Eng. 43, 707–716. doi:10.1007/s00603-010-0092-7
Maiti, S., Rangaswamy, K., Geubelle, P.H., 2005. Mesoscale analysis of dynamic
fragmentation of ceramics under tension. Acta Mater. 53, 823–834.
doi:10.1016/j.actamat.2004.10.034
Medvedovski, E., 2010. Ballistic performance of armour ceramics: Influence of design and
structure. Part 1. Ceram. Int. 36, 2103–2115. doi:10.1016/j.ceramint.2010.05.021

123
Meyers, M.A., 1994a. Dynamic failure : mechanical and microstructural aspects. J. Phys. IV
4, 597–621.
Meyers, M.A., 1994b. Dynamic Behavior of Materials, Second. ed. John Wiley & Sons.
Meyers, M.A., Subhash, G., Kad, B.K., Prasad, L., 1994. Evolution of microstructure and
shear-band formation in Alpha-hcp titanium. Mech. Mater. 17, 175–193.

Miller, O., Freund, L.B., Needleman, A., 1999. Modeling and simulation of dynamic
fragmentation in brittle materials 101–125.
Mohammed, E., Smilowitz, R., Rittenhouse, T., 1996. Blasr Resistance Design of
Commercial Building. Pract. Period. Struct. Des. Constr. 31–39.
Munjiza, A., Owen, D.R.J., Bicanic, N., 1995. A combined finite-discrete element method in
transient dynamics of fracture solids. Eng. Comput. 12, 145–174.
Nemat-Nasser, S., Deng, H., 1994. Strain-Rate Effect on Brittle Failure in Compression. Acta
Met. Mater. 42, 1013–1024.
Norville, H.S., Conrath, E.J., 2006. Blast-Resistant Glazing Design. J. Archit. Eng. 12, 129–
136. doi:10.1061/(ASCE)1076-0431(2006)12:3(129)
Park, S.W., Xia, Q., Zhou, M., 2001. Dynamic behavior of concrete at high strain rates and
pressures: II. Numerical simulation. Int. J. Impact Eng. 25, 887–910.
doi:10.1016/S0734-743X(01)00021-5
Perrodon, A., Marshall, N., 1983. Dynamics of oil and gas accummulations. Elf-Aquitaine.
Pugno, N., Ciavarella, M., Cornetti, P., Carpinteri, a, 2006. A generalized Paris’ law for
fatigue crack growth. J. Mech. Phys. Solids 54, 1333–1349.
doi:10.1016/j.jmps.2006.01.007
Raghupathy, R., Gazonas, G.A., Molinary, J.F., Zhou, F., 2006. Numerical Convergence of
the Cohesive Element Approach in Dynamic Fragmentation Simulations, in: Shock
Compression of Condensed Matter. pp. 654–657. doi:10.1063/1.2263407
Rocco, C., Guinea, G. V., Planas, J., Elices, M., 2001. Review of the splitting-test standards
from a fracture mechanics point of view. Cem. Concr. Res. 31, 73–82.
doi:10.1016/S0008-8846(00)00425-7
Rouabhi, A., Tijani, M., Moser, P., Goetz, D., 2005. Continuum modelling of dynamic
behaviour and fragmentation of quasi-brittle materials: application to rock fragmentation
by blasting. Int. J. Numer. Anal. Methods Geomech. 29, 729–749. doi:10.1002/nag.436
Saksala, T., 2010. Damage-Viscoplastic consistency model with parabolic cap for rocks with
brittle and ductile behaviour under low-velocity impact loading. Int. J. Numer. Anal.
Methods Geomech. 34, 1362–1386. doi:10.1002/nag
Saksala, T., Hokka, M., Kuokkala, V.T., Mäkinen, J., 2013. Numerical modeling and
experimentation of dynamic Brazilian disc test on Kuru granite. Int. J. Rock Mech. Min.
Sci. 59, 128–138. doi:10.1016/j.ijrmms.2012.12.018

124
Sands, J.M., Fountzoulas, C.G., Gilde, G. a., Patel, P.J., 2009. Modelling transparent
ceramics to improve military armour. J. Eur. Ceram. Soc. 29, 261–266.
doi:10.1016/j.jeurceramsoc.2008.03.010
Schuler, H., 2007. Evaluation of an exprimental method via numerical simulation, in:
Internation Conference on Fracture Mechanics of Concrete and Concrete Structures.
Taylor & Francis, London, pp. 607–611. doi:10.1227/01.NEU.0000223495.39240.9A
Schuler, H., Mayrhofer, C., Thoma, K., 2006. Spall experiments for the measurement of the
tensile strength and fracture energy of concrete at high strain rates. Int. J. Impact Eng.
32, 1635–1650. doi:10.1016/j.ijimpeng.2005.01.010
Sharon, E., Fineberg, J., 1999. Confirming the continuum theory of dynamic brittle fracture
for fast cracks. Nature 397, 333–335. doi:10.1038/16891
Shen, L., 2009. A Rate-Dependent Damage / Decohesion Model for Simulating Glass
Fragmentation under Impact using the Material Point Method. Cmes 49, 23–45.
Shenoy, V.B., Kim, K.-S., 2003. Disorder effects in dynamic fragmentation of brittle
materials. J. Mech. Phys. Solids 51, 2023–2035. doi:10.1016/j.jmps.2003.09.010
Shockey, D. A., Curran, D.R., Seaman, L., Rosenberg, J.T., Petersen, C.F., 1974.
Fragmentation of rock under dynamic loads. Int. J. Rock Mech. Min. Sci. Geomech.
Abstr. 11, 303–317. doi:10.1016/0148-9062(74)91760-4

Suaris, W., Shah, S.P., 1984. Rate‐Sensitive Damage Theory for Brittle Solids. J. Eng. Mech.
110, 985–997. doi:10.1061/(ASCE)0733-9399(1984)110:6(985)
Taylor, L.M., Chen, E., Kuszmaul, J.S., 1986. Microcrack-Induced Damage Accummulation
in Brittle Rock Under Dynamic Loading. Comput. Methods Appl. Mech. Eng. 55, 301–
320.
Tedesco, J.W., Ross, C.A., 1998. Strain rate dependent constitutive equations for concrete. J.
Press. Vessel Technol. 120, 389–405.

Throne, B.J., 1990. A damage model for Rock fragmentation and comparision of calculation
with blasting in granite, Sandia national laboratories.

Vocialta, M., Molinari, J., 2015. Influence of internal impacts between fragments in dynamic
brittle tensile fragmentation. Int. J. Solids Struct. 58, 247–256.
Wagner, J., Holian, B.L., Voter, A.F., 1992. Molecular-dynamics simulations of two-
dimensional materials at high strain rates. Phys. Rev. A 45.
Wang, Q.Z., Li, W., Xie, H.P., 2009. Dynamic split tensile test of Flattened Brazilian Disc of
rock with SHPB setup. Mech. Mater. 41, 252–260. doi:10.1016/j.mechmat.2008.10.004
Wang, Z., Li, Y., Shen, R.F., 2007. Numerical simulation of tensile damage and blast crater
in brittle rock due to underground explosion. Int. J. Rock Mech. Min. Sci. 44, 730–738.
doi:10.1016/j.ijrmms.2006.11.004
Wang, Z., Li, Y., Wang, J.G., 2008. A method for evaluating dynamic tensile damage of
rock. Eng. Fract. Mech. 75, 2812–2825. doi:10.1016/j.engfracmech.2008.01.005
Wu, C., Shen, L., 2013. Numerical Simulation of Concrete Spalling Under impact, in:

125
Computer Methods in Applied Mechanics and Engineering. Taylor & Francis, London,
UK, pp. 781–786.
Wu, H., Zhang, Q., Huang, F., Jin, Q., 2005. Experimental and Numerical Investigation on
the Dynamic Tensile Strength of Concrete. Int. J. Impact Eng. 32, 605–617.
doi:10.1016/j.ijimpeng.2005.05.008
Yang, R., Bawden, W.F., Katsabanis, P.D., 1996. A New Constitutive Model For Blast
Damage. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 33, 245–254.
Yarin, a. L., Roisman, I. V., Weber, K., Hohler, V., 2000. Model for ballistic fragmentation
and behind-armor debris. Int. J. Impact Eng. 24, 171–201. doi:10.1016/S0734-
743X(99)00048-2
Yatom, H., Ruppin, R., 1989. Dynamic fragmentation model with internal damage. J. Appl.
Phys. 65, 112. doi:10.1063/1.343408
Yazdchi, M., Valliappan, S., Zhang, W., 1996. Continuum model for dynamic damage
evolution of anisotropic brittle materials. Int. J. Numer. Methods Eng. 39, 1555–1583.
Yew, C.H., Taylor, P.A., 1994. A Thermodynamic Theory of Dynamic Fragmentation. Int. J.
Impact Eng. 15, 385–394.
Zhang, Q.B., Zhao, J., 2013. Determination of mechanical properties and full-field strain
measurements of rock material under dynamic loads. Int. J. Rock Mech. Min. Sci. 60,
423–439. doi:10.1016/j.ijrmms.2013.01.005
Zhang, Q.B., Zhao, J., 2014. A review of dynamic experimental techniques and mechanical
behaviour of rock materials. Rock Mech. Rock Eng. 47, 1411–1478.
doi:10.1007/s00603-013-0463-y
Zhang, Y., Lu, Y., Ma, G., 2006. Investigation of dynamic response of brittle materials under
high-rate loading. Mech. Res. Commun. 33, 359–369.
doi:10.1016/j.mechrescom.2005.02.019
Zhang, Y.-Q., Lu, Y., Hao, H., 2004. Analysis of fragment size and ejection velocity at high
strain rate. Int. J. Mech. Sci. 46, 27–34. doi:10.1016/j.ijmecsci.2004.03.002
Zhou, F., Molinari, J.-F., Ramesh, K.T., 2006. Analysis of the brittle fragmentation of an
expanding ring. Comput. Mater. Sci. 37, 74–85. doi:10.1016/j.commatsci.2005.12.017
Zhou, X.Q., Hao, H., 2008. Mesoscale modelling of concrete tensile failure mechanism at
high strain rates. Comput. Struct. 86, 2013–2026. doi:10.1016/j.compstruc.2008.04.013
Zhou, Y.X., Xia, K., Li, X.B., Li, H.B., Ma, G.W., Zhao, J., Zhou, Z.L., Dai, F., 2012.
Suggested methods for determining the dynamic strength parameters and mode-I
fracture toughness of rock materials. Int. J. Rock Mech. Min. Sci. 49, 105–112.
doi:10.1016/j.ijrmms.2011.10.004
Zhou, Z., Zou, Y., Li, X., Jiang, Y., 2013. Stress evolution and failure process of Brazilian
disc under impact. J. Cent. South Univ. 20, 172–177. doi:10.1007/s11771-013-1473-3
Zhu, W.C., Tang, C.A., 2006. Numerical simulation of Brazilian disk rock failure under static
and dynamic loading. Int. J. Rock Mech. Min. Sci. 43, 236–252.

126
APPENDIX A

THE FORTRAN CODE FOR THE PROPOSED DAMAGE


MODEL

cccccccccccccccccccccccccccccccccccccccccccccc

c c

c SL+ Drucker-prager Model c

c THE UNIVERSITY OF SYDNEY c

cccccccccccccccccccccccccccccccccccccccccccccc

subroutine vumat(

C Read only -

1 nblock, ndir, nshr, nstatev, nfieldv, nprops, lanneal,

2 stepTime, totalTime, dt, cmname, coordMp, charLength,

3 props, density, strainInc, relSpinInc,

4 tempOld, stretchOld, defgradOld, fieldOld,

5 stressOld, stateOld, enerInternOld, enerInelasOld,

6 tempNew, stretchNew, defgradNew, fieldNew,

C Write only -

7 stressNew, stateNew, enerInternNew, enerInelasNew )

include 'vaba_param.inc'

127
C

C The state variables are stored as:

C All arrays dimensioned by (*) are not used in this algorithm

dimension props(nprops), density(nblock), coordMp(nblock,*),

1 charLength(nblock), strainInc(nblock,ndir+nshr),

2 relSpinInc(nblock,nshr), tempOld(nblock),

3 stretchOld(nblock,ndir+nshr),

4 defgradOld(nblock,ndir+nshr+nshr),

5 fieldOld(nblock,nfieldv), stressOld(nblock,ndir+nshr),

6 stateOld(nblock,nstatev), enerInternOld(nblock),

7 enerInelasOld(nblock), tempNew(nblock),

8 stretchNew(nblock,ndir+nshr),

8 defgradNew(nblock,ndir+nshr+nshr),

9 fieldNew(nblock,nfieldv),

1 stressNew(nblock,ndir+nshr), stateNew(nblock,nstatev),

2 enerInternNew(nblock), enerInelasNew(nblock)

character*80 cmname

integer i

real st11,st22,st33,VolSt,trace,xVR,xMnVS,xMnVSR,

1 c1,c2,c3,dcrack,crack,Dpois,f1,f2,D,dt

c..................STATE VARIABLES .......................

c state(i,1)=Strain along x at the end of time t

c state(i,2)=Strain along y at the end of time t

128
c state(i,3)=Strain along z at the end of time t

c state(i,4)=Strain along xy at the end of time t

c state(i,5)=Strain along yz at the end of time t

c state(i,6)=Strain along zx at the end of time t

c state(i,7)=Microcrack density at the end of time t

c state(i,8)=Damage at the end of time t

c state(i,9)=Element deletion

c state(i,10)=Plastic strain

c state(i,11)=Volumetric strain at the end of time t

c state(i,12)=Pressure at the end of time t

c state(i,13)=Maximum Vol strain by the end of time t

c...........................................................

c Mat constant should not have more than 6 letters

xE=props(1) !Youngs modulus

Pois=props(2) ! Poisson ratio

xcst=props(3) ! critical strain

xb=props(4) ! strain rate parameter

xalp=props(5) !other parameter

Kic=props(6) ! fracture toughness

xc=props(7) !Bulk wave speed

xden=props(8) !density

c.....DP material constants

slim0=props(9) ! Yield strength

slimp=props(10) ! Failure strength

hm=props(11) ! Hardening Modulus

sm=props(12) ! Softening Modulus

pc=props(13) ! Pressure Coefficient (Positive)

129
c

xK=xE/3.0/(1.0-2.0*Pois)

xG=xE/2.0/(1.0+Pois)

alamda=xK-2.0/3.0*xG

ep=(slimp-slim0)/hm

ef=ep+slimp/sm

xhm=hm/slim0 !normalized hardening

xsm=sm/slim0 !normalized softening

tol=1.0d-6

tol1=1.0d-3

tol2=1.0d-12

if (STEPTIME. EQ. 0.0) then !if1

do i = 1,nblock !do1

c When timeStep zero the operation is elastic

trace=strainInc(i,1)+strainInc(i,2)+strainInc(i,3)

stressNew(i,1)=stressOld(i,1)+alamda*trace+2.0*xG*strainInc(i,1)

stressNew(i,2)=stressOld(i,2)+alamda*trace+2.0*xG*strainInc(i,2)

stressNew(i,3)=stressOld(i,3)+alamda*trace+2.0*xG*strainInc(i,3)

stressNew(i,4)=stressOld(i,4)+2.0*xG*strainInc(i,4)

stressNew(i,5)=stressOld(i,5)+2.0*xG*strainInc(i,5)

stressNew(i,6)=stressOld(i,6)+2.0*xG*strainInc(i,6)

stateNew(i,1)=0.0

130
stateNew(i,2)=0.0

stateNew(i,3)=0.0

stateNew(i,4)=0.0

stateNew(i,5)=0.0

stateNew(i,6)=0.0

stateNew(i,7)=0.0

stateNew(i,8)=0.0

stateNew(i,9)=0.0

stateNew(i,10)=0.0

stateNew(i,11)=0.0

stateNew(i,12)=0.0

stateNew(i,13)=0.0

end do !do1

else !1

do 100 i = 1,nblock

St11=stateOld(i,1)+strainInc(i,1)

St22=stateOld(i,2)+strainInc(i,2)

St33=stateOld(i,3)+strainInc(i,3)

St44=stateOld(i,4)+strainInc(i,4)

St55=stateOld(i,5)+strainInc(i,5)

St66=stateOld(i,6)+strainInc(i,6)

131
c logarithmic strain

c.....xv1=logarithmic strain along x

xv1=ALOG(1.0+St11)

c.....xv2=logarithmic strain along y

xv2=ALOG(1.0+St22)

c.....xv3=logarithomic strain along z

xv3=ALOG(1.0+St33)

c Incremental strain

xv4 = ALOG(1.0+St44)

xv5 = ALOG(1.0+St55)

xv6 = ALOG(1.0+St66)

c....along x=xdv1

xdv1=ALOG((1.0+St11)/(1.0+stateOld(i,1)))

c....along y=xdv2

xdv2=ALOG((1.0+St22)/(1.0+stateOld(i,2)))

c....along z=xdv3

xdv3=ALOG((1.0+St33)/(1.0+stateOld(i,3)))

c Strain increment at each time step=trace

trace=xdv1+xdv2+xdv3

trace1=strainInc(i,1)+strainInc(i,2)+strainInc(i,3)

c Calculate the cumulative sum of volumetric strain VolSt

VolSt1=St11+St22+St33

VolSt=xv1+xv2+xv3

xpre=(stressOld(i,1)+stressOld(i,2)+stressOld(i,3))/3.0

132
c V_m=Maximum Vol strain of the history at the end of time t

V_m=amax1(VolSt,stateOld(i,13))

c...When D=1 the stress is zero in all possible scenarios


(tension+compression+loading+unloading)

If(stateOld(i,8).ge.1.0)then

stressNew(i,1)=0.0

stressNew(i,2)=0.0

stressNew(i,3)=0.0

stressNew(i,4)=0.0

stressNew(i,5)=0.0

stressNew(i,6)=0.0

St11=0.0

St22=0.0

St33=0.0

St44=0.0

St55=0.0

St66=0.0

c..Pressure=xp_end

xp_end=0.0

c..Update the state variables

stateNew(i,1)=St11

stateNew(i,2)=St22

stateNew(i,3)=St33

stateNew(i,4)=St44

133
stateNew(i,5)=St55

stateNew(i,6)=St66

stateNew(i,7)=stateOld(i,7)

stateNew(i,8)=stateOld(i,8)

c stateNew(i,9)=element deletaion

stateNew(i,10)=stateOld(i,10)

stateNew(i,11)=VolSt

stateNew(i,12)=xp_end

stateNew(i,13)=V_m

goto 100

end if

c..when the failure strength is reached in compression stress zero

c If(stateOld(i,10).gt.ep)then

c stressNew(i,1)=0.0

c stressNew(i,2)=0.0

c stressNew(i,3)=0.0

c stressNew(i,4)=0.0

c stressNew(i,5)=0.0

c stressNew(i,6)=0.0

c..Pressure=xp_end

c xp_end=0.0

c..Update the state variables

c stateNew(i,1)=St11

c stateNew(i,2)=St22

c stateNew(i,3)=St33

c stateNew(i,4)=St44

c stateNew(i,5)=St55

134
c stateNew(i,6)=St66

c stateNew(i,7)=stateOld(i,7)

c stateNew(i,8)=stateOld(i,8)

c stateNew(i,9)=element deletaion

c stateNew(i,10)=stateOld(i,10)

c stateNew(i,11)=VolSt

c stateNew(i,12)=xp_end

c stateNew(i,13)=V_m

c goto 100

c end if

c..UNLOADING FOR TENSION

If (VolSt.gt.0.0 .and.trace.le.0.0)then

trace=strainInc(i,1)+strainInc(i,2)+strainInc(i,3)

stressNew(i,1)=stressOld(i,1)+alamda*trace+2.0*xG*strainInc(i,1)

stressNew(i,2)=stressOld(i,2)+alamda*trace+2.0*xG*strainInc(i,2)

stressNew(i,3)=stressOld(i,3)+alamda*trace+2.0*xG*strainInc(i,3)

stressNew(i,4)=stressOld(i,4)+2.0*xG*strainInc(i,4)

stressNew(i,5)=stressOld(i,5)+2.0*xG*strainInc(i,5)

stressNew(i,6)=stressOld(i,6)+2.0*xG*strainInc(i,6)

c..Pressure=xp_end

xp_end=(stressNew(i,1)+stressNew(i,2)+stressNew(i,3))/3.0

c UPDATE STATE VARIABLES

stateNew(i,1)=St11

135
stateNew(i,2)=St22

stateNew(i,3)=St33

stateNew(i,4)=St44

stateNew(i,5)=St55

stateNew(i,6)=St66

stateNew(i,7)=stateOld(i,7)

stateNew(i,8)=stateOld(i,8)

c stateNew(i,9)=element deletaion

stateNew(i,10)=stateOld(i,10)

stateNew(i,11)=VolSt

stateNew(i,12)=xp_end

stateNew(i,13)=V_m

goto 100

end if

c..UNLOADING FOR COMPRESSION

If (VolSt.lt.0.0 .and.trace.ge.0.0)then

trace=strainInc(i,1)+strainInc(i,2)+strainInc(i,3)

stressNew(i,1)=stressOld(i,1)+alamda*trace+2.0*xG*strainInc(i,1)

stressNew(i,2)=stressOld(i,2)+alamda*trace+2.0*xG*strainInc(i,2)

stressNew(i,3)=stressOld(i,3)+alamda*trace+2.0*xG*strainInc(i,3)

stressNew(i,4)=stressOld(i,4)+2.0*xG*strainInc(i,4)

stressNew(i,5)=stressOld(i,5)+2.0*xG*strainInc(i,5)

stressNew(i,6)=stressOld(i,6)+2.0*xG*strainInc(i,6)

136
c... Pressure=xp_end

xp_end=(stressNew(i,1)+stressNew(i,2)+stressNew(i,3))/3.0

c UPDATE STATE VARIABLES

stateNew(i,1)=St11

stateNew(i,2)=St22

stateNew(i,3)=St33

stateNew(i,4)=St44

stateNew(i,5)=St55

stateNew(i,6)=St66

stateNew(i,7)=stateOld(i,7)

stateNew(i,8)=stateOld(i,8)

c stateNew(i,9)=element deletaion

stateNew(i,10)=stateOld(i,10)

stateNew(i,11)=VolSt

stateNew(i,12)=xp_end

stateNew(i,13)=V_m

goto 100

end if

ccccccccccccccccccccccccccccccccccccc

if (VolSt.gt.0.0) then !if2 .and. xpre.gt.0.0

ccccccccccccccccccccccccccccccccccccc

c V_pre= volumetric strain at the beginning of the step

V_pre=stateOld(i,11)

c xVR=Extensional vol strain rate

137
xVR=trace/dt

xra=strainInc(i,3)/dt

xra_lg=xdv3/dt

If(VolSt.ge.V_m)then

c.. xEVOL=effective vol strain in tension

If(V_pre.le.xcst)then

dcrack=0.0

elseif(V_pre.gt.xcst.and.VolSt1.lt.xcst)then

xEVOL=(V_pre+xcst)/2.0

dcrack=xalp*(xEVOL-xcst)**xb

else

dcrack=xalp*(VolSt-xcst)**xb

end if

else

dcrack=0.0

end if

if(dcrack.lt.0.0)then

dcrack=0.0

end if

c...Crack density "crack"

crack=stateOld(i,7)+dcrack*dt

c3=9.0/16.0

If (crack.gt.c3) then

crack=c3

138
end if

c DPois=Damaged Poisson Ratio

DPois=Pois*(1.0-(16.0/9.0)*crack)

C Damage rate = DRt

c Simplification of Eqn for damage rate. simplified with f1 and f2

f1=(1.0-DPois**2.0)/(1.0-2.0*DPois)

D=1.0/c3*f1*crack

If (D.gt.1.0) then

D=1.0

end if

c...ELEMENT DELETION CONDITION

if (D.eq.1.0) then

stateNew(i,9)=0

else

stateNew(i,9)=1

end if

c..Update the damaged material property

xKD=xK*(1.0-D)

xGD=3.0*xKD*(1.0-2.0*Dpois)/2.0/(1.0+Dpois)

c trace in secant is current strain form

trace=xv1+xv2+xv3

139
stressNew(i,1)=(xKD-2.0/3.0*xGD)*trace+

* 2.0*xGD*xv1

stressNew(i,2)=(xKD-2.0/3.0*xGD)*trace+

* 2.0*xGD*xv2

stressNew(i,3)=(xKD-2.0/3.0*xGD)*trace+

* 2.0*xGD*xv3

stressNew(i,4)=2.0*xGD*xv4

stressNew(i,5)=2.0*xGD*xv5

stressNew(i,6)=2.0*xGD*xv6

xp_end=(stressNew(i,1)+stressNew(i,2)+stressNew(i,3))/3.0

If (i.eq.3023)then

write(*,*)"@@@ Not expected ="

write(*,*)"VolSt1,trace1",VolSt1,trace1

end if

c..UPDATE STATE VARIABLES

stateNew(i,1)=St11

stateNew(i,2)=St22

stateNew(i,3)=St33

stateNew(i,4)=St44

stateNew(i,5)=St55

stateNew(i,6)=St66

stateNew(i,7)=crack

stateNew(i,8)=D

c stateNew(i,9)=element deletion

140
stateNew(i,10)=stateOld(i,10)

stateNew(i,11)=VolSt

stateNew(i,12)=xp_end

stateNew(i,13)=V_m

stateNew(i,15)=xVR

stateNew(i,16)=xra

stateNew(i,17)=xra_lg

cccccccccccccccccc

else !2

cccccccccccccccccccc

c MATERIAL AT COMPRESSION SUBROUTINE

c...Initializing variables

dpint=0

yfc=0

yf=0

yf1=1.0

iter=0

c...Plastic strains initializing

depsp1=0.0

depsp2=0.0

depsp3=0.0

depsp4=0.0

depsp5=0.0

depsp6=0.0

141
c

c... Start iteration loop

300 iter=iter+1

depse1=strainInc(i,1)-depsp1

depse2=strainInc(i,2)-depsp2

depse3=strainInc(i,3)-depsp3

depse4=strainInc(i,4)-depsp4

depse5=strainInc(i,5)-depsp5

depse6=strainInc(i,6)-depsp6

c Trial stress

trace=depse1+depse2+depse3

sig1=stressOld(i,1)+alamda*trace+2.0*xG*depse1

sig2=stressOld(i,2)+alamda*trace+2.0*xG*depse2

sig3=stressOld(i,3)+alamda*trace+2.0*xG*depse3

sig4=stressOld(i,4)+2.0*xG*depse4

sig5=stressOld(i,5)+2.0*xG*depse5

sig6=stressOld(i,6)+2.0*xG*depse6

c...sum of equivalent plastic strain "pintc"

pintc=stateOld(i,10)+dpint

c... Deviatoric part of trial stress calculated from back stress

142
p=(sig1+sig2+sig3)/3.0

ds1=sig1-p

ds2=sig2-p

ds3=sig3-p

ds4=sig4

ds5=sig5

ds6=sig6

xA=ds1*ds1+ds2*ds2+ds3*ds3+2.0*ds4*ds4+2.0*ds5*ds5+2.0*ds6*ds6

j2r = sqrt(xA)

c Equivalent stress=sigeq

sigeq=3.0*xA/2.0

pf=slim0-pc*p

If(pintc.lt.ep) then

chf=pf*(1.0+xhm*pintc)

elseif (pintc .lt. ef) then

chf=pf*(1+xhm*pintc)*(1.0-xsm*(pintc-ep))

else

chf=0.0

sig1=0.0

sig2=0.0

sig3=0.0

sig4=0.0

sig5=0.0

sig6=0.0

c write(*,*)"t=",stepTime,"iter=",iter,"point #",i,"failed"

143
goto 2000

end if

if (chf.lt.0.d0) chf=0.d0

yf=yfc

c... Check yield condition

yfc=sigeq-chf*chf

if((yfc .lt. 0.0 .and. iter .eq. 1)

* .or. (abs(yfc/yf1) .le. 1.0d-4)) goto 2000

c... Plasticity occurs

if (iter .eq. 1) yf1=yfc

c write(*,*)"t=", stepTime, "Yields at #", i , "iter=", iter

ccc...flow rule....ccccc

if(pintc.lt.ep) then

hf=1.0+xhm*pintc

else

144
hf=(1.0+xhm*ep)*(1.0-xsm*(pintc-ep))

end if

hf=2.0*pc*pf*hf*hf

temp=3.0*(27.0*xA+hf*hf)

temp=sqrt(temp)

if (abs(temp) .lt. tol) then

flow1=0.0

flow2=0.0

flow3=0.0

flow4=0.0

flow5=0.0

flow6=0.0

else

flow1=(9.0*ds1+hf)/temp

flow2=(9.0*ds2+hf)/temp

flow3=(9.0*ds3+hf)/temp

flow4=9.0*ds4/temp

flow5=9.0*ds5/temp

flow6=9.0*ds6/temp

endif

c..calculate "dlamb"

145
If(iter.eq.1)then

dlamb=1.0d-6

else

if(abs(yf) .lt. tol) then

dlamb=yf/abs(yf)*1.0d-8

elseif(yfc .gt. yf)then

dlamb=dlamb*yfc

elseif (abs((yfc-yf)/yfc) .lt. tol) then

dlamb=1.0d-8*yfc/abs(yfc)

else

dlamb=-dlamb*yfc/(yfc-yf)

end if

end if

c...calculate plastic variables

depsp1=depsp1+dlamb*flow1

depsp2=depsp2+dlamb*flow2

depsp3=depsp3+dlamb*flow3

depsp4=depsp4+dlamb*flow4

depsp5=depsp5+dlamb*flow5

depsp6=depsp6+dlamb*flow6

dpint=dpint+dlamb

146
if (iter.lt.30) goto 300

2000 continue

stressNew(i,1)=sig1

stressNew(i,2)=sig2

stressNew(i,3)=sig3

stressNew(i,4)=sig4

stressNew(i,5)=sig5

stressNew(i,6)=sig6

c..Pressure=xp_end

xp_end=(stressNew(i,1)+stressNew(i,2)+stressNew(i,3))/3.0

If(pintc.ge.ep)then

stateNew(i,14)=0

else

stateNew(i,14)=1

end if

c UPDATE STATE VARIABLES

stateNew(i,1)=St11

stateNew(i,2)=St22

stateNew(i,3)=St33

stateNew(i,4)=St44

stateNew(i,5)=St55

stateNew(i,6)=St66

147
stateNew(i,7)=stateOld(i,7)

stateNew(i,8)=stateOld(i,8)

c stateNew(i,9)=element deletaion

stateNew(i,10)=pintc

stateNew(i,11)=VolSt

stateNew(i,12)=xp_end

stateNew(i,13)=V_m

end if

100 continue

End if !1

return

end

148
APPENDIX B:

LIST OF PUBLICATION

Das, S., Shen, L. (2014). Experimental and numerical investigation of dynamic failure of
sandstone under high strain rates. 23rd Australasian Conference on the Mechanics of
Structures and Materials (ACMSM23), Lismore, NSW: Southern Cross University.

Das, S., Shen, L., Flores-Johnson, E., Gan, Y.,and Nguyen, G. D. “A rate-dependent damage
model for simulating fragmentation of brittle materials under impact loading” under
preparation to be submitted to” International Journal of Solids and Structures.

149

You might also like