Sunteți pe pagina 1din 30

Rock Mech. Rock Engng.

(2008) 41 (5), 641–670


DOI 10.1007/s00603-007-0136-9
Printed in The Netherlands

Numerical Procedure for Dynamic Simulation


of Discrete Fractures Due to Blasting
By
M. R. Saharan and H. S. Mitri

1
Central Mining Research Institute – Regional Centre, Nagpur, India
2
Department of Mining, Metals and Materials Engineering, McGill University,
Montreal, Canada
Received January 5, 2006; accepted January 3, 2007
Published online April 22, 2007 # Springer-Verlag 2007

Summary
This paper describes development of a generic nonlinear, dynamic modelling technique to simu-
late discrete rock fractures due to blasting using the finite element method. The element elimina-
tion technique together with a brittle, Rankine failure-type material model are used as a means to
simulate the initiation and growth of fractures in the rock under the effect of blast-induced
dynamic pressure pulse. Dynamic loads representing ideal and non-ideal detonations are simu-
lated and a new method, termed as optimised pressure profile, is proposed to approximate the
pressure-time profile of the blast load to model the dynamic load. Comparison of numerical model
results with previously reported observations from the literature reveals the ability of the model as
a predictive tool and supports the validity of the developed modelling procedure.
Keywords: Finite element method, dynamic analysis, nonlinear modelling, element elimination
technique, discrete fractures, rock fracturing.

List of Symbols
[B] Strain-displacement matrix
[C] Damping matrix
[C] Consistency condition crack tensor written in the crack direction coordinate system
cd Dilatational wave speed at a particular time step in an element
Cp, Cs Longitudinal and transverse shock wave velocity, respectively
[D] Elasticity matrix
E Elastic modulus
f(t) Body forces
f int and f ext Internal resistance and external force vectors
G Unit charge length (kg=m) (TNT equivalent)
[K] Time dependant stiffness matrix
[M] Mass matrix
[N] Shape function matrix
642 M. R. Saharan and H. S. Mitri

fP(t)g Nodal force vector


Pb Borehole pressure
PPV Peak particle velocity
q Explosive weight in TNT equivalent
q(t) Surface tractions
rc Coupling ratio (explosive diameter=borehole diameter)
T Transformation matrix

u, u_ , u Acceleration, velocity and displacement, respectively
VOD Velocity of detonation of an explosive
t Time step in finite difference scheme
rcr , rf, rs Radius of crushing zone, fracture zone and seismic zone, respectively
,  Mass and stiffness proportional Rayleigh damping constants, respectively
" Mechanical strain
 Stress
c Uniaxial compressive strength
 Material density
 Coefficient of adiabatic exponent
! Circular frequency of natural modes
^, ^ Effective Lame constants

1. Introduction
Drilling and blasting is the main and most economical procedure to extract valuable
mineral resources from the earth. Researchers still make different hypotheses in ex-
plaining the fundamental operative mechanisms responsible for rock fracturing by
explosive energy in spite of its prevalent use at a large scale. Enormous experimental
research efforts have been made over the last five decades to understand the rock
fracturing mechanisms. Table 1 summarises the different rock breaking mechanisms
as propounded by various authors. As can be seen, there is no agreement on a unified
theory. The differences may stem from observational difficulties associated with
experimental techniques due to the extremely short duration of the explosive shock
wave (in the order of micro-seconds), a very fast fracturing process covered under
explosive gaseous products and rock debris (fracture propagation speed up to one third
of the shear wave speed in the rock), as well as the heterogeneous nature of the rock.
Moreover, the prevalent experimental techniques developed so far do not have full
control on the experiments. Continuous efforts are being made to develop better ex-
perimental techniques in order to understand the fundamental operative rock fractur-
ing mechanisms. Such an improvement will be a great help in the development of
better explosive products as well as safer and more economical blast designs.
Rapid advances made with the numerical modelling tools and the availability of
powerful computational resources at affordable cost have made numerical simulation
nowadays a most promising tool to study the dynamic rock fracturing processes. The
use of numerical simulation for the dynamic rock fracturing is appealing, essential and
most suitable due to the large number of complex variables involved.
A numerical procedure is established in this study, which encapsulates the relevant
phenomena related to the rock fracturing due to explosive energy. The dynamic load
pulse used in this study considers the role of shock wave in rock fracturing, while the
pressure exerted by gas energy is indirectly considered in the process of the element
elimination technique (EET) using the finite element method. Experimental results
indicate that shock waves are the primary cause of rock fracturing and the gas pressure
Numerical Procedure for Dynamic Simulation 643

Table 1. Blasting theories and their breakage mechanisms (revised after Anon, 1987)

Reference Breakage mechanism


( added references)
Tensile Compressional Gas Flexural Nuclie
reflected stress waves pressure rupture stress
waves flow

Obert and Duval (1949) X


Hino (1956) X
Duval and Atchison (1957) X
Rinehart (1958) X
Langefors and Kihlstrom (1963) X X
Starfield (1966) X
Porter and Fairhurst (1970) X X
Persson et al. (1970) X
Kutter and Fairhurst (1971) X X
Field and Ladegarrd-Pederson (1971) X X
Johansson and Persson (1972) X X
Lang and Favreau (1972) X X X X
Ash (1973) X X
 Bhandari and Vutukuri (1974) X X X
Hagan and Just (1974) X
Barker et al. (1978) X
Winzer et al. (1983) X
Margolin and Adams et al. (1983) X
 McHugh (1983) X
 Brinkmann (1987) X
 Daehnke et al. (1996) X
 Nie and Olsson (2000) X

energy has more roles in the fragmentation process (Brinkmann, 1990; Nie and
Olsson, 2000). The role of thermal energy in fracturing is not considered in the current
study due to the fact that the expansion of explosive gaseous products takes place
adiabatically (Hustrulid, 1999). The discrete fracture network growth in the developed
procedure is simulated without introducing geometric or material imperfections. Char-
acteristic models of blast simulation as an explosion in an unbounded thick plate are
presented in this paper, while considering two broad explosive categories of pressure
pulse, namely, ideal- and non-ideal detonation.

2. Basics of Dynamic Modelling


Dynamic simulation with any numerical method essentially involves the solution of
the equations of motion. In a finite element formulation, these can be written as
(Hughes, 2000)
½Mf€
ug þ ½Cfu_ g þ ½Kfug ¼ fPðtÞg ð1Þ
where, u ¼ displacement, u_ ¼ velocity, and €
u ¼ acceleration.
The mass matrix
ð
½M ¼ ½NT ½NdV ð2Þ
v

In the above,  is the material density and [N] is the shape function matrix.
644 M. R. Saharan and H. S. Mitri

The damping matrix, [C], for the Rayleigh damping is in the form
½C ¼ ½M þ ½K ð3Þ
 and  are pre-defined constants.
The time-dependant stiffness matrix is defined as
ð
½K ¼ BT ½D½BdV ð4Þ
v

[D] and [B] represent the constitutive and strain-displacement matrices, respectively.
The nodal force vector due to surface tractions, [q(t)], and body forces, [f(t)], is
given by
ð ð
fPðtÞg ¼ ½NT qðtÞS þ ½NT f ðtÞdV ð5Þ
s v

Equation (1) can be rewritten in the following form,


ug ¼ ff ext g  ff int g
½Mf€ ð6Þ

where, ff ext g ¼ fPðtÞg; and f int ¼ ½Kfug þ ½Cfu_ g.


Equations (1) and (6) can be solved either by explicit methods (the most popular
one is the central difference method) or by implicit methods (such as, Hubolt algo-
rithm and Nemark algorithm). Solution by explicit methods requires finding an ini-
tial solution for set of algebraic equations. No global matrix formulation is needed in
the explicit methods and no iterations are performed. In contrast, implicit methods
require the solution of system matrices, which sometimes can become computation-
ally costly. The most efficient explicit dynamic analysis procedure is the second
order central difference operator, which is based on the implementation of an ex-
plicit integration rule together with the use of diagonal or lumped element mass
matrix. Half time step values are used to calculate velocity and acceleration. Both
values are accurate to an order of t2 (time step square). Values of the derivatives
at the centre of a time interval are obtained from the difference in the function values
at the ends of the interval. The central difference integration operator is explicit in
1
that the kinematic state may be advanced using known values of u_ ði2Þ and €uðiÞ
from the previous time increment. A key to the computational efficiency of the
explicit solution procedure is the use of diagonal element mass matrices because
the inversion of such matrices, which is required in the beginning of each time
increment, makes the solution of algebraic equations very simple. Thus, Eq. (6)
becomes
ug ¼ ½M1 ðff ext g  ff int gÞ
f€ ð7Þ

The explicit procedure integrates through time by using many small increments. The
central difference operator is conditionally stable, and stability limit for the operator
(with no damping) is given in terms of the highest eigen value in the system as,
2
t  ð8Þ
!max
where ! is the circular frequency of the natural modes.
Numerical Procedure for Dynamic Simulation 645

The current dilatational wave speed in the element is calculated with the following
expression
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ð^ þ 2^Þ
cd ¼ ð9Þ


where ^ and ^ are effective Lame constants and  is the material’s mass density.
Details of the integration procedure are given in finite element textbooks (e.g. see,
Hughes, 2000).

3. Description of Numerical Model


Plane stress modelling provides a convenient means to understand fundamental studies
for the dynamic rock fracturing processes because the results can be directly compared
with observed fracturing that takes place on the free surface of the blasthole. Blast
loading is essentially radial to the blasthole axis and hence it would be planar in a
2-dimensional section. In the past, laboratory scale studies were undertaken using this
plane stress concept by conducting experiments on thick plates (e.g., see Kutter
and Fairhurst, 1971; Fourney et al., 1993; Jung et al., 2001). Blasting in the present
context, therefore, will be simulated as a problem of explosion from a central borehole
in an unbounded thick plate. Single borehole in an unbounded medium is chosen to
simplify the case under investigation and in order to preclude extraneous impact of the
blasting waves on the fracturing process resulting from the interaction of either more
than one hole or reflection from free faces. A 38 mm diameter blasthole is selected for
the numerical simulations, which commensurate with common blasting practice of
many mining operations.

3.1 Model Domain


Accurate numerical representation of unbounded media has always been a concern in
dynamic loading problems. Many strategies have been proposed such as the use of
infinite elements proposed by Lysmer and Kuhlemeyer (1969). Application of infinite
elements has some pitfalls and therefore is not considered in the present case. These
limitations are discussed in a subsequent section. Another commonly used arrange-
ment, which is adopted in the current study, is restricting displacements perpendicular
to the model boundaries by placing rollers. It can be shown by elastic analytical
solutions that the radius of influence of any underground opening is approximately
six times the opening size (e.g. see Obert and Duvall, 1967). Roller boundaries placed
at a distance of six to ten times the opening size from the opening should not affect the
results of geostatic problems. In dynamic modelling such distance is not sufficient due
to wave propagation and reflection. Therefore, it has been decided to place the far-field
boundaries at a distance of approximately 250 times the opening size. Such distance
will ensure a) avoidance of spurious wave reflections from boundaries and their
subsequent deleterious impact on the final model behaviour, and b) allocation of
sufficient analysis time for fracture propagation due to transient blast wave loadings.
646 M. R. Saharan and H. S. Mitri

Fig. 1. Plane stress model layout with a central 38 mm diameter blasthole

Full model domains are simulated and model boundary truncation due to sym-
metry are not considered. A schematic of the numerical model geometry is shown
in Fig. 1.

3.2 Material Model


The tensile (extension) failure mode is considered as the basic rock failure mechanism
(Batzle et al., 1980; Blair and Cook, 1998; Kranz, 1983). Fracture formation in a rock
material subjected to direct tension is related to the breaking of molecular bonds ahead
of the advancing fractures. Fractures generated under compressive stress states are
observed to be also due to extension in nature and follow the direction of maximum
compression. The extension and coalescence of these cracks lead to macroscopic fail-
ure pattern. It is a known fact that rock brittle failure due to blasting will be governed
by its tensile strength since the tensile strength of the rock is much lower than its
compressive strength. Exceeding the tensile strength near the blasthole periphery is
overwhelming and that leads to a crushing zone. Beyond the crushing zone, blasting
results in the formation of a discrete fracture networks. A detailed discussion on the
rock fracturing mechanism due to transient dynamic loading can be found elsewhere
(e.g., Saharan et al., 2006).
ABAQUS, a general-purpose finite element modelling commercial software, pro-
vides a brittle cracking model to simulate brittle rock failure (ABAQUS, 2003). This
material model is selected for the present development as it fulfills the tensile failure
representation requirements mentioned above. The model considers anisotropy induced
by cracking and ignores hardening and subsequent softening characteristics of com-
pressive failure. In compression, the model assumes elastic behaviour. The ABAQUS
cracking model assumes fixed, orthogonal cracks with the maximum number of cracks
at a point limited by the number of direct stress components present at that point.
Internally, once cracks exist at a point, the component forms of all vector- and tensor-
valued quantities are rotated so that they lie in the local system defined by the crack
orientation vectors. In physical sense, macro-cracks or fractures are obtained in the
numerical procedure with the use of element elimination technique. While brief
Numerical Procedure for Dynamic Simulation 647

Fig. 2. The Rankine failure criterion in the plane stress space

mathematical expressions are given below, more detailed information of the brittle
cracking model can be obtained from ABAQUS manual (ABAQUS, 2003),
(a) Crack detection
A simple Rankine criterion is used to detect crack initiation. This states that a
crack forms when the maximum principal tensile stress exceeds the tensile strength of
the material. The Rankine crack detection surface in the plane stress space is shown in
Fig. 2.
(b) Element strain
This is defined as follows:
de ¼ deel þ deck ð10Þ
where de is the total mechanical strain, deel is the elastic strain representing the
uncracked rock, and deck is the cracking strain associated with any existing cracks.
The intact continuum between the cracks is modelled with isotropic, linear elasticity.
(c) Constitutive equation
ds ¼ Del ðde  Tdeck Þ ð11Þ
el
where ds is the stress increment, D is the isotropic linear elasticity matrix and T is a
transformation matrix constructed from the direction cosines of the local coordinate
system. T is constant in the present fixed crack model.
(d) Consistency condition
A consistency condition for cracking (analogous to the yield condition in classical
plasticity) written in the crack direction coordinate system has the following tensor
form:
½C ¼ Cðt; sI; II Þ ¼ 0 ð12Þ
where [C] ¼ [Cnn Ctt Css Cnt Cns Cts]T and sI,II represents a tension softening
model (Mode I fracture) in the case of the direct components of stress and a shear
648 M. R. Saharan and H. S. Mitri

softening=retention model (Mode II fracture) in the case of the shear components of


stress. The matrices @C=dt and @C=@sI,II are assumed to be diagonal, implying that
there is no coupling between cracks in the cracking conditions. Two cracking states

Fig. 3. Schematic representation of the functioning of the EET approach using a non-linear brittle failure
model
Numerical Procedure for Dynamic Simulation 649

Table 2. Typical granite rock properties used for the numerical modelling

Rock property Value Source reference Remarks

Density (), kg=m3 2650.0 Diederichs (1999) Tested and compiled


Young’s modulus (E), GPa 60.0 representative values
Poisson’s ratio () 0.24
Tensile strength (UTS), MPa 15.0
Ultimate crack opening 1  105 Dawding et al. (1985) Tested value
displacement (COD), m

are possible: actively opening crack state and closing=reopening crack state. The brit-
tle cracking model is characterized by a stress-displacement response rather than a
stress-strain response. This characterization is based on Hilleborg et al. (1976) fracture
energy proposal to avoid unreasonable mesh-sensitive numerical results. Thus, crack
opening displacement is selected as a criterion for element elimination from the
model.
The rock material chosen for the present study is granite, whose properties are well
tested and known. Although crack detection in the present development is based
purely on Mode I fracture considerations due to crack detection by the Rankine failure
criterion, ensuing cracking behaviour includes both Mode I (tension softening) and
Mode II (shear softening=retention) behaviour.
The cracked element is removed from the calculations after an amount of crack
opening displacement is attained at which it can no longer sustain stresses. To
eliminate an element, all components of stress tensors in this element are set to
zero. As a result, all forces in this element become zero as well and therefore, this
element stops to transmit load to neighbouring non-eliminated elements. In order to
avoid numerical problems related to strong local loss of equilibrium, the stress is set
equal to zero in several relaxation steps. The modulus of elasticity in the eliminated
elements is set to zero in the last relaxation step (Mishnaevsky and Schmauder,
2001). A removed element represents a macro crack or fracture in the context of
the present development. Figure 3 provide information about logical steps used
by the EET using the brittle failure material model and Table 2 gives a summary
of the rock properties used for the dynamic simulation as well as reference of
these values.

3.3 Numerical Solution Procedure


Explicit method is ideally suited for the dynamic simulation of rock fracturing for the
following reasons:
– The duration of dynamic load is short – in the order of milliseconds.
– Small time steps ensure numerical accuracy.
– Explicit method does not involve storage and solution of large matrices.
– No iteration procedure is involved.
The explicit dynamic analysis procedure is based on the implementation of an explicit
integration rule together with the use of diagonal or lumped element mass matrices.
650 M. R. Saharan and H. S. Mitri

Velocity components are calculated using the explicit central difference integration
rule as follows:
1 1 tðiþ1Þ þ tðiÞ ðiÞ
u_ ðiþ2Þ ¼ u_ ði2Þ þ €u
2 ð13Þ
1
ðiþ1Þ
u_ ¼ u_ ðiÞ þ tðiþ1Þ €
uðiþ2Þ
where u_ and €u are velocity and
 acceleration,
  respectively.
 The superscript (i) refers to
the increment number and i  12 and i þ 12 refer to mid-increment values. The
central difference integration operator
 isexplicit in that the kinematic state may be
advanced using known values of u_ i  12 and € u from the previous increment. Details
of the integration procedures are introduced in the theory manual of ABAQUS (2003).

3.4 Dynamic Load


There are two broad categories for detonation in rocks, namely, ideal- and non-ideal
detonation; see Fig. 4. The ideal detonation profile corresponds to emulsion-type
explosive where the peak pressure rise time is very short and the post peak pressure
drop is steep. The non-ideal detonation profile corresponds to ANFO type explosive
where the rise time for the peak blasthole pressure is longer and the post peak pressure
drop is much slower than emulsion type explosive. Numerically, these blasthole pres-
sure profiles can be approximated by one of three procedures namely, 1) Equation-of-
State (EOS), 2) pressure-decay functions or 3) direct input of dynamic pressure as a
function of time; see Table 3.
The EOS describes material behaviour in the high-rate intense pressure environ-
ment and the equation relates different material quantities as a single function regard-
less of prior history of deformation. Several EOS have been developed and are in use
(Braithwaite et al., 1996) but the EOS by Jones, Wilkins and Lee (JWL equation) is
the most popular in geomechanics due to its simple form, experimental basis and ease
of calculations (He et al., 2002; Etoh et al., 2002). The JWL equation-of-state contains
the parameters describing the relationships between volume, energy and pressure of

Fig. 4. Pressure pulse shapes for two categories of detonation (Aimone, 1992; Olsson et al., 2001)
Numerical Procedure for Dynamic Simulation 651

Table 3. Definition of blast pressure-time profile for dynamic modelling

Method Equation Example reference


       
John-Wilkinson- ! 0 ! 0 !2 Liu (1997)
Lee (JWL) P¼A 1 e R1  þ B 1  e R2  þ Em0
R1 0 R2 0 0
Pressure decay P ¼ P0 ½et  et  Cho et al. (2003)
functions P ¼ P0 ½et  et  Lima et al. (2002)
P ¼ P0 et Kutter (1967)
Direct input Gaussian function Donze et al. (1997)
of pressure-time Triangular load shape Valliapan et al.
profile (1983)
Optimized pressure profile Proposed

detonation. While the JWL EOS takes into account the rock-explosive interaction,
estimating the correct parameters for non-ideal detonation or in softer rocks poses con-
siderable difficulties (Liu, 1997). The detonation process in most rocks is non-ideal,
thus, the accuracy of using the JWL EOS is questionable.
Input in the form of decay functions is also reported (Duvall, 1953; Jung et al.,
2001; Lima et al., 2002; Olatidoye et al., 1998; Robertson et al., 1994). These equa-
tions are used to replicate the exact waveform, but they require assumptions of some
parameters whose physical significance is unknown.
The use of Gaussian function and triangular load function has also been attempted
to approximate measured dynamic-pulse load. These procedures, however, are not
close to the physical characteristics of the dynamic load and hence carry no physical
meaning. The Gaussian function is mainly introduced to avoid numerical errors asso-
ciated with the application of very high pressure, which is in the order of GPa, in a
very short duration in the order of microseconds.
A new and simple method is proposed by the authors and is termed herein as
‘‘optimized pressure-time profile’’. In this method, the peak borehole pressure and the
pressure at different time scales are required to specify the optimized pressure-time
profile. The peak borehole pressure can be estimated from the following equation
(Atlas, 1987; Clark, 1987; Nie and Olsson, 2000).
Bore Hole Pressure,
 
VOD2  2 
Pb ¼  rc ; Pa ð14Þ
8
where,
 ¼ explosive density, kg=m3
VOD ¼ Velocity of detonation, m=s
explosive diameter
rc ¼ coupling ratio ¼
bore hole diameter
 ¼ adiabatic exponent ¼ 1:5 ðPersson et al:; 1994Þ:
Values of detonation velocity (VOD), explosive density () and coupling ratio (rc, ratio
of charge diameter to borehole diameter) are required for estimating the peak borehole
pressure. Table 4 provides values of VOD and  along with the source information.
652 M. R. Saharan and H. S. Mitri

Table 4. Details of explosive properties for borehole pressure calculations

Borehole Emulsion explosive ANFO explosive


diameter, mm
VOD, m=s , kg=m3 VOD, m=s , kg=m3

38 4700 1050 3300 950


Source reference Sun et al. (2001) Brinkmann (1990) Mohanty (2003) Brinkmann (1990)

According to Eq. (14), the ideal detonation and non-ideal detonation in the present
case will have 2.9 and 1.29 GPa peak borehole pressure, respectively. The pressure-
time profile is constructed with the help of peak pressure value and by applying it over
a time period in a magnitude and a manner consistent with the explosive character-
istics. The complete process and basis of its construction is briefly described below.
The explosion pressure in a borehole decays to a stand-off pressure within a couple
of milliseconds. The stand-off pressure is much smaller than the tensile strength of the
rock. The rise time of explosion pressure to its peak is very short and varies primarily
according to the explosive characteristics and secondly according to the blasthole
diameter, blasthole confinement, rock strength, etc. Generally, the rise time for the
ideal detonation in 38 mm diameter hard brittle rocks is around 25 and 100 ms for the
non-ideal detonation characteristics (e.g. see Jung et al., 2001). The subsequent decay
in the peak borehole pressure down to the stand-off pressure is sharp in the case of the
ideal detonation and slow for the non-ideal detonation. Therefore, to construct pres-
sure-time profile, the peak borehole pressure is fully applied at the respective rise time
of the particular explosive characteristics which increases from zero at time zero. The
peak pressure magnitude drops by 90, 99 and 99.9% over a time period consistent with
the two types of detonation characteristics as shown in Fig. 5. The optimized pressure-
time profile constructed in this way is in good agreement with the observed profiles by
Frantzos (1989), Fourney et al. (1993), Jung et al. (2001) and Daniel (2003). As can be

Fig. 5. Optimized pressure-time profiles of different explosive characteristics simulated in the study for a
38 mm diameter blasthole
Numerical Procedure for Dynamic Simulation 653

Fig. 6. Comparison of different methods for the approximation of blast pressure-time profile

seen from the comparison of Fig. 5, the optimized pressure-time profile better approx-
imates the real pressure pulse than the profiles obtained by Gaussian and triangular
shape functions.

3.5 Model Boundary Conditions


There are two primary responses of any dynamic loading out of a real blasting problem.
First, the rock undergoes a crushing and fracturing process in response to the dynamic
loading. Most of the wave energy dissipates in this fracturing process. The remain-
ing wave energy continues to travel through the rock domain until it hits a free face
where it is partly reflected and partly transmitted to the air causing air over-pressure
and noise.
The majority of the dynamic modelling analyses are done with elastic material
models. This representation leads to two major concerns from a numerical simulation
point of view, namely, suitable arrangements at far-field boundaries to represent
unbounded medium and damping of the wave energy to spatially decrease it in the
model domain in a similar fashion to the physical damping. Otherwise oscillations
would occur unabatedly in the elastic medium causing deleterious effects on the final
results. These two concerns are addressed here in detail in order to establish a valid
and representative numerical modelling procedure.
(a) Representation of the far-field boundaries
One of the major concerns of any numerical simulation involving wave propaga-
tion in solids of unbounded medium is to avoid spurious reflection of waves at the
model boundaries. Several schemes have been suggested, developed and implemented
but all have their own limitations in absorbing the outgoing waves. One of the popular
arrangements is the use of infinite elements to represent the unbounded medium such
as developed by Lysmer and Kuhlemeyer (1969). These infinite elements in fact
654 M. R. Saharan and H. S. Mitri

represent infinitesimal dashpots oriented normal and tangential to the boundary. This
arrangement works well provided the waves impinge orthogonally on the boundary.
This means that the boundary should be far enough from the wave source. The wave
source should also behave like a point source in this case. Further, infinite elements
provided by Lysmer and Kuhlemeyer (1969) and adopted in ABAQUS (2003) absorb
almost all of the energy of P (primary) and S (secondary) waves but they are less
efficient in absorbing R (Rayleigh) wave energy (Ramshaw et al., 1998). These infinite
elements therefore represent ‘quiet boundaries’, and not ‘silent boundaries’. The im-
petus set in the current development is the simulation of an explosive column charge
which is in contrast to the point excitation charge and its effect on the near borehole
fracturing process. So it will not be practical to construct computationally unmanage-
able large model domain to represent point source for the waves reaching at the
boundary. The infinite elements, therefore, are not suitable and an alternate arrange-
ment needs to be considered.
Another suggested approach is the insertion of a time-damping term in the wave
equations and an attenuation boundary zone around the discretization mesh. The wave
field goes to zero with time when passing through the attenuation zone. Bing and
Greenhalgh (1998) and Sochacki et al. (1987) studied five kinds of such damper
(linear, power, cubic, Gaussian and exponential) for 2D acoustic and elastic modelling
in the time domain and concluded that the artificial reflections are best reduced by a
linear damper. The Bing and Greenhalgh (1998) study indicates that a larger mesh and
a wider absorbing zone are required for effective results from this linear time damper.
Further, the use of different type of element formulations, one for the solid medium
and another for the wave-absorbing zone, may pose numerical instability problems
during the dynamic analysis due to mesh incompatibility issues. Therefore, this meth-
od is not considered for the present development.
The present development is considering a discrete simulation of fracturing process
around the blasthole. This means that the model itself should absorb the majority of
the energy as it happens in nature. So, prevalent roller boundaries should be accurate
enough to represent the far-field boundary conditions. These boundaries however
should be far enough and far greater than the established norms for static analysis
in order to avoid any effect on the fracturing zone after wave reflection while provid-
ing enough time for the fracture propagation. This can be made possible by consider-
ing full model domains and no truncation of model size due to symmetry planes. As
per the rock material properties (see Table 2) and Eq. (9), the estimated maximum
primary compressional wave speed is 5166 m=s. Also, as per the following relations
by Mosinets and Garbacheva (1972), the likely maximum radius of the fracturing zone
is 1.0 m in the present numerical analyses.
sffiffiffiffiffiffi
Cs p ffiffiffi
Crushed zone radius; rcr ¼  3q ð15Þ
Cp
rffiffiffiffiffiffi
Cp p ffiffiffi
Fracture zone radius; rf ¼  3q ð16Þ
Cs
pffiffiffiffiffiffi
Cp p ffiffiffi
Seismic zone radius; rs ¼  3q ð17Þ
10
Numerical Procedure for Dynamic Simulation 655

where,
q ¼ explosive weight in TNT equivalent, kg
Cp ¼ longitudinal shock wave velocity, m=s
Cs ¼ transverse shock wave velocity, m=s.
Now considering a point 1.0 m away from the borehole centre and far-field boundaries
placed at 10 m away from the borehole boundary, the fastest wave will take approxi-
mately 3.5 ms with a wave speed of 5166 m=s to reach the point under consideration
after reflection from the far-field boundaries. This time is long enough to study the
fracturing process by the primary waves as the waves have peak pressures at 25 and
100 ms, respectively for the two detonation characteristics. So, far-field roller bound-
aries can be placed at 10 m away from the central borehole of 38 mm diameter and this
distance would be equivalent to a more than 250 times the opening size. Furthermore,
a numerical solution time of about 2 ms will provide enough time to capture the
fracturing process by the primary waves and hence it will be used.
(b) Damping of wave energy
There are two reasons for adding damping to a numerical model: 1) to limit
numerical oscillations and 2) to add physical damping to the system (ABAQUS,
2003). ABAQUS has provisions of bulk viscosity in linear and quadratic forms to
limit numerical oscillations. The suggested values for bulk viscosity may be re-
tained during the analysis. Choosing a stiff element having a single integration point
can be the other source of damping to the numerical models to limit numerical
oscillations.
The use of Rayleigh damping with elastic material models is a common practice
in simulating geomechanics applications to simulate the effect of the physical damp-
ing. The damping coefficients are selected either as per the PPV (peak particle veloc-
ity) measurements from specific site data or assign 1 to 15% values (Biggs, 1964;
Massarsch, 1992). It is also suggested that the damping coefficient should not be more
than 0.5% if a linear failure criterion like Mohr-Coulomb is used (Itasca, 2001). This
is because the linear failure criterion itself induces early failure or plasticity. Also, in
rock and soil, the natural damping is independent of frequency (Gemant and Jackson,
1937) so the use of frequency dependent Rayleigh damping may not be justifiable in
such analyses. Another problem reported with the frequency dependent damping is
misrepresentative results due to sometimes over damping of the low frequency modes
(Metzger, 2003), which usually control the solution.
Other types of damping procedures are also reported, such as mass proportional
damping (Bing and Greenhalch, 1998; Takewaki, 2000), dynamic relaxation damping
(Metzger, 2003), etc., but the validity of these methods for hysteretic damping me-
dium such as rock is not well established. Also, such damping arrangements are mass
proportional rather than stiffness proportional. A physical system like rock has natural
frequencies, which are beyond the effective range of the mass proportional damping
methods. Stiffness proportional damping is much needed either in the form of plas-
ticity or by specifying stiffness proportional Rayleigh damping parameter for such
systems. Therefore, arrangements involving mass proportional damping will not be
incorporated in the present development.
656 M. R. Saharan and H. S. Mitri

It is reported (ABAQUS, 2003) that the energy absorbed by plasticity is signifi-


cantly higher than it can be absorbed by any artificial damping method like the
Rayleigh damping. The proposed development aims to capture the rock fracturing,
which is the result of plasticity, on a real time frame as it happens in reality so no
artificial damping needs to be considered in the current development. Failed elements,
estimated by the Rankine failure criterion, can be removed from the calculations to
give the effect of crack=fractures generation and propagation as a source of plasticity
in the analyses (Refer to Fig. 3 for the logistic steps).
It can be concluded from the above discussion that no artificial damping as a
material property is needed if the material model considers plasticity. Therefore,
the application of artificial material damping is ruled out with the current numerical
procedure.

3.6 Element Type


One requirement for modelling rock fracturing using the finite element method is that
the element should provide a rich enough set of possible fracture paths. Such require-
ment can be met in the triangular and tetrahedron elements, for 2-dimensional and
3-dimensional analyses, respectively. Kikuchi (1983) demonstrated the superiority of
constant stress-strain triangular element (CST) over the 4-node isoparametric quad-
rilateral element with reduced integration scheme in 2-dimensional analysis of plas-
ticity problems. Therefore, the CST plane stress element is adopted for the current
study. This element has constant stress and hence it uses a single integration point for
stiffness calculation. A lumped mass matrix is used for the element with the total mass
divided equally over the nodes. Boundary tractions are integrated with two points along
the element edge.
A displacement convergence study is required to determine a suitable element
size. The area of interest is the one, which surrounds the blasthole where fracturing
takes place. Therefore, suitable size elements are required for this zone, shown in
Fig. 7, while the rest of the model domain could have larger size elements for

Fig. 7. Plane stress model for the displacement convergence analysis


Numerical Procedure for Dynamic Simulation 657

Fig. 8. Displacement convergence analysis results

Table 5. Results of displacement convergence analysis

No. of elements Side length of borehole Radial displacement (mm)


skin elements

602 21.75 0.496


1676 11.28 0.619
2588 8.5 0.685
4422 4.59 0.713
4560 4.59 0.714
5156 4.59 0.714
5784 4.25 0.715
11474 1.94 0.718

computational efficiency. A displacement convergence study is carried out with static


analysis with due consideration of elastic material properties and a static borehole
pressure of 2 GPa. The results of displacement convergence analysis are plotted in
Fig. 8 and the numerical values are recorded in Table 5. As can be seen, an element
size of 4.59 mm is appropriate in the present case. The suitability of such element size
based on static convergence analysis for use in dynamic analysis should also be
evaluated based on low frequency modes. According to the nomogram provided by
Valliapan et al. (1983), a maximum element size of 6.25 mm is required for the critical
dynamic pulse of ideal detonation characteristics with the rock material properties as
shown in Table 2. Therefore, the selected element size of 4.59 mm is equally suitable
for the low frequency modes. Furthermore, at least 10 nodal points are required for
the fastest travelling wave amplitude for its smooth propagation through the model
(Ramshaw et al., 1998). This requires a maximum element size of 4.7 mm given
that the primary compressional wave speed is 5166 m=s. Thus, an element size of
4.59 mm as concluded by the static convergence analysis is also appropriate for the
dynamic analysis.
658 M. R. Saharan and H. S. Mitri

4. Analysis of Results
The dynamic model set-up above now needs to be validated. It is difficult to compare
numerical model results with an exact fracturing pattern because such a precise knowl-
edge base does not exist. Prevalent experimental techniques have observational difficul-
ties due to the reasons explained in the introduction section. Therefore, the dynamic
model results will be compared with empirical observations reported in the literature.
Figure 9 shows the finite element mesh generated for the dynamic simulation; it
consists of 4560 CST elements and 2312 nodes. A zoom-in view of the model showing
the inner zone of interest around the borehole is displayed in Fig. 10. The central
borehole is subjected to peak borehole pressures of 2.9 and 1.29 GPa to represent ideal

Fig. 9. FE mesh used for the dynamic model

Fig. 10. Zoom-in view of the FE mesh showing the 38 mm diameter blasthole
Numerical Procedure for Dynamic Simulation 659

and non-ideal detonation, respectively, using the optimized pressure-time profile de-
picted in Fig. 5. The material model has the values as shown in Table 1 and the model is
solved using an explicit integration scheme with the dynamic load duration of 2 ms.
The numerical model results for the two types of dynamic load are shown in
Figs. 11 through 17. These results bring out a number of important phenomena, which
are discussed below.
Strain rate dependant rock response
As can be observed from Fig. 11, the stress required to open the first crack with a
high shock load of the ideal detonation is 600 MPa while only 159 MPa is required
to open the first crack in the case of non-ideal detonation. Several researchers have
argued in favour of strain rate dependant rock properties (e.g., Prasad et al., 2000). It is
interesting to note that in the present analysis, the material behaviour, which remains
essentially elastic throughout the calculations as well as the brittle failure law using
the element elimination technique are both rate independent. The element elimination

Fig. 11. Stress magnitudes associated with fracture initiation. (a) Ideal detonation – peak load ¼ 2.9 GPa.
(b) Non-ideal detonation – peak load ¼ 1.29 GPa
660 M. R. Saharan and H. S. Mitri

technique in conjunction with inertia effects, endows the material with attributes that
ultimately account for the ability to capture the load rate effects accurately.
Characteristics of fracturing zone
Comparison of model results between ideal and non-ideal detonations are shown
in Fig. 12 for the fracture pattern at the moment of the application of the peak load. It
has been well observed and documented that ideal detonation leads to more crushing
around the borehole, which follows a large numbers of short length radial cracks (e.g.,

Fig. 12. Fracture pattern at the moment of peak load application. (a) Ideal detonation, time ¼ 1 ms. (b) Non-
ideal detonation, time ¼ 817 ms
Numerical Procedure for Dynamic Simulation 661

McHugh and Keough, 1982). In contrast to this, the non-ideal detonation results in
a smaller crushing zone followed by a few long radial cracks. As can be seen from
Fig. 12, the model results are in good agreement with the reported fracturing behav-
iour. Similar trends are observed for the final fracturing pattern, shown in Fig. 13, at
the end of the blast pressure pulse.
Extent of fracturing zone
Figure 14 presents the extent of fracturing zones for the two detonation types. As
can be seen, the numerical model predicts fracturing zones having radii of 487 and

Fig. 13. Final fracturing pattern. (a) Ideal detonation. (b) Non-ideal detonation
662 M. R. Saharan and H. S. Mitri

341 mm for the ideal and non-ideal detonation, respectively. Mosinets and Garbacheva
(1972) and Kexin (1995) provide empirical relations to predict the extent of crushing
and fracturing zone. Kexin (1995) provides the following relation
 18
G 1
fractures zone radius; r ¼ 96 ð10EÞ6 ð19Þ
10c
where, r is the fractures zone radius (mm), c is the uniaxial compressive strength
(MPa), E is the elastic modulus (MPa), G is the unit charge length (kg=m) (TNT

Fig. 14. Fracturing pattern at the end of solution time (2 ms). (a) Ideal detonation. (b) Non-ideal detonation
Numerical Procedure for Dynamic Simulation 663

equivalent). Using G ¼ 1.435 and 1.13 kg=m (for respective explosive characteristics),
c ¼ 200 MPa and E ¼ 60  103 MPa, Eq. (19) predicts 356 and 348 mm for the
respective quantities thus showing close agreement with the numerical model pre-
dictions. It is noteworthy that the present modelling procedure makes it easy to dis-
tinguish between the crushing zone and the fracturing zone. It is also important to

Fig. 15. Kinetic, internal and artificial energy plots. (a) Ideal detonation, PPV at 0.5 m from the borehole
(point coordinates ¼ 0.5, 0.0). (b) Non-ideal detonation, PPV at 0.25 m from the borehole (point coor-
dinates ¼ 0.25, 0.0)
664 M. R. Saharan and H. S. Mitri

mention that in the present modelling procedure, fractures are neither pre-placed in the
model domain nor their growth paths are pre-specified.
Energy utilisation in fracturing
Numerically, the kinetic energy, which is provided by the explosive energy, in the
case of the ideal detonation drops from the peak value of 14.994 to 14.00 kJ (Fig. 15a)
while it drops from 5.97 to 5.80 kJ (Fig. 15b) in the case of the non-ideal detonation.
The consumption of kinetic energy amounts to 6.63% in the case of ideal detonation
and 2.85% in the case of non-ideal detonation. The fracturing process is the only
source of energy absorption in the numerical models as no artificial damping is used.
These observations are in line with the observations reported by Langefors and
Kihlstorm (1978) and Nicholls and Hooker (1962). Furthermore, the numerical mod-
elling results validate the well acknowledged fact that a very small explosive energy is
used for fracturing and the majority of the explosive energy is lost in wave propaga-
tion, heat, sound and air-overpressure (Lownds and Du Plessis, 1984).
ABAQUS (2003) reports that the ratio of kinetic energy to internal energy (which
includes recoverable strain energy and energy consumed in plastic work) should not
be more than 10% to ensure accuracy of the dynamic numerical procedure. Also,
artificial strain energy, which is an indicator for hourglass effect, should be negligible

Fig. 16. Peak particle velocity plots (PPV) in the X-axis direction
Numerical Procedure for Dynamic Simulation 665

to assure accuracy of the dynamic numerical modelling. These conditions are also well
complied with as can be seen from Fig. 15.
Fracturing distance and PPV
Typical single point PPV observations are plotted in Fig. 16 for a point 0.5 m away
from the borehole in the case of ideal detonation and 0.25 m away in the case of non-
ideal detonation. PPV plots over a 10 m distance away from the borehole are also
plotted for the two types of detonation in Fig. 17. The plots shown in Fig. 16 are in
good resemblance with the routine field PPV measurements where PPV measurements
show reduction in magnitude and frequency with the increase of distance from the
source. This reduction is due to damping of the wave energy provided by the earth. It
is noteworthy here that no artificial damping is used in the present procedure yet the
explosive energy rapidly damps out, which is similar to field observations.
Prediction of rock fracturing based on PPV observations is a common practice in
mining. Several predictor equations are based on this context and most notably by
Holmberg and Persson (1979). They predict that fractures are most likely to take place
where PPV is more than 1000 mm=s. The PPV trends presented in Fig. 17 predict
similar trends as reported by Holmberg and Persson (1979). Also the extent of frac-
tured zones is limited to the distance with PPV values greater than 1000 mm=s.

Fig. 17. PPV trends for the two types of detonation characteristics
666 M. R. Saharan and H. S. Mitri

The results presented in Figs. 11–17 also validate the assumptions made during
the development of the new numerical procedure. The following illustrates these
validations.
(i) Modeled domain size
PPV plots presented in Fig. 16 indicate that the fracture network obtained is due to
the primary shock waves only. The PPV plots do not indicate the presence of second-
ary loading due to the reflected waves. Therefore, the choice of model domain size
was appropriate for such analyses.
(ii) Damping of the wave energy
PPV plots presented in Fig. 16 as well as PPV trends obtained by Fig. 17 illustrate
that the selected material model represents damping characteristics associated with the
natural material; no artificial damping was considered in the developed procedure.
Also, the material model shows enough potential to accurately simulate two distinct
fracturing characteristics, crushing and cracking, with a single parameter for element
elimination.
(iii) Duration of the analysis
Energy plots presented in Fig. 15 as well as stable fracture network illustrated in
Figs. 13 and 14 indicate that the results obtained are stable in nature. These results also
indicate that enough time is provided for fracture propagation. Therefore, the simulation
time adopted in the analysis is appropriate to obtain a stable fracture network.
(iv) Model boundary conditions
Roller boundaries were selected to represent unbounded rock medium. Stable
fractures networks, energy plots and PPV plots presented in Figs. 15–17 bring out
that the results are nowhere affected by the spurious reflection of waves from the
artificial boundaries. Hence, the selection of roller boundaries in the present modelling
procedure is justified.

5. Conclusions
This paper presents in detail a nonlinear numerical modelling procedure for the dy-
namic modelling of rock fracturing by blasting. This includes description and justifi-
cation of adopted model domain size and its boundary conditions, material model,
dynamic load as well as element type and size. A new method, named as optimized
pressure pulse, is introduced to approximate the observed pressure-time profile due to
the blast load. The results obtained from the dynamic numerical model are evaluated
and validated against established empirical knowledge base and observed dynamic
rock behaviour. The numerical results also confirm the validity of the assumptions
made during the development of the numerical procedure, particularly with regard to
model size, roller-type boundary conditions and damping. The developed procedure
thus presents a promising potential for many engineering blasting research problems.
One important limitation of the current modelling technique is that it can not be
used for investigating fracturing which ultimately results in complete fragmentation.
This is due to the fact that the model domain essentially remains continuum through-
out the solution time despite the use of the element elimination technique.
Numerical Procedure for Dynamic Simulation 667

Acknowledgements
The work presented in the paper is a part of research work done for PhD thesis of the first author.
The research work is supported by several organisations and institutes, most notably, Central
Mining Research Institute, Dhanbad, India (CMRI-India); Natural Sciences and Engineering
Research Council of Canada, Canada (NSERC-Canada) and J. W. McConnel Foundation, McGill
University, Canada. The authors are grateful for their generous support. Authors also express their
gratitude towards anonymous reviewers of the paper for improving the quality of the paper.

References
ABAQUS (2003): Abaqus 6.4-1 user’s manuals. Abaqus Inc., Pawtucket, RI, USA.
Aimone, C. T. (1992): Rock breakage: Explosives, blast design. In: Hartman, H. L. (ed.), SME
Mining Engineering Handbook. Society of Mining Engineers, Littleton, 722–746.
Anon (1987): Explosives and rock blasting. Atlas Powder Company. Washington, USA.
Atlas (1987): Explosives and rock blasting. Atlas Powder Company. Washington, USA.
Batzle, M. L., Simmons, G., Siegfried, R. W. (1980): Microcrack closure under stress: direct
observation. J. Geophys. Res. 85, 7072–7090.
Bhandari, S., Vutukuri, V. S. (1974): Rock fragmentation with longitudinal charges. In: Proc. 3rd
Int. Cong. Rock Mech. Denver, US, 1337–1342.
Biggs, J. M. (1964): Introduction to Structural Dynamics. New York, McGraw-Hill.
Bing, Z., Greenhalgh, S. (1998): A damping method for the computation of the 2.5 green’s
function for arbitrary acoustic media. Geophys. J. Int. 133(1), 111–120.
Blair, S. C., Cook, N. G. W. (1998): Analysis of compressive fracture in rock using statistical
techniques: Part I. A non-linear rule-based model. Int. J. Rock Mech. Min. Sci. 35, 837–848.
Braithwaite, W., Brown, W. B., Minchinton, A. (1996): The use of ideal detonation computer
codes in blast modelling. In: Mahanty, B. (ed.), Proc. 5th Int. Sym. Rock Fragmentation by
Blasting (FRAGBLAST 5), Montreal, Quebec, Canada. Rotterdam, Balkema, 37–44.
Brinkmann, J. R. (1987): Separating shock wave and gas expansion breakage mechanisms. In:
Proc. 2nd Int. Symp. Rock Fragmentation by Blasting.
Brinkmann, J. R. (1990): An experimental study of the effects of shock and gas penetration in
blasting. In: Proc. 3rd Int. Symp. Rock Fragmentation by Blasting, FRAGBLAST 3, Brisbane,
Australia, August 26–31, Published by AusIMM, 55–66.
Cho, S. H., Miyake, H., Kimura, T., Kaneko, K. (2003): Effect of the waveform of applied
pressure on rock fracture process in one free-face. J. Sci. Tech. Energetic Mat. 64(3),
116–125.
Clark, G. B. (1987): Principles of Rock Fragmentation. John Wiley & Sons Inc., London, p 610.
Daehnke, A., Rossmanith, H. P., Knasmillner, R. E. (1996): Blast-induced dynamic fracture
propagation. In: Mohanty, B. (ed.), Proc. 5th Int. Symp. Rock Fragmentation by Blasting,
FRAGBLAST 5, Montreal, Canada, 619–626.
Daniel, R. (2003): Pressure-time history of emulsion explosives. Personal communications.
Dawding, C., Labuz, J. F., Shah, S. P. (1985): Experimental analysis of crack propagation in
granite. Int. J. Rock Mech. Min. Sci. 22, 85–99.
Diederichs, M. S. (1999): Instability of hard rockmasses: The role of tensile damage and
relaxation. (Unpublished) PhD thesis, University of Waterloo, Canada, p 610.
668 M. R. Saharan and H. S. Mitri

Donze, F. V., Bouchez, J., Magnier, S. A. (1997): Modeling fractures in rock blasting. Int. J. Rock
Mech. Min. Sci. 34(8), 1153–1163.
Duvall, W. I. (1953): Strain wave shapes in rock near explosions. Geophysics 18(2), 310–323.
Etoh, S., Hamashima, H., Murata, K., Kato, Y. (2002): Determination of JWL parameters from
underwater explosion tests. In: 12th Int. Detonation Symp. (poster session). San Diego,
California.
Fourney, W. L., Dick, R. D., Wang, X. J., Wei, Y. (1993): Fragmentation mechanism in crater
blasting. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 30(4), 413–429.
Frantzos, D. C. (1989): Finite element analysis of radial cracking mechanism around blastholes
using measured pressure-time curves for low density ammonium nitrate=fuel oil. (Unpub-
lished) PhD thesis. Dep. Mining Eng., Queen’s Univ., Canada, p 290.
Gemant, A., Jackson, W. (1937): The Measurement of Internal Friction in Some Solid Dielec-
tric Materials. The London, Edinburgh and Dublin Philosophical Magazine J Science XXII,
960–983.
He, H., Liu, Z., Nakamura, K., Abe, T., Wakabayashi, K., Okada, K., Nakayama, Y., Yoshida, M.,
Fujiwara, S. (2002): Determination of JWL equation of state parameters by hydrodynamically
analytical method and cylinder expansion test. Jl. Japan Explosives Soc. 63(4), 197–203.
Hilleborg, A., Modeer, M., Petersson, P. E. (1976): Analysis of Crack Formation and Crack
Growth in Concrete by Means of Fracture Mechanics and Finite Elements. Cement Concrete
Res. 6, 773–782.
Holmberg, R., Persson, P. A. (1979): Design of tunnel perimeter blasthole pattern to prevent rock
damage. Tunneling ‘79. Inst. Min. Metall., London, 280–283.
Hughes, T. J. R. (2000): The finite element method: linear static and dynamic finite element
analysis. Dover publications, NY, p 682.
Hustrulid, W. (1999): Blasting principles for open pit mining, Vol. I – general design concepts.
Balkema, Rotterdam, p 382.
Itasca (2001): FLAC3D 2.0 manuals. Itasca consulting Group Inc. Minneapolis, MN, USA.
Jung, W. J., Utagava, M., Ogata, Y., Seto, M., Katsuyama, K., Miyake, A., Ogava, T. (2001):
Effects of rock pressure on crack generation during tunnel blasting. Jl. Japan Explosives Soc.
62(3), 138–146.
Kexin, D. (1995): Maintenance of roadways in soft rock by roadway-rib destress blasting. China
Coal Society 20(3), 311–316 (In Chinese).
Kikuchi, N. (1983): Remarks on 4CST elements for incompressible materials. Comput. Methods
Appl. M. 37, 109–123.
Kranz, R. I. (1983): Microcracks in rocks: a review. Tectonophysics 100, 449–480.
Kutter, H. K. (1967): The interaction between stress wave and gas pressure in the fracture
process of an underground explosion in rock, with particular application to presplitting.
(Unpublished) PhD Thesis. Univ. of Minnesotta, Minneapolis, p 234.
Kutter, H. K., Fairhurst, C. (1971): On the fracture process in blasting. Int. J. Rock Mech. Min.
Sci. 8, 181–202.
Langefors, U., Kihlstorm, B. (1978): The modern techniques of rock blasting. 3rd edn., John
Wiley & Sons Inc., New York, p 438.
Lima, A. D. R., Romanel, C., Roehl, D. M., Araujo, T. D. (2002): An adaptive strategy for the
dynamic analysis of rock fracturing by blasting. In: Proc. Int. Conf. Computational Eng. &
Sci. (ICES’02). Reno, Nevada.
Numerical Procedure for Dynamic Simulation 669

Liu, L. (1997): Continuum modelling of rock fragmentation by blasting. (Unpublished) PhD


thesis. Queen’s University, Canada, p 240.
Lownds, C. M., Du Plessis, M. P. (1984): The behaviour of explosives in intermediate-diameter
boreholes. In: Proc. Inst. Quarrying Trans. December, 709–804.
Lysmer, J., Kuhlemeyer, R. L. (1969): Finite dynamic model for infinite media. J. Eng. Mech. Div.
ASCE 92, SM1, 65–91.
Massarsch, K. R. (1992): Static and dynamic soil displacements caused by pile driving. In:
Barends, F. B. J. (ed.), Application of stress-wave theory to piles. Balkema, Rotterdam, 15–24.
McHugh, S. (1983): Crack extension caused by internal gas pressure compared with extension
caused by tensile stress. Int. J. Fracture 21, 139–145.
McHugh, S., Keough, D. (1982): Use of laboratory derived data to predict fracture and
permeability enhancement in explosive pulse tailored field tests. In: Proc. US Symp. Rock
Mech., Issues in Rock Mech. 504–514.
Metzger, D. R. (2003): Adaptive damping for dynamic relaxation problems with non-monotonic
spectral response. Int. J. Numer. Meth. Engng. 56, 57–80.
Mishnaevsky, L. L. Jr., Schmauder, S. (2001): Continuum mesomechanical finite element modell-
ing in materials development: a state-of-the-art-review. Appl. Mech. Rev. 54(1), 49–69.
Mohanty, B. (2003): Personnel communications.
Mosinets, V. N., Garbacheva, N. P. (1972): A seismological method of determining the parameters
of the zones of deformation of rock by blasting. Sov. Min. Sci. 8(6), 640–647.
Nicholls, H. R., Hooker, V. E. (1962): Comparative studies of explosives in salt. U.S. Bur. Mines
R.I. 6041, p 46.
Nie, S., Olsson, M. (2000): Study of fracture mechanism by measuring pressure history in
blastholes and crack lengths in rock. In: Proc. 27th Annual Conf. Explosives and Blasting
Technique, Orlando, USA, 291–300.
Obert, L., Duvall, W. (1967): Rockbursts, Bumps, and Gas Outbursts, chapter 19 in ‘Rock
Mechanics and the Design of Structures in Rock’. John Wiley & Sons Inc., 582–593.
Olatidoye, O., Sarathy, S., Jones, G., McIntyre, C., Milligan, L. (1998): A representative survey of
blast loading models and damage assessment methods for buildings subject to explosive
blasts. Report by Nichols Research Corporation. Report No. CEWES MSRC=PET TR=98-36,
p 14.
Olsson, M., Nie, S., Bergqvist, I., Ouchterlony, F. (2001): What causes cracks in rock blasting? In:
Proc. EXPLO2001. Hunter valley, NSW, Australia, 191–196.
Persson, P. A., Holmberg, R., Lee, J. (1994): Rock blasting and explosives engineering. Boca
Raton, CRC Press, Florida.
Prasad, U., Mohanty, B., Nemes, J. A. (2000): Dynamic Fragmentation of Selected Rocks Under
Impact Loading. In: Proc. 4th North American Rock Mech. Symp., Seattle, Washington,
577–581.
Ramshaw, C. L., Selby, A. R., Bettess, P. (1998): Computation of the transmission of waves from
pile driving. In: Skipp, B. O. (ed.), Ground dynamics and man made processes. T. Telford
Pub., London, 115–128.
Repetto, E. A., Radovitzky, R., Ortiz, M. (2000): Finite element simulation of dynamic fracture
and fragmentation of glass rods. Comput. Methods Appl. Mech. Engrg. 183, 3–14.
Robertson, N. J., Hayhurst, C. J., Fairlie, G. E. (1994): Numerical simulation of explosion
phenomena. Int. J. Computer Appl. Tech. 7(3–6), 316–329.
670 M. R. Saharan and H. S. Mitri: Numerical Procedure for Dynamic Simulation

Saharan, M. R. (2004): Dynamic modelling of rock fracturing by blasting. (Unpublished) PhD


thesis. Mining, Metals and Materials Dep., McGill Univ., Canada, p 274.
Saharan, M. R., Mitri, H., Jethwa, J. L. (2006): Rock fracturing by explosive energy: Review of
state-of-the-art. Fragblast, Vol. 10, No. 1–2, March–June 2006, pp. 61–81.
Sochacki, J. R., Kubicheck, R., George, J., Fletcher, W. R., Smitheson, S. (1987): Absorbing
boundary conditions and surface wave. Geophysics 52, 60–71.
Sun, C., Later, D. W., Chen, G. (2001): Analysis of the effect of borehole size on explosive energy
loss in rock blasting. Int. J. Rock Fragmentation and Blasting – FRAGBLAST 5(4), 235–246.
Takewaki, I. (2000): Soil-structure random response reduction via TMD-VD simultaneous use.
Comput. Methods Appl. Mech. Engrg. 190, 677–690.
Valliapan, S., Lee, I. K., Murti, V., Ang, K. K., Ross, A. H. (1983): Numerical modelling of rock
fragmentation. In: Stephansson, O. (ed.), Proc. 1st Int. Symp. Rock Frag. By Blasting –
FRAGBLAST 1. Balkema, Rotterdam, 375–390.

Author’s address: Hani S. Mitri, Department of Mining, Metals and Materials Engineering,
McGill University, 3450 University Street, Montreal, Canada H3A 2A7; e-mail: hani.mitri@
mcgill.ca

S-ar putea să vă placă și