You are on page 1of 183

FRAGM: A BLASTING FRAGMENTATION MODEL OF ROCKS

by
Sergey Victorovych Zagreba






A Thesis submitted to the College of Engineering and
Mineral Resources at West Virginia University in partial
fulfillment of the requirements for the degree of

Master of Science in Mining Engineering





Approved by


Syd S. Peng, Ph. D., Committee Chairperson
Felicia Peng, Ph. D.
Yi Luo, Ph. D











Department of Mining Engineering
Morgantown, West Virginia
2003


Keywords: rock blasting, fragmentation prediction, burden influence
ABSTRACT
FRAGM: A BLASTING FRAGMENTATION MODEL OF ROCKS
by
Sergey Victorovych Zagreba


Fragmentation is a major concern of any blasting operation. Information on the
degree and size distribution of fragments within a blasted rock mass is essential for
efficient rock loading and crushing operations.
Detailed literature review shows that majority of the previous researchers estimated
blast fragmentation by considering four basic variables, i.e. rock properties, explosive
properties, drilling pattern and bench geometry. However, when considering the effect of
burden on rock fragmentation, simplified assumptions were made and its variability along
the bench height was often ignored.
In reality, because of the non-uniform burden along the bench height, the actual
powder factor in the front row of holes could differ significantly from the one estimated
assuming uniform burden. This will be more so if the bench is highly irregular and has
significantly different toe and crest burdens. Ignoring this fact may result in a poor fit of
the existing fragmentation models for the actual data. Research in this thesis
demonstrates this fact and the results underscore the importance of considering the true
bench profile for the blast fragmentation analysis.
For the present work, the irregularity of the bench profile and the resulting non-
uniform burden was estimated using the laser profiling technique. A new rock
fragmentation model, FRAGM, which considers the variability in burden, was developed
and verified by comparing with actual blast results from a West Virginian limestone
quarry. The developed model can be used as a quick and reliable means to predict or
assess the rock fragmentation before or after a blast.




iii
Acknowledgment


I wish to express my sincere gratitude to Dr. Syd S. Peng, Thesis Advisor and
Chairman of Department of Mining Engineering of the West Virginia University, for his
patience and guidance throughout this study. Without his support this work would not
have been possible. I would like to convey my thanks and appreciation to the other
committee members, Dr. Felicia Peng and Dr. Yi Luo for their constructive criticisms and
concrete suggestions at several stages of this work.
A special thanks to Ms. Karen Centofanti for her assistance in the official work.
Finally, I am greatly indebted to my wife, Mary Beth, and my entire family whose
support can never adequately be expressed in words. Thank you for encouraging me
throughout my life to meet challenges with determination and to strive for success.




iv
Table of Contents

Abstract .. ii

Acknowledgment iii

Table of Contents ... iv

List of Tables . vi

List of Figures vii

Chapter 1 Introduction .. 1

Chapter 2 Literature Review 4

2.1 Mechanism of rock breakage by blasting . 4

2.1.1 Radial cracking theory . 6
2.1.2 Shock wave theory 11
2.1.3 The contribution of the shock wave and gas pressure .13

2.2 Extent of blast damage zones .. 19
2.3 Effect of controllable blast parameters on fragmentation .. 22
2.4 Effect of discontinuities on rock fragmentation by blasting 30
2.5 Methods of study of rock fragmentation by blasting .. 31
2.5.1 Kuz-Ram model 34
2.5.2 Hole-by-hole analysis 40

Chapter 3 Development of the Proposed Model FRAGM .. 41

Chapter 4 Description of the Rock Fragmentation Prediction
Engineering Model FRAGM . 55

4.1 Bench and borehole information . 57
4.1.1 Burden 57
4.1.2 Spacing ... 59
4.1.3 Powder factor . 60
4.1.4 Loading density . 60

v
4.1.5 Diameter of blastholes ... 61
4.1.6 Stemming ... 62
4.1.7 Subdrilling .. 63

4.2 Explosive types .. 64
4.3 Rock strength .. 64
4.4 Delay time .. 68
4.5 Geological conditions .. 69
4.6 Output data description .. 70

Chapter 5 Verification of the Proposed Engineering Model 71
5.1 Introduction . 71
5.2 Image analysis technique and sampling .. 72
5.3 Constraints .. 74
5.4 Verification of proposed development in the field . 75

Chapter 6 Quality Control .. 104
6.1 Introduction . 104
6.2 Ammonium nitrate based dry explosives . 104
6.3 Slurries 105

Chapter 7 Conclusions and Recommendations . 106
References 109
Appendix A Verification of the Proposed Methodology
by Hand Calculations 120

Appendix B Profiling the Quarry Bench Face . 134

Appendix C Using an Image-Processing Program .. 162
Split-Desktop

vi
List of Tables




Table 2.1 Spalling-related values by Petkof et al. (1961) 18
Table 4.1 Explosive strength based on composition 65
Table 4.2 Rock properties by Mohanty (1987) 66
Table 4.3 Rock properties presented by Cook (1976) . 67
Table 5.1 Predicted and actual size distribution for the blast # 1 79
Table 5.2 Predicted and actual size distribution for the blast # 2 85
Table 5.3 Predicted and actual size distribution for the blast # 3 91
Table 5.4 Predicted and actual size distribution for the blast # 4 96
Table 5.5 Predicted and actual size distribution for the blast # 5 .. 102
Table B.1.1 Results of the bench face profiling (Blast # 1) .. 137
Table B.1.1.1 Results of the estimation of rock volume in the first
row of blastholes (Blast # 1) .. 140

Table B.1.2 Results of the bench face profiling (Blast # 2) ... 141
Table B.1.2.1 Results of the estimation of rock volume in the first
row of blastholes (Blast # 2) .. 145

Table B.1.3 Results of the bench face profiling (Blast # 3) . 146
Table B.1.3.1 Results of the estimation of rock volume in the first
row of blastholes (Blast # 3) .. 149

Table B.1.4 Results of the bench face profiling (Blast # 4) ... 150
Table B.1.4.1 Results of the estimation of rock volume in the first
row of blastholes (Blast # 4) . 155

Table B.1.5 Results of the bench face profiling (Blast # 5) ... 156
Table B.1.5.1 Results of the estimation of rock volume in the first
row of blastholes (Blast # 5) .. 161





vii
List of Figures




Figure 2.1 Diagrammatic representation of the interaction of a spherical
wave with the radial crack system modified by Roberts and
Wells (1954) . 7

Figure 2.2 Favorable reflection geometry for extending cracks towards the
free face by Roberts and Wells (1954) 7

Figure 2.3 a)incidence of the reflected wave at the crack tip
b) region influenced by the reflected waves
c) theoretical crater development a single hole; presented by
Roberts and Wells (1954) 10

Figure 2.4 Typical distance-time plot for granite by Petkof et al. (1961) . 15
Figure 2.5 Typical distance-time plot for marble by Petkof et al. (1961) . 16
Figure 2.6 Typical distance-time plot for limestone by Petkof et al. (1961) . 17
Figure 2.7 The damage zones surrounding an explosive charge 20
Figure 2.8 Crack system for the square pattern with 80 mm holes and
0 mm standard deviation in drilling by Lownds (1976) . 24

Figure 2.9a Crack system for the square pattern with 80 mm holes and
200 mm standard deviation in drilling by Lownds (1976) 25

Figure 2.9b Crack system for the square pattern with 80 mm holes and
300 mm standard deviation in drilling by Lownds (1976) 26

Figure 2.10a, b Effective circles around holes for a square pattern: a) no
deviation in drilling; b) 30 % of the burden deviation in
drilling; by Lownds (1976) 28

Figure 2.11 a, b Increased radius of effective circles: a) no deviation in
drilling; b) 30 % of the burden deviation in drilling;
by Lownds (1976) 29

Figure 3.1 Diagram of a blasting pattern and the geometry of
charged blastholes 42

Figure 3.2 Typical representation of the drillhole inclination versus
the bench face angle presented by Oloffson (1990) . 43


viii
Figure 3.3 Variation of burden in case of vertical holes by Bhandari (1997) .. 45
Figure 3.4 Parts of a bench presented by Hustrulid and Kuchta (1998) 46
Figure 3.5 Two areas of breakage .. 48
Figure 3.6 Three cases in main breakage area ... 51
Figure 4.1 Procedures of blast round design and evaluation of blasting
performance . 56

Figure 4.2 X-Y-Z coordinates data of the bench boreholes 58
Figure 5.1.1 The quarry highwall before actual shot (blast # 1) . 76
Figure 5.1.2 Muckpile immediately after the shot (blast # 1) . 77
Figure 5.1.3 Cumulative size distribution (blast # 1) .. 80
Figure 5.2.1 The quarry highwall before actual blast (blast # 2) . 82
Figure 5.2.2 Muckpile a few days after the shot (blast # 2) . 83
Figure 5.2.3 Cumulative size distribution (blast # 2) .. 86
Figure 5.3.1 Quarry highwall and muckpile approximately week after
the shot (blast # 3) ... 88

Figure 5.3.2 Muckpile after the shot (blast # 3) ... 89
Figure 5.3.3 Cumulative size distribution (blast # 3) ... 92
Figure 5.4.1 Muckpile immediately after the shot (blast # 4) .. 94
Figure 5.4.2 Cumulative size distribution (blast # 4) .. 97
Figure 5.5.1 The quarry highwall before actual shot (blast # 5) . 99
Figure 5.5.2 Muckpile immediately after the shot (blast # 5) .. 100
Figure 5.5.3 Cumulative size distribution (blast # 5) 103
Figure A-1 Typical bench face profile of borehole in the first row .. 123

Figure B-1 Laser surveying equipment used for profiling the face
(courtesy Measurement Devices Ltd.) 135

Figure C-1 Typical JPEG image is taken in the field for the calculation
of size distribution .. 164

Figure C-2 Typical grayscale image .. 165
Figure C-3 Typical binary image ... 166



1
Chapter 1 Introduction


Adequate rock fragmentation is the major objective of any blasting operations.
Fragmentation is the basic concern in rock blasting and serves as the main measure of
blasting effectiveness. Good fragmentation is the key to successful mining operation and
equipment maintenance. It is desirable to have a uniform fragment size distribution,
avoiding both fines and oversizes. It is very important that blast pattern can be quickly
and accurately analyzed before actual blast. Any mining operators can minimize total
production costs per ton of rock blasted. This requires an evaluation of the component
costs, which include drilling, blasting, loading, hauling and crushing costs.
The drilling and blasting are the first unit operations in the mining process and have
a major impact on the performance and cost of subsequent unit operations. An increase in
the degree of fragmentation will give the loading equipment a higher rate of productivity.
This will result in lower costs per ton or cubic yard moved. The effect of wear and tear
will also decrease, giving lower operating cost per hour.
Under similar conditions of haul, lift, size and type of truck, and haul road
condition, truck production per hour will increase with greater degree of fragmentation
due to faster shovel or loader loading rates and a decrease in bridging at the crusher.
There will be a consequent decrease in cycle time. At a standard operating cost per hour,
this increase in truck speed or productivity will result in lower unit operating costs.
An increase in the degree of fragmentation gives lower crushing costs as more
material passes through as undersize. Liner costs, repair and maintenance, and bridging
time will decrease and the crushing rate per hour will increase. The decreased bridging
time also cuts down on truck delay time at the crusher, which, in turn, gives higher truck

2
and shovel (loader) productivity. Any increase in the degree of fragmentation means less
work for the crusher.
These have been the easiest to explain since the unit costs always decrease with
increasing fragmentation. The same is not true for the drilling and blasting costs. There
are many possible combinations, which can occur depending upon the particular design.
For a given rock type, geologic structure, and firing sequence, an increase in the degree
of fragmentation may be achieved by (a) increasing the consumed quantity of a given
explosive, (b) changing to an explosive having greater energy content per unit hole
volume (higher energy content/ density), or (c) combinations of both.
For blasting case (a) the associated drilling cost would increase if the explosive
quantity were to be increased by simply drilling the same diameter drill holes but on a
tighter pattern. Thus there would be more drill holes required to blast a given volume. If
larger diameter drill were substituted and the increased hole volume achieved in this way
then the rate of increase or decrease would depend upon the comparative drilling cost per
foot of hole.
For case (b), assuming that the same hole diameter and pattern are used, the drilling
cost would remain constant independent of the fragmentation.
For case (c) the drilling cost could remain constant, increase or decrease depending
upon the situation.
The objective of this research work are: 1) to establish a methodology for blast
fragmentation prediction; 2) to develop an engineering model for fragmentation size
prediction based on the methodology established in the first step, and to provide an

3
algorithm for prediction; and 3) to compare field data with predicted values to verify the
model.
The engineering model considers four main factors: explosive properties, rock
properties, drilling pattern and actual bench geometry. Field data would be used to verify
the theoretical development.



















4
Chapter 2. Literature Review

2.1 Mechanism of rock breakage by blasting


Blasting theory is one of the most interesting, challenging, and controversial areas of
the explosives engineering. It encompasses many areas in the science of chemistry,
physics, thermodynamics, shock wave interactions, and rock mechanics. In broad terms,
rock breakage by explosives involves the action of an explosive and the response on the
surrounding rock mass within the realms of energy, time and mass. This chapter content
will emphasize the concept associated with blasting theories, rather than a rigorous
mathematical, physical, or chemical treatment through formulae. Where formulae are
used, they are merely to enhance the concept presented.
In spite of the tremendous amount of research conducted in the last few decades, no
single blasting theory has been developed and accepted that adequately explains the
mechanisms of rock breakage in all blasting conditions and material types. There is as yet
no consistent and widely applicable theory of blasting, but only a number of limited
theories, many of which are empirical in nature and based on ideal situations. The
reflected theory has been chosen in this thesis for its simplicity and ease of applications.
Rock fracture resulting from explosion process of explosives load in drill holes
depend on the number of free faces, the burden, the hole placement and rock geometry,
the physical properties and loading density of the explosive, the type of stemming, the
rock structure and mechanical strength, and other factors. Final fragmentation in a bench
blasting operation can be attributed to a combination of:
1. crushing of the rock immediately around the explosive cavity;

5
2. initial radial fracturing due to tensile tangential stress component in the outgoing
stress wave;
3. secondary radial fractures formed at the surface, propagating inward, due to
enhanced tangential stress accompanying free surface displacement;
4. extension of the initial radial fractures by reflected radial tensile strain at oblique
angles to the surface;
5. joining of inward propagating radial fractures with initially created outward radial
fractures;
6. tangential fractures formed at the surface, propagating parallel to the free surface;
7. tensile separation and shear of rock at places of weakness in the rock mass;
8. separation of the rock due to reflected radial tensile strain;
9. fracture and acceleration of fragments by strain energy release;
10. further fracture and acceleration of broken rock by late expanding gases; and
11. pre- existing discontinuities in the rock mass.

While none of these mechanisms can be ignored, explosive- generated radial
fractures are crucial in determining the overall fragmentation as Harries and Hengst
(1977) and Lownds (1983) showed using simulation models. Those studies describe
computer simulation models, which calculate blast results providing information on the
fragmentation. The models are based on Harries hypothesis (1977) that radial fractures
are primarily responsible for fragmentation in rock.




6
2.1.1 Radial cracking theory

The practical implications for blasting are described by Roberts and Wells (1954).
They explained that the cracks grew outward away from the hole under the action of the
gas pressure. The authors pointed out that the cracks would have extended a distance of
only 0.38 times the burden width (0.38B) when that portion of the incident wave directly
in front of the charge meets the free surface (Fig 2.1a). The crack tip and the front of the
reflected wave will meet at the distance (X).


c
X
c
X B
38 . 0
2
=

; (2.1.1.1)
X = 0.55B (2.1.1.2)
where B- the burden, ft
c- the crack velocity, ft/ sec.

As indicated, the crack growing in this direction should stop or be retarded at this
point. Cracks growing at an angle to the surface will also be influenced by the
reflected wave. If this wave meets the tip of the static crack it can accelerate the increase
of the crack growth (Roberts and Wells, 1954). The length e will depend upon the crack
orientation and crack velocity.

As can be seen (Fig. 2.2)
tan 2 =
h
e
(2.1.1.3)

7







Figure 2.1 Diagrammatic representation of the interaction of a spherical wave
with the radial crack system modified from Roberts and Wells (1954).








Figure 2.2 Favorable reflection geometry for extending cracks towards the free
surface by Roberts and Wells (1954).





8
and sin 2 =
g
e
(2.1.1.4)
hence
h =
2 tan
e
(2.1.1.5)
and
g =
2 sin
e
(2.1.1.6)

The time required for the wave to travel a distance h = g at velocity c must be the
same as that needed by the crack traveling at a velocity V
crack
to travel distance e. Thus

t
crack
=
crack
V
e
=
c
h g +
= t
wave
(2.1.1.7)


Assuming that
V
crack
= K c (2.1.1.8)
where K is a constant, then

e
h g +
=
K
1
(2.1.1.9)

Substituting the values for h and g into equation (2.1.1.9) one finds that


9

K
1
2 sin
1
2 tan
1
=
|
|
.
|

\
|
+

(2.1.1.10)

Simplifying yields
tan = K (2.1.1.11)
or
= tan
1
K (2.1.1.12)

If the crack velocity is known then the orientation of cracks which will be effected
can be calculated. For the case (Robert and Wells, 1954) when

V
crack
= 0.38c (2.1.1.13)

The most favorable crack orientation would be
= tan
1
0.38
= 20.8
0
(2.1.1.14)
The length e to which the crack would have grown to its encounter with the
reflected wave is
e = 2B sin (2.1.1.15)
and the distance, which the wave would have traveled is
g + h = | |

2 cos 1
cos
+
B
(2.1.1.16)



10












Figure 2.3a) incidence of the reflected wave at the crack tip; b) region influenced by the
reflected waves; and c) theoretical crater development of a single hole; presented by
Roberts and Wells (1954).





(a)
(b)
(c)

11
As can be seen in the expanded view of the crack tip (Figure 2.3a), it is the radial
(tensile) component of the reflected wave, which acts to extend the crack tip. The
tangential component is compressive and acts in the direction of crack propagation. Other
cracks in the burden region will also be affected by the reflected radial component and
encouraged to extend. Crack oriented behind the line of blastholes (Figure 2.3b) will not
be affected.
For a single hole shot in close proximity to a free face (Figure 2.3c), the expected
included angle (B) for the broken rock is expected to be

B = 180
0
2 = 138.4
0
(2.1.1.17)

Assuming that V
crack
= 0.38 c, it agrees quite well with what has been observed in
practice (Roberts and Wells, 1954).


2.1.2 Shock wave theory

Shock wave theory of rock breakage has been proposed in many forms by different
researchers. This model states that most of the rock breakage in a blast occurs at a free
face as a result of spalling, which occurs when a compressive wave is reflected at a free
boundary. The slabs, which are spalled from the rock edge, are formed in a succession of
increasing thickness, where the number of slabs depends on the amplitude and duration of
the stress wave.

12

The United States Bureau of Mines conducted a series of experiments to study the
explosion crater to confirm the shock wave theory. Atchison and Duvall (1957) studied
the relationship between the radial strain and explosive energy for a concentrated charge.
Their results may be expressed by the equation below

=
n
m
W
D
C
P

3 / 1 2

(2.1.2.1)

where D/W
1/3
is the scaled distance; P
m
is the explosion pressure; is the explosive
density; c is the wave velocity; and is the strain wave. They found n was 1.56 for the
transition zone and 1.24 for the seismic zone in Greenstone granite using 60% ammonia
gelatin at a density 1.41 g/cc. This relationship should apply quite generally since it does
not depend on any specific type of wave form. Therefore, the factor n will vary from one
set of conditions to another.

For cylindrical charges, the empirical formula to determine the relationship between
the radial strain and the explosive properties was studied by the Soviet researchers.
Adushkin et al. (1987) published the following equation:


C E
U
rm
/
= 6.62 * 10
-3

2 . 1

qe
R
(2.1.2.2)


13
where:
U
rm
- the radial component of the maximum particle velocity, m/sec;
R - the distance between observation points, m;
qe - the linear density of the explosive for a Trotyl charge equivalent in energy,
kg/m,
- the material density, kg/m
3
;
c - the elastic lontitudinal wave velocity in the material, m/sec;
E - the concentration of explosive energy in the charge, J/m
3
.

One argument against the shock wave theory proposed by Langefors and Kihlstrom
(1963) is the fact that fragmentation could occur in the absence of a high intensity stress
pulse, which cannot be explained by the shock wave theory. However, there is no doubt
that strain magnitude in blast waves is a very important factor in rock breaking
mechanism and must therefore be included in any useful model.

2.1.3 The contribution of the shock wave and gas pressure

Some of the earliest photographic studies regarding bench blasting were reported by
the United States Bureau of Mines in the late 1950s (Atchison and Duvall, 1956; Blair,
1960). At that time the emphasis was on studying the shock wave aspect of blasting
action. With this concept, the shock wave generated by the blast travels to the free
surface from which it reflects. If the tensile strength is low, compared to the amplitude of
the tensile portion of the wave, the rock face will spall. The spalled rock then travel away

14
from the remaining rock with a certain velocity. This is obviously a contribution to the
overall heave of burden rock.

The United States Bureau of Mines conducted a comprehensive blasting research
program in the 1960s, which has resulted in findings of great interest. The test performed
and reported by Petkof et al., (1961) was to study the motion of bench faces during
blasting. The techniques used in these studies were (1) high-speed (1,000 frames per
second) photography of the quarry face in front of one charge hole of a quarry blasting
round and (2) measurement of the radial strain wave generated in the rock by the charge
at a distance equal to the burden.
The high-speed camera provided data in the form of a motion picture of the blast.
Frame-by-frame projection of the film permits the measurement of the rock movement as
a function of time. The time of charge detonation could be obtained by observing the
flash of light from the detonation cord. In cases where the detonating cord could not be
used as a timer, the origin of the time axis was the first observable motion of the rock.
The slope of the line drawn through the points of the distance- time plot is the horizontal
velocity of the broken rock or the fly-rock velocity. Typical results for the three rock
types are given in Figure 2.4, 2.5, and 2.6. The point indicated as calculated on the
figures is that corresponds to the shock wave at the face.
In general, the overall curve consists of two straight line segments. The second of
these segments has a significantly greater slope than the first indicating a stepwise
increase in the velocity at the inflection point.



15




Figure 2.4 Typical distance-time plot for granite by Petkof et al., (1961).





16




Figure 2.5 Typical distance-time plot for marble by Petkof et al., (1961).



17




Figure 2.6 Typical distance-time plot for limestone by Petkof et al., ( 1961).




18



Table 2.1. Spalling-related values (Petkof et al., 1961)

Rock type

Shot

Measured Strain,
m

Spalling Velocity,
ft/sec

Time,
msec

L - 1

1150

43

0.065

L - 2

1000

37

0.44

L - 3

750

28

0.46


Granite

L - 4

1350

50

0.52

T - 1

270

11

0.67

T 2A

700

29

0.78



Marble

T 2B

510

22

1.02

Limestone

L - 1

650

18

0.92




19
The measured initial velocity of the block coming off the rock face has the same
order of magnitude as expected according to the reflection theory of rock breakage
(Petkof et al., 1961). The increasing curve slopes suggest that after some period of time
the rock at the rear of the blasted burden is moving faster than the spalled blocks. This is
consistent with the acceleration of the main portion of the burden by the gas pressure.
Because of its higher velocity but delayed starting time, the bulk of the blast catches up to
the spalled blocks and accelerates them. In the impact, which occurs between these two
parts of the burden some additional breaking is possible.
The conclusion from this series of tests is that under normal bench blasting
conditions spalling is a minor contributor to the heaving process. This process is
dominated by the action of the gas pressure.

2.2 Extent of blast damage zone

The prediction and observation of the nature and extent of the damage produced in
the surrounding rock when an explosive charge detonates in a borehole is of major
practical significance for engineered rock excavation.
The radius of the damage zone formed when a cylindrical charge detonates in a rock
mass is one of the most important parameters required in the development of a
scientifically based method for designing blast patterns. The process taking place in the
rock surrounding a charge are so complex that an exact mathematical description is
presently impossible (Hustrulid, 1999). The damage mechanisms change as the distance
from the explosion increases, and many investigators therefore distinguish a number of

20
















Figure 2.7 The damage zones surrounding an explosive charge.






21
zones within which the stressed state and the fracture pattern differ. The paper
Calculation of fracture zones created by exploding cylindrical charges in ledge rocks
by Drukovanyi et al., (1976) is in the authors opinion of special importance in this
regard.
In this paper the investigators distinguish a number of zones within the stressed state
and their fracture patterns differ:
In zone 1 immediately adjoining the charge (Figure 2.7) the compressive stresses
exceed the compressive strength of the rock. In this zone considerable deformations
occurs and fracture is of overcrushing type, i.e., fine crushing is observed. Current
assessments of the size of this zone are conflicting, but most foreign investigators hold
the view that the size of the zone of plastic flow does not exceed 3-5 charge radii.
Zone 2, the zone of radial fissures, has a rather complex structure. The intensity of
crushing of the rock in this zone as a result of the intersection of (a) the natural radial
fissures by the radial ones, and (b) possible circular ruptures forming around the charge
cavity, decreases with increasing distance from the explosion center. The outer boundary
of this zone is located outside the region of individual radial fissures.
Zone 3 is the zone of elastic deformation. For practical purposes, it is of maximum
interest to establish the boundaries of the region with a specific crushing intensity
(fragment characteristics) or the zone of so-called controllable crushing. The theoretical
solution of this problem is extremely complex because one of the governing factors of the
process of crushing is the jointing pattern of the rock mass. This means that the use of
the theory of mechanics, which is applicable to a continuous medium, cannot provide an
exact solution.

22
Experimental assessment of fracture zone extent obtained by different methods vary
quite widely. However, to a first approximation one can assume that when high
explosives are fired in direct contact with the hole wall the boundary of the zone of radial
fissures is located at a distance of 40-50 charge radii from the explosion center
(Azarkovich, 1965, 1981; Anonymous, 1971; Vovk et al., 1973).


2.3 Effect of Controllable Blast Parameters on Fragmentation

This research study is focused on the prediction of the rock fragmentation
distribution resulting from a given bench blast operation. This is not an easy task, since
theoretical developments of rock breakage are hindered by the numerous variables
influencing the phenomenon. More than twenty factors appear to affect fragmentation in
a blast (Da Gama, 1983). The effect of interaction between several blast design variables
on fragmentation results had been studied by many researchers. These variables are
powder factor, drilling pattern, borehole diameter, delay timing, and drilling inaccuracy.
Fragmentation results described by the mean fragment size alone are inadequate and
a full description of the entire size range is needed. It was realized that the distribution
curve (the Rosin-Rammler Curve) had been generally recognized as giving a reasonable
description of fragmentation in blasted rock (Faddeenkov, 1975; Cunningham, 1983,
1987; Harries and Hengst, 1977). To define the Rosin-Rammler curve two parameters are
needed, namely, the characteristic fragment size (X
c
) and the exponent n which
characterizes the root-mean-square deviation from the mean or in other words the

23
uniformity of crushing. If n>1 then the characteristic fragment size (X
c
) is approximately
equal to the mean fragment size (50% passing).
The behavior of n with inaccuracy in drilling had been studied by Lownds (1976,
1983). He stated that increasing inaccuracy in hole position results in a significant
decrease of degree of uniformity of the blasted material. From Figures 2.8 and 2.9 it can
be seen that greater borehole diameters or, in other words, higher specific consumption of
explosives gives better and more uniform fragmentation (lower X
c
and higher n).
On the other hand, as can be seen from Figure 2.9 that increasing inaccuracy in hole
position had little effect on the characteristic fragment size, except for the case of square
pattern with 50 mm holes. It seems that when the amount of explosive per hole is such
that the radius of affected rock mass from the explosion is small, drilling misalignment
not only gives non- uniform fragments but also bad fragmentation. This statement will be
demonstrated by the following simple example.
When holes are fired independently, there will effectively be a cylinder of broken
rock mass around each hole after firing (Lownds, 1976). In a horizontal section through
the bench, each cylinder can be represented as a circle. For fracture of the whole rock
mass during blasting every point in the section must be within at least one of these
circles. For example, in Figure 2.10a, a square drilling pattern with effective circles
around the holes is shown. A 30% of the burden deviation in drilling resulted in the
pattern shown in Figure 2.10b. It can be seen from this figure that a large proportion of
the area intended to be fractured remained unaffected from radial cracks emanating from
the holes. This resulted in a larger characteristic fragment size compared to the case with



24






















Figure 2.8 Crack system for the square pattern with 80mm holes and 0mm standard
deviation in drilling by Lownds (1976).








25



















Figure 2.9a Crack system for the square pattern with 80mm holes and 200mm standard
deviation in drilling by Lownds (1976).






26



















Figure 2.9b Crack system for the square pattern with 80mm holes and 300mm standard
deviation in drilling by Lownds (1976).





27
no drilling inaccuracy, and also in a non- uniform fragmentation because of the obvious
asymmetry of the track system.
A 33% increase in the radius of the effective circles around blastholes (by increasing
the amount of explosives per hole) and with no faulty drilling is shown in Figure 2.11a,
while a 30% of the burden deviation in drilling is shown in Figure 2.11b.
From Figure 2.11b, it can be observed that with this radius of effective circles there
are some small areas unaffected by the explosives action. That is, the mean fragment is
expected to remain the same, but the uniformity of fragmentation will decrease because
of the non-uniform distribution of holes on that section.
The analysis of the effect of controllable blast parameters on fragmentation using the
literature review lead to the following conclusions:
1. For the effect of powder factor on fragmentation, it was found:
- the predicted behavior of characterictic fragment size (63.9% passing)
with powder factor match well with that predicted by Kuznetsovs
equation (Kuznetsov, 1973; Cunningham, 1983, 1987; Lownds, 1983).
- a decrease of the uniformity of fragmentation was found with decreasing
powder factor (Lownds, 1983).
2. The drilling pattern has the following effect on fragmentation:
- staggered pattern gives lower characteristic fragment size compared to the
square pattern for the same powder factor, because of the better
distribution of explosives in the rock mass in the former case.
- staggered pattern gives more uniform distribution (higher value of the
Rosin-Rammler exponent n).

28


Figure 2.10 Effective circles around holes for a square pattern: a) no deviation in drilling;
b) 30 % of the burden deviation in drilling; by Lownds (1976).


29


Figure 2.11 Increased radius of effective circles: a) no deviation in drilling; b) 30 % of
the burden deviation in drilling; by Lownds (1976).



30
3. A greater borehole diameter produces a better and more uniform fragmentation
for a given blast pattern.
4. Inaccuracy in drilling had a negative effect on uniformity of fragmentation.
However, no effect on the characteristic fragment size was observed, except in the
case of low usage of explosives where the characteristic fragment size was found
to increase as the drilling deviation was increased.

2.4 Effect of Discontinuities on Rock Fragmentation By Blasting

Rock properties are the uncontrollable variables in blast design. Blast performance is
influenced by geologic structure and rock strength. In almost every mining practice, the
rocks are far from homogeneous. There are joints, bedding planes, mud or soft seams,
which have a major effect on blasting performance. These are defined as planes of
weakness within a rock mass along which there has been no visible movement. There
will be a difference in transmission of the stress waves through the joints depending on
whether the joint is tight, open or filled (Obert and Duvall, 1950; Goldsmith, 1967). Tight
joints do not affect the transmission of stress waves whereas the open and filled joints
introduce an acoustic impendance mismatch and reflect the stress waves. If the reflected
wave is sufficiently strong, internal spalling takes place. The radial cracks, which the
strain wave would have formed in a continuous rock, are prematurely interrupted by the
joint.
Many investigators have studied the effect of discontinuities on rock breakage
induced by blasting. A brief review of their results is presented below:

31
- Fourney et al., (1983) has found in his model scale experiments a joint initiated
fragmentation mechanism. For a layered medium this mechanism of joint initiated
cracking yields a much smaller average fragment size than would be obtained in a
homogeneous media. This reduction in fragment size is at least 1.5 times.
- Da Gama (1983) found in full- scale bench blasts that less energy is required to
fragment a discontinuous rock than a homogeneous rock and used the Bonds
third law of comminution to estimate this energy reduction.
- Harries (1983) in full-scale bench blasts found that any increase in the mean
spacing between joints and/or bedding planes partings demands that a greater
degree of new breakage is created in a blast. An increase in the degree of fissuring
usually encourage the use of greater burdens, blasthole spacings and collar (or
stemming) length and correspondingly lower energy factors.

Ash (1973) stated that better fragmentation occurs when drill holes are oriented
along lines perpendicular to the most prominent joint face of the rock mass. Large
fragments result from those lines of drill holes parallel to that joint face.

2.5 Methods of Study of Rock Fragmentation by Blasting
In general there are three methods by which rock fragmentation by blasting can be
studied:
1. Using a full- scale blasting process as a model
2. Using scale models
3. Using numerical models.

32
The actual full-scale blasting operation can be too large for experimentation. For
most small-scale model blasting, sizing of 100% of the blasted material with sieves for
particle size determination is usually feasible. However, sieving is not possible with full-
scale production blasts although a photographic method can be used instead.
Photographic methods of evaluating fragmentation are currently successful
(measurements of section through the muckpile) but still the relation between the
distribution measured by photographic or image analysis and the actual particle size
distribution (measured sieving) has to be found. In general, full-scale experimentation is
expensive and time consuming, and only a limited number of parameters can be varied.
Models from concrete, rock or photoelastic materials such as plexiglass have been
used by many workers to study rock fragmentation by explosives or fracturing and crack
propagation problems (Fourney et al., 1983; Bjarnholt and Skalare, 1983).
However, scale models must make the following assumptions:
(a) cracking in photoelastic materials used as models is similar to cracking in rock
(Harries and Hengst, 1977);
(b) the effect of discontinuities on the fragmentation process in model blast is similar
to that in full scale blasts; and
(c) the actual structure of the rock, as determined by the existing discontinuities,
cannot be simulated in model scale blast.
Many industrial explosives will not detonate reliably in small diameters because the
critical diameter depends on the type of explosive. Each explosive has its own diameter-
velocity curve. If the borehole diameter is less than the critical diameter, the detonation
process will not support itself and will be extinguished. As a result PETN based

33
explosives have to be used. This applies particularly to the experimentation and
evaluation of aluminized explosives where the aluminum requires a finite time to react
(Porter, 1974).
Numerical modeling does not have the disadvantages of the above two methods.
Finite element codes were developed that model both the shock and stress wave
propagations through the rock and the nucleation and growth of cracks in the affected
rock mass (McHugh, 1983; Margolin and Adams, 1983). While still a long way from use
in routine blast design, these approaches help direct experimental work in a cost-
effective way, and greatly improve our understanding of fundamentals mechanisms.
An example of the more empirical models which can be used was given by
Cunningham (1983). His model incorporates Kuznetsovs work (1973) on relating
explosive energy, hole size and rock characteristics to mean fragment size, and the Rosin-
Rammler curves for assessing fragment size distribution. This approach is used
extensively by AECI, the major South-African industrial group, for designing blast and,
providing the chose of values for the rock is correct, gives a fairly good match to data
obtained in actual tests.
Harries and Hengst (1977) constructed a digital simulation model to study rock
fragmentation due to blasting. This model was the basic for further models that can be
used in routine blasting work such as the SABREX program (Scientific Approach to
Blasting Rock by Explosives), and Lownds FRAG model (1983). In these models
various assumptions were made for the propagation of cracks and these were
programmed into a simulation model of the blasting area.

34
BLASPA has been used by Favreau (1983) and others for modeling blasting for
many years. Recently, Favreau (1993) has described some of the aspects of the swell
module used in the model and the results obtained. This model can be only applied to
description of particle motion during a blast.
The spherical element computer program DMC_Blast, developed by Preece in 1989,
has been modified a few times. A new version of that program (Preece et al., 1997)
performs coupled gas flow and rock motion simulations in a bench blasting environment.
Several different equations of state are included for modeling the behavior of the
explosive.
The equation of state for explosive gases traditionally used by the British company
Imperial Chemical Industries, also known as ICI, allows modeling of many different
explosives. The program muckpile contours are in good agreement with those observed
in the field (Preece et al., 1993).

2.5.1 Kuz-Ram model

An empirical equation of a relationship between the mean fragment size and applied
blast energy per unit volume of rock (powder factor) has been developed by Kuznetsov
(1973) as a function of rock type. He reported that initial studies had been done with
models of different materials and the results were later applied to open pit mines and an
atomic blast. A degree of scatter between fragmentation measurements and prediction
was shown, and was to be expected, considering the nature of mining and the variability

35
of rock. The model predicts fragmentation from blasting in terms of mass percentage
passing through versus fragment size. His equation is given below

X
m
= A
8 . 0
0
|
|
.
|

\
|
Q
V
Q
1/6
(2.5.1.1)

where:
X
m
- mean fragment size, cm;
A - the rock factor, 7- for medium hard rocks, 10- for hard highly fissured
rocks, 13- for hard, weakly fissured rocks;
V
0
- the rock volume (cubic meters) broken per blasthole = Burden x Spacing
x Bench Height;
Q - the mass of TNT containing the energy equivalent of the explosive charge
in each blasthole, kg.

The relative weight strength of TNT compared to ANFO (ANFO = 100) is 115.
Hence equation (2.5.1.1) based upon ANFO instead of TNT can be written as
(Cunningham, 1987):
X
m
= A
8 . 0
0
|
|
.
|

\
|
e
Q
V
Qe
1/6

30 / 19
115

|
|
.
|

\
|
anfo
S
(2.5.1.2)
where:
Q
e
- the mass of explosive being used, kg;
S
anfo
-

the relative weight strength of the explosive to ANFO (ANFO =100)


36
Since

K Q
V
e
1
0
= (2.5.1.3)
where K - the powder factor (specific charge), kg/m
3


Equation (2.5.1.2) can be rewritten as
X
m
= A (K)
0.8
Q
e

1/6

30 / 19
115
|
|
.
|

\
|
anfo
S
(2.5.1.4)

Equation (2.5.1.4) can now be used to calculate the mean fragmentation (X
m
) for a
given powder factor. Solving equation (2.5.1.4) for K:
K =
25 . 1
30 / 19
6 / 1
115

|
|
.
|

\
|
anfo
e
m
S
Q
X
A
(2.5.1.5)

One can calculate the powder factor required to yield the desired mean
fragmentation. Cunningham (1983) indicated that in his experience the lower limit for A
even in very weak rock types was A = 4 and the upper limit was A = 12.
In an attempt to better quantify the selection of rock factor A, the Blastability Index
initially proposed by Lilly (1986) has been adapted for Kuznetsovs model (Cunningham,
1987). Cunningham stated that the evaluation of rock factors for blasting should at least
take into account the density, mechanical strength, elastic properties and structure. The
equation is given below:


37
A = 0.06 (RMD + JF + RDI + HF) (2.5.1.6)
where:
RMD: rock mass description
- powdery/ friable 10
- vertically jointed JF
- massive 50
JF: vertical joint factor
JF = JPS + JPA (2.5.1.7)
JPS: vertical joint plane spacing
- 0.1 m 10
- 0.1 m to MS 20
- MS to DP 50
MS: oversize, m
DP: drilling pattern size, m
JPA: joint plane angle
- dip out of face 20
- strike perpendicular to face 30
- dip into face 40
RDI: rock density influence, tons/m
3

RDI = 25 RD 50 (2.5.1.8)
HF: hardness factor
Y: Youngs modulus, GPa
If Y < 50GPa, HF = Y/3

38
If Y > 50GPa, HF = UCG/5
UCG: uniaxial compressive strength, MPa

Then the Rosin-Rammler formula is used to predict the fragment size distribution. It
has been generally recognized as giving a reasonable description of fragmentation in
blasted rock. This equation is:

n
x
x
m
c e R
|
|
.
|

\
|

=
(2.5.1.9)
where:
R
m
- the proportion of material retained on the screen;
X - the screen size, cm;
X
c
- the characteristic size (scale factor), cm;
n - the index of uniformity, a blasting parameter.

The characteristic size (X
c
) is that through which 63.9 % of the particles pass. If the
characteristic size (X
c
) and the index of uniformity (n) are known then a typical
fragmentation curve can be plotted.
Equation (2.5.1.9) can be rearranged to yield the following expression for the
characteristic size

n
m
c
R
X
X
1
ln
=
(2.5.1.10)
Since the Kuznetsov formula gives the screen size X
m
for which 50% of the material
would pass, substituting these values:

39
X = X
m

R = 0.5
into equation (2.5.1.10) one finds that

n
m
c
X
X
693 . 0
=
(2.5.1.11)

A useful indirect check on the index of uniformity has been done by Cunningham
(1983). He based his prediction of fragmentation on the Kuznetsov equation and used the
relationship between fragmentation and drilling pattern to calculate this blasting
parameter of the Rosin-Rammler formula. The blasting parameters, n, is estimated by:

|
.
|

\
|
|
.
|

\
|

|
|
|
|
.
|

\
|
+
|
.
|

\
|
=
H
L
B
W
B
S
D
B
n 1
2
1
14 2 . 2
5 . 0
(2.5.1.12)
where:
B - burden, m;
S - spacing, m;
D - borehole diameter, mm;
W - standard deviation of drilling accuracy, m;
L - total charge length, m;
H - bench height, m.
. The interrelationship between the variables in a blast and the results of blasting
established in this work are in good agreement with experiments. (Cunningham, 1983,
1987).


40
2.5.2 Hole-by-hole analysis method

In bench blasting, the actual drilled pattern is not the exact designed pattern because
the drilling can cause some unexpected fragment sizes. The goal of the hole-by-hole
method is to be able to predict the fragment size distribution and effect of faulty drilling
when the exact drilling pattern is known. Each borehole is assigned a volume, that it has
to fracture for the actual drill pattern used. The fragmentation distribution of each
blasthole is calculated and they are summed up to give the total fragment size
distribution. This method is used to consider the effect of drilling pattern.
Some rules of dividing the volume for each hole have been suggested by Hjelmberg
(1983). He believed that those rules could improve the prediction. The triangle was
chosen 90 degrees for simplicity. However, the real angle of breakage in the triangle
depends on the distance to the nearest free face and on the distance to the nearest holes,
that detonate simultaneously (Bhandari, 1975b).
The literature review showed that the effect of variations in drilling and ignition
patterns can be added to the present formulas. With the idea of finding the volume of
each hole, this volume is used in the Kuz-Ram model to predict the fragment size
distribution of each hole. Summing up the total fragment size distribution would take care
of the influence of drilling and ignition patterns.





41
Chapter 3 Development of the Proposed Model FRAGM



Optimum blast fragmentation is fundamental to all phases of the surface mining
operation. Changes in the blast design may affect the efficiency and the productivity of
downstream processes such as loading, hauling, crushing and milling, resulting in cost
savings or losses. The basic steps of blasting engineering are design, implementation and
observation of the blast results. From the literature review it was found out that previous
researchers have calculated the fragmentation and the size distribution by considering
rock properties, explosive properties, and the influence of actual drilling pattern.
The influence of true burden is very important in rock fragmentation. The true
burden is one of the most important parameters for the execution of an efficient blasting.
Measuring the true burden has, until recently, been a difficult and almost an impossible
operation. One way of doing this was to use a long stick and string hanging over the face
of a bench to establish precisely the location of rock face in relation to the front line of
blastholes. This way, even a qualified surveyor, using standard surveying equipment
would find it difficult to obtain an accurate profile of the rock face.
As a result many researchers assumed that the rock volume in the first row of
blastholes could be calculated by multiplying the burden, spacing and height of a bench,
and consequently irregularities of a bench face were not taken into account (Fig. 3.1).
However, in the real life this procedure could produce a good agreement with the results
of a blast if all holes in the first row were drilled at an angle equal to the bench face angle
(Fig. 3.2). It simply means that the burden will be the same at different parts of a bench
such as the toe and the crest.


42




















Figure 3.1 Diagram of a blasting pattern and the geometry of charged blastholes.





43
















Figure 3.2 Typical representation of the drillhole inclination versus the bench face
angle presented by Olofsson (1990).








44
Generally many mining operators prefer to use vertical blastholes in their surface
mines as inclined drilling has many disadvantages, and some of them are listed below:

1) increased deviation when drilling long blastholes;
2) difficulty in positioning of the drills in collaring operations;
3) increased drilling length;
4) necessity of close supervision which creates work lapses;
5) lower drill feed, which means that in hard rock the penetration rate is limited in direct
proportion to the angle of inclination of the mast;
6) more wear on the beads, drill steel and stabilizers;
7) less productivity with rope shovels due to the lower height of the muckpile;
8) poorer flushing of drill cuttings due to the friction forces, requiring an increase in
airflow;
9) problems in charging an explosive, especially in blasthole with water.

As one can see from Figure 3.3 and Figure 3.4, the burden varies and commonly has
the larger value in the bottom part of a bench. Due to the nonuniform burden, the actual
powder factor, in general, is smaller than the one computed by the estimating the rock
volume assuming a vertical bench face.
Better prediction of rock fragmentation can be achieved by a hole-by-hole analysis
for the first line of blastholes. The critical parts in the hole-by-hole analysis are: 1) how
to estimate the rock volume in the first row of blastholes; and 2) how to divide the
volume for each hole.

45












Figure 3.3 Variation of burden in case of vertical hole on an inclined face by
Bhandari (1997).




46




















Figure 3.4 Parts of a bench presented by Hustrulid and Kuchta (1998).




47
To accurately estimate the rock volume for the first row of blastholes, the face
profile must be known as realistically as possible. One way of doing this is by using a
laser profiler, which can receive a beam reflected directly off the rock (as opposed to
using specially designed reflectors). It is even possible to locate the collars of the holes
already drilled or to be drilled. Where the blast holes have been already drilled, the true
profile for each hole could be obtained. This method produces a better quantitative profile
of the bench face and leading to the possibility of making a better prediction of rock
fragmentation. This method is described in details in Appendix B of this thesis.
The second critical part with the hole-by-hole analysis is dividing the volume for
each hole. The assigned 90 degrees angle of breakage is inadequate and most often too
small for each borehole in the first line in bench blasting. The angle of breakage depends
on the rock properties, burden, spacing and the quantities of explosive charges applied
(Bhandari, 1975a, 1975b, 1996, 1997). However, calculating the angle of breakage alone
is not sufficient to find the bench volume broken by each borehole in the first line.
Joints and discontinuities have a tendency to influence the angle of breakage. In this
situation, the predicted angle of breakage can be different from the actual one. In some
cases, if there are a number of badly weathered joints in the bench, the actual angle of
breakage is determined by the angle between the sets of joint faces.
In order to find the bench volume broken by each hole on the first row, it is
proposed to divide the bench into different areas of breakage according to the different
mechanisms of rock breakage for each borehole (Fig.3.5). This way, two different areas
for each borehole could be identified and shown in Figure 3.5. They are the main
breakage area and the secondary breakage area.


48
Plan view

Case 1: Unbroken rock between two boreholes








Case 2: Completely broken rock between two boreholes





Figure 3.5 Two areas of breakage.




49
The main breakage area is due to the reflected stress waves at the free face and the
discontinuities, and also due to the redistribution of the pressure that enlarge fractures
when the burden is uncoupling from the main rock mass. Rock is weaker in tension than
in compression. Hence, when the reflected shock waves pass through the rock, they cause
tensile failure around the discontinuities in the rock. Most of the rock breakage occurs in
tension. Therefore, the tensile strength of the rock is a key factor in determining the
resistance of rock. Worsey and Rustan (1987) pointed out that the compressive strength
of rock has no major effect on the blasting performance and that the tensile strength,
estimated by the Brazilian test, is related to the blasting performance.
The detonation wave is a strong shock wave deriving the energy from the chemical
reaction of the explosive composition. These waves cause crushing and compression of
the rock depending on the explosion pressure and the strength and stiffness of the rock.
The stress waves act proportional to the pressure generated by the explosive. When the
stress waves hit the discontinuity face or the free face, the stress waves get reflected and
become tensile. The breakage occurs when the stress exceeds the tensile strength of the
rock. If the magnitude of the stress wave is small, the reflected tensile stress will not
cause any breakage, and the stress will travel as a seismic wave. The main breakage area
is defined between the blastholes in this tensile breaking area. Then the rock starts
uncoupling from the rock mass and is pushed forward by the borehole gas pressure.
The stress wave behavior explained above can be estimated by the empirical
equation (A.9) developed by Adushkin (1987) and described in details in Appendix A of
this thesis. This equation may be used for calculating the location of breaking points

50
associated with the above explained process of rock fragmentation. Connecting the
breaking points to the borehole will give the angle of the main breakage area.
The breakage of rock in the secondary breakage area is primarily due to the
reflection of stress waves at the discontinuities in the main breakage area. Also due to the
intersection of the stress waves generated by different boreholes, a compressive stress
wave is converted to a tensile stress wave. The numbers of fractures in the secondary
breakage area may be the same as they are in the main breakage area. However, the rate
of load release in this area is lower than in the main breakage area. As a result, the
fragment size in this area is commonly larger than in the main breakage area.
From Figure 3.6 one can see that there are three cases of the main breakage area in
relation to the burden and the amount of explosives used: 1) the borehole is too far away
from the bench face; 2) the borehole is too close to the bench face; and 3) correct burden
and powder factor.
If the borehole is too far away from the bench face, the powder factor is too low to
completely break the rock between the bench face and the borehole. The detonation of
the explosive charge will only create a crater around the borehole. When the shock wave
hits the bench face, it is too weak to cause any tension failure at the bench face, and,
therefore, leaves the bench unbroken. Consequently, the gas pressure will only be able to
push the rock upward instead of forward until the borehole pressure is released.
If the borehole is too close to the face, the angle of breakage is very large. Since the
burden is very small, the borehole pressure can be very easily released and, hence, does
not have enough time to enlarge the tension cracks initiated by the shock waves. A lot of


51
Plan view

Case 1: Borehole too far away from the bench face







Case 2: Borehole too close to the bench face





Case 3: Correct burden and powder factor





Figure 3.6 Three cases in main breakage area.




52
explosive energy may be wasted in throwing a small quantity of rock, and flyrock
problem can occur in such situations.
If the burden and powder factors are chosen properly, the burden allows much better
utilization of stress waves and gas energy, resulting in better fragmentation. In a typical
bench blasting, there are tens of holes detonated simultaneously. In most cases, an
overlap between two adjoining main breakage areas is expected. In such cases, dividing
the overlapped area by two and adding it to the remaining main breakage areas seems a
reasonable approach.

3.1 Overall breakage

The purpose of defining the main breakage area and the secondary breakage area for
the first row of boreholes is to calculate these areas for each hole in the first row in a
quantitative way. First, the main breakage points can be found by using equation (A.9).
Then, connecting the breaking points on each side with the borehole location gives the
triangle of the main breakage area. By repeating the same procedure, the main breakage
area can be found for each borehole in the first line. These areas should be different in
different areas of the bench such as the toe and the crest. After the main breakage areas
are found for all holes in the first row, the average fragment size can be calculated for
every hole in the first row by using the Kuznetsov equation (2.5.1.2). By summing up the
average size for each hole in the first row, the average fragment size and size distribution
can be calculated for the first line of blastholes. Further calculations of the blasted rock
fragmentation are done using the Kuz-Ram method, described in Section 2.5.1. The

53
example of step by step calculations of the fragment size distribution for a given blast is
presented in details in Appendix A of this thesis.
Given below is a summary and the sequence of steps followed in the proposed
FRAGM model to determine the size distribution for a given blast using the bench profile
data to estimate the volume of rock to be blasted:
1) using the bench profile data to estimate the volume of rock to be blasted;
2) using statistical tools the average crest burden for a given set of the bench
profiles data is determined;
3) the results of the step 1 and step 2, provide information to estimate the average
bench face profile for a given blast;
4) divide the blasted volume into different boundary areas corresponding to each
row of boreholes. The boundary of these areas will be the face of a bench and the
axis of the borehole rows. For the first row of blastholes, the boundaries of this
areas will be the face of the bench and the axis of the first row of boreholes. For
the remaining row of blastholes, the boundaries are between the axis for a given
row of blastholes and its previous row;
5) use the blastability index method, initially proposed by Lilly (1986), to calculate
rock factor A by using empirical equation (2.5.1.6);
6) calculate the angle of breakage based on the rock properties, explosive
properties, explosive charge and drilling pattern for each hole in the first row. In
some cases, if there are a number of badly weathered joints in the bench, the
actual angle of breakage is determined by the angle between the sets of joint
faces. In such cases the predicted angle of breakage can be different from the

54
actual one and, hence, the geologic structure must be considered before designing
a blast and using the results of observation for the fragmentation prediction
calculations;
7) divide the blasted volume for each hole in the first row by the boundary
conditions. Calculate those areas (the main and secondary breakage areas);
8) predict the mean fragment size for each hole in the first row by using equation
(2.5.1.2);
9) predict the average fragment size and size distribution for the first row of
blastholes by using the method described in Section 2.5.1;
10) calculate the average fragment and size distribution for a given blast by using the
method described in Section 2.5.1.













55
Chapter 4 Description of the Rock Fragmentation Prediction
Engineering Model FRAGM


Proper blasting designs require careful consideration of many variables. To
determine the effect of simple design changes, iterative calculations are required. To
reduce the time in calculation, an engineering model FRAGM, which is based on the
methodology and algorithm proposed in the third chapter, is used to predict the size
distribution for a given blast. The time required for calculations reduces from several
hours to several minutes and reliable prediction can be obtained. With the capability of
modern computer to perform tremendous amount of computations in a short period of the
time, FRAGM can calculate the blasting performance in an efficient manner, which
allows the mine engineer to estimate the blasting effect and optimize blasting round
design.
Some of these variables are controllable, such as the explosive properties and the
drilling pattern. The rock properties and bench geometry are uncontrollable factors for a
given blast. Other uncontrollable parameters are joints, fractures, tensile strength, rock
density, P-wave velocity and the true burden. The main task for mine engineer is to find
the optimal values of controllable factors to obtain the best fragmentation.
The basic steps involved in choosing proper blast parameters that ensure desired rock
fragmentation in FRAGM model are depicted in Figure 4.1. Though some degree of
flexibility is possible with regard to having a suitable bench profile that provides
optimum fragmentation, often stability consideration limit this possibility. Similarly, once
a curtain kind of drilling equipment is chosen, the hole diameter also becomes fixed.


56





















































Figure 4.1 Procedures of blast round design and evaluation of blasting performance.

Profiling Bench Face
Bench Shape Survey
Rock Volume Estimation
Borehole Size and
Type Selection
Explosive Selection
Drilling Pattern
Blasting Performance
Prediction
By Using FRAGM
Improved Drilling
Pattern Design,
Charge Design, or
Explosive Selection
NO
Accept Blast
Round Design
YES
Satisfied ?

57
Hence, the alternative lies in varying the explosive type or drilling pattern to get the
desired rock fragmentation. The FRAGM model provides an easy and reliable way to
examine the effect of changing the blast parameters on rock fragmentation without
resolving to expensive and time consuming field trials. Also, it helps the blast designer to
examine the degree of fragmentation obtained in a blast on a routine basis. This chapter is
a general description of the input and output data of FRAGM.

4.1 Bench and borehole information
To determine the rock fragmentation, the input data of FRAGM consists of bench
and borehole information. The bench information contains the bench face profile data and
the location of the boreholes. An X-Y-Z relative coordinate system is utilized to reference
the bench (Fig. 4.2). The left upper corner is (0,0,0) and X values increase from left to
right. Y values increase toward the bench face. Z values increase downwords. The
borehole information contains loading density, powder factor, diameter of the borehole,
diameter of the charge column, length of the powder column, length of the borehole,
stemming length, and X-Y coordinates of each hole. FRAGM uses these groups of
information to determine the actual burden, spacing, angle of breakage, and rock volume
being broken in a given blast.

4.1.1 Burden
Burden is the nearest distance from a blast hole to a free face or between rows of
blast holes. Experience and empirical equations obtained from field tests can lead to an
appropriate distance for burden.


58




Plan View

















Figure 4.2 X-Y-Z coordinates data of the bench boreholes.

59
Burden is determined by the strength of rock in the bench, geometry of the bench
and the bench face, and the specific explosive energy applied, which in turn is related to
the diameter of a charge, weight strength and loading density, and spacing-to-burden
ratio. The observation by Preece et al., (1991) using motion picture photograph proved
that premature gas leakage in small burden occur. Therefore, large burdens with heavier
charge is recommended only for casting because it will reduce the possibility of
premature leakage of gases from the blast hole through the fragmented rocks while
providing sufficient heave energy to throw the overburden material. The push on the
overburden material by the gas pressure in the blast hole will be at its most effective
situation.
It is necessary to measure the true burden in the first row of boreholes for effective
prediction of rock fragmentation. In this model, the bench face profile data are collected
and stored. Then the surface area of each profile is calculated and analyzed. As a result
one can estimate the actual volume in the first row of holes and evaluate the powder
factor distribution. It is very important to remember that the uniformity of fragmentation
is found with decreasing powder factor (Lownds, 1983).

4.1.2 Spacing
Spacing is the distance between adjacent blast holes in the same row. In most cases,
evidence show that spacing should not be greater than two times the burden in order to
make a clean uniform free face whether it is for conventional bench blasting or casting.
Spacing smaller than the burden can be efficient in blasting of very hard rock (Shapurin
and Kutuzov, 1990).

60
Traditionally burden, spacing, and bench height determine the rock volume broken
by each blasthole. However, in a typical bench this estimation cannot be applied for the
first row of blastholes unless the boreholes position are parallel to the bench face. The toe
burden, the crest burden, the spacing, and the bench height should all be considered to
determine the rock volume broken by each borehole in the first line of blastholes. A good
and useful model must be able to consider the actual burden and spacing. Especially this
is a case in the first row of blastholes. The design values of the first line of blastholes are
often misleading. The only way to find those values is by examining them hole-by-hole
in the actual bench. The hole-by-hole analysis method can measure the effects of burden
and spacing, and determine the angle of breakage of rock broken.

4.1.3 Powder factor
Powder factor represents the amount of explosive energy applied to each specific
volume of blasted material. It can be expressed in terms of pounds of explosive used for
each cubic yard of material or in kilograms of explosive used for each cubic meter in the
bench. It is the indicator of consumption of explosive in a shot. In general, the value
varies between 0.9 and 1.5 pounds per cubic yard. A heavier powder factor not only
provides a better fragmentation, but also provides energy for throwing, which is truly an
advantage to surface coal mining operations.

4.1.4 Loading density
The explosive density is one of the important properties that should be considered in
blast design. The density of most commercial explosives ranges from 0.8 to 1.6 gram per

61
cubic centimeter. ANFO and aluminized ANFO are in the low density range 0.8 to 1.15
gram per cubic centimeter. Cartridge explosives (slurry or dynamite) are in the high
density range 0.9 to 1.6 gram per cubic centimeter.
Loading density is commonly measured by dividing the weight of explosive over the
borehole volume. Loading density affects the explosive quantity in the borehole,
detonation velocity, pressure, and energy. However, an explosives sensitivity can be
reduced or destroyed by an excessive increase in density. If the density becomes too high,
exceeding the critical density, the explosive will not detonate. This phenomenon is called.
dead pressed.
The loading density is determined by the loading equipment and the skills of the
loaders. Since these factors are fixed, the loading density in the same bench should not
differ significally from borehole to borehole.

4.1.5 Diameter of blastholes
Blastholes are drilled in the bench to contain a charge column as well as the
stemming material. When a blasthole is coupled with the charge column, the effect of a
shot is enhanced. This is a desirable situation. Therefore, the larger the diameter of a
blasthole, the more explosive can be loaded into the borehole per linear distance. The
practical range of diameter is between 7 and 14 inches. For casting design, a large
diameter is preferred because a higher powder factor is required. However, the diameter
of drillholes is not the only contributor to a high powder factor. Loading density, burden,
and spacing may also change the powder factor.

62
The blast designer has to remember that each explosive has its own diameter-
velocity curve. If the borehole diameter is less than the critical diameter, the detonation
process will not propagate. The effect of borehole diameter on detonation properties is
significant. Most blast designers are aware of this and avoid using small borehole
diameters. The practical range of borehole diameter in surface mining practices is from 5
to 14 inches, which produces a detonation velocity from 4000 to 5000 meters per second
for ANFO. The difference of detonation velocity and pressure within this range is not
significant.

4.1.6 Stemming
Stemming distance refers to the top of the blasthole normally filled with inert
material to confine the explosive gases. In order for a high explosive charge to function
properly and release the maximum energy, the charge must be confined in the borehole.
Adequate confinement is also necessary to control airblast and flyrock. In most cases, a
stemming distance of 0.7 times the burden is adequate to keep the material from ejecting
prematurely from the hole. It must be remembered that stemming distance is proportional
to the burden. Therefore charge diameter, specific gravity of explosives, and specific
gravity of rock are all needed to determine the burden, and stemming distance is also a
function of these variables. If drilling cuttings are used as a stemming material the
stemming distance equals to 0.7 times the burden may not be adequate to keep the
stemming from blowing out (Konya and Walter, 1985).
On the other hand, doubling and tripling the stemming distance may not ensure the
holes to function properly. If the stemming distances are excessive, poor top breakage

63
will result and the amount of backbreak will increase (Konya and Walter, 1985). The
common material used for stemming is drill cuttings, since they are conveniently located
at the collar of the blasthole. However, very fine cuttings commonly called drilling dust
make a poor stemming material. In case such as this, it is common to bring crushed stone
to the job site to use as a stemming material. If drilling dust were used instead of
crushed stone or drilling chips, it may be necessary to increase the stemming depth to
equal the burden distance. Drilling dust makes a poor stemming material since it will
not lock into the borehole walls and is easily ejected. It was also pointed out that the
optimum size of stemming material would be material that has an average diameter of
approximately 0.05 times the diameter of the blasthole and the material must be angular
to function properly (Konya and Walter,1985).

4.1.7 Subdrilling
Subdrilling is the depth that blastholes will be drilled below the proposed grade to
ensure that breakage will occur to the grade line. Blastholes normally do not break to full
depth. Most surface mines are using subdrilling. The subdrilling will lead to a result of a
flat bottom in an excavation. If drilling is done slightly deeper than required and some
holes are too deep at the time of loading, the blaster can always place drill cuttings in the
bottom of those holes to bring them up to the desired height. The blaster, however, does
not have the ability at the time of loading to remove excessive cuttings or materials that
has fallen down into the hole.



64
4.2 Explosive types
There are basically four different types of commercial explosives, i.e. emulsion,
Heavy ANFO, ANFO, and ALANFO (Table 4.1). Each explosive has its own detonation
velocity, pressure, strength and energy. These properties vary with borehole diameter,
temperature, loading density, and weather conditions. The ideal detonation velocity
varies from 3,750 to 6,000 meters per second, and the detonation pressure from 0.055 to
0.13 Mega bars, for different types of explosives.
Comparing the different types of explosives in field conditions and selecting the
right type of explosive suitable for the specific field conditions are the major tasks of the
mine engineers and blasters. FRAGM can be used to predict the performance of
different explosives that can be very helpful in the changing surface mine environments.

4.3 Rock strength
Rock strength is an important factor in blasting performance. Rock has three
different types of the strength: compressive, tensile, and shear strength. Rock is much
weaker in tension than in compression. When the shock waves pass through rock, they
cause tensile failure around the discontinuities in the rock. Most of the rock breakages
occur in tension. The tensile strength of rock is the key factor to determine rock
resistance to blasting.
Compression failure only occurs around the borehole. Worsey and Rustan (1987)
stated that the compressive strength of rock has no major effect on blasting performance
in surface mining operations. They also pointed out that the Brazilian tensile strength is
related to blasting performance.


65










Table 4.1 Explosive strength based on composition

Explosive
name
Water
%
AN
%
CN
%
FO
%
Al
%

AN(pp)
%
Density
g/cc
Strength
Emulsion 15.00 69.50 9.00 6.50 0.00 0.00 1.20 78-91
Emulsion 15.00 79.50 0.00 5.50 0.00 0.00 1.20 80-94
Emulsion 14.75 75.50 0.00 5.25 5.00 0.00 1.20 97-109
Emulsion 13.50 71.50 0.00 5.00 10.00 0.00 1.20 112-126
Heavy
ANFO
9.00 38.00 0.00 6.00 0.00 47.00 1.30 89-108
Heavy
ANFO
9.00 38.00 0.00 6.00 0.00 47.00 1.10 88-100
Heavy
ANFO
9.00 25.50 12.50 7.00 0.00 46.00 1.30 88-103
Heavy
ANFO
9.00 25.50 12.50 7.00 0.00 46.00 1.10 87-96
Heavy
ANFO
12.00 33.00 0.00 8.00 0.00 47.00 1.30 82-98
Heavy
ANFO
12.00 33.00 0.00 8.00 0.00 47.00 1.10 79-91
ANFO 0.00 94.00 0.00 6.00 0.00 0.00 0.80 100
ANFO 0.00 97.00 0.00 3.00 0.00 0.00 0.80 68-77
ANFO 0.00 92.00 0.00 8.00 0.00 0.00 0.80 92-96
ANFO 0.00 94.00 0.00 6.00 0.00 0.00 0.90 100-105
ALANFO 0.00 90.00 0.00 5.00 5.00 0.00 0.90 110-120
ALANFO 0.00 86.00 0.00 4.00 10.00 0.00 0.90 118-132








66









Table 4.2 Rock properties published by Mohanty (1987)


Rock type

Granite

Quartz-diorite

Griesen

Limestone

Density (g/cc)

2.58

2.81

2.84

2.78

Longitudinal Wave (m/sec)

5,110

5,000

5,570

5,100

Shear Wave (m/sec)

3,020

3,240

3,610

3,230

Poissons Ratio

0.23

0.14

0.14

0.17

Youngs Modulus (GPa)

58.0

45.2

84.2

67.5

Compressive Strength (MPa)

125

180

215

198

Brazilian Tensile Strength (MPa)

8

15

16

11

Dynamic Tensile Strength (MPa)


32

56

67

51







67







Table 4.3 Rock properties presented by Cook (1976)


Rock Type

Marble

Limestone

Granite

Sandstone

Shale
Density
(g/cc)
2.70-2.90 2.30-2.50 2.60-2.70 2.10-2.60 2.70
Longitudinal Wave
(km/s)
5.3-6.4 2.9-5.0 5.6 3.05 ------
Poissons Ratio ------- 0.24-0.32 0.20-0.30 ------ 0.26
Youngs Modulus
(GPa)
20-100 10-70 50-90 5-45 70-100
Compressive
Strength (MPa)
60-250 30-250 150-290 30-240 70-230
Brazilian Tensile
Strength (MPa)
2-6 3-8 7 9 ------



Rock Type

Greenstone

Basalt

Taconite

Quartzite
Density
(g/cc)
3.02 3.00 3.23-3.44 2.17
Longitudinal Wave
(km/s)
5.2 6.6 5.1-5.9 5.0
Poissons Ratio 0.15 0.33 0.26-0.24 0.28-0.15
Youngs Modulus
(GPa)
80 85 91-102 69
Compressive
Strength (MPa)
300 80-360 330-340 380
Brazilian Tensile
Strength (MPa)
------ 15 20-30 18







68
FRAGM uses the Brazilian tensile strength value to predict the blasting performance.
The rock property information is usually obtained by laboratory testing. Mohanty (1987)
presented the rock properties for four types of rock as shown in Table 4.2. Cook (1976)
also presented the rock properties for nine types of rock as listed in Table 4.3.

4.4 Delay time
The availability of short-delay initiating makes it possible to create desired firing
intervals among charges. The rule of thumb for delay of initiation between rows is
generally one to two milliseconds for every foot of burden in conventional blasting.
In case of casting design, the tendency is to increase the delay to provide appropriate
room for the subsequent rows of overburden material to move. Delay of five or even ten
milliseconds for each foot of burden was reported (Favreau, 1983). Providing adequate
forward relief between rows of blastholes is very important for casting. The prolonged
delay between rows of blastholes was also identified as one factor that actually reduces
ground vibrations in surface mining operations.
A delay of initiation between holes in the same row can be arranged by using short-
delay initiating devices. However, in bench blasting simultaneous initiation of all holes in
a row can have an enhanced effect. To preserve heave energy of a shot, this is also a
sound approach. In casting design, the delay of initiation along holes in a row should be
maintained at a minimum or at zero. The studies on the effect of delay time on
fragmentation have been conducted by Konya and Walter (1985). In their experiments,
the best delay time was 2 milliseconds per foot of spacing. They stated that the effect of
delay time on fragmentation would not differ significally in the normal range of delay

69
time in blasting practice. FRAGM and the previous fragmentation models do not
consider the effect of delay time.

4.5. Geological conditions
The geological conditions of a blasted bench is the most important and complicated
factor to be examined when a surface mine is being planned and at the beginning of
designing a blast round. The strength of rock in the bench, existence of weak and
fractured layers, and dip and strike of the bench are to be considered.
An open pit should be developed to accommodate the needs of blasting and material
handling among considerations. The distribution of haulage and access roads is one
concerns of mine planning. At the same time, the requirements of successful blasting are
subject to the geological conditions of the bench and those requirements must be strictly
observed. In order to achieve the desired fragmentation, the strength of rocks, the
direction of strike and dip, and the location and direction of fractured zones or soft layers
such as clay are important factors to be evaluated. Soft strata, such as mud layers, cause
more problems than other geological features. They allow an almost instant release of the
explosion gas pressure, due to their low shear strength among the layers. The fragments
in the mud strata can be thrown a significant distance and cause poor fragmentation.
The joint direction and orientation effect blasting performance. The research work
by Ash (1973) showed that a better fragmentation occurs when drillholes are oriented
along the lines perpendicular to the most prominent joint face of the rock mass. Large
fragments result from those lines of drillholes parallel to that joint face. Local geological
conditions are often very difficult to assess. The effect of geological conditions vary from

70
borehole to borehole. Due to the complex conditions of geological structure in mining
practice, it is very difficult to predict the precise geological effect on blasting
performance. The use of empirical formulas to predict geologic effect would certainly
cause some errors, but it might best handle some general blasting effects due to
geological structure. In this model equations 2.5.1.7, 2.5.1.8, 2.5.1.9 are used to predict
the effect of the geological structures on blasting performance.

4.6. Output data description
The average fragment size and size distribution predicted for a given blast are
computed by following the algorithm in the Chapter 3. The broken areas for each
blasthole in the first row can be found on the computer screen. These areas are the major
concerns for blast designers. For an irregular bench, the broken areas for each blasthole in
the first row are different. Therefore, the average fragment size for each blasthole is
different. Sometimes, the difference might be very significant for a poor designed drill
bench. The main task of blast designers is to keep this difference small and keep the
average fragment size close to the designed value. The average fragment size is an
important factor in subsequent operations. If there are oversized fragments that require
secondary blasting, the time and money wasted in this operation can be very costly.
FRAGM can be used to predict these areas. The ones with large areas are most likely to
produce oversized fragments. Increasing the powder factor in these blasthole areas, or
drilling a new borehole into these areas might assist in correcting these problems. The
average fragment size and size distribution are presented in an Excel tabular form. The
size distribution curve can be plotted on the computer screen for a given blast.

71
Chapter 5 Verification of the Proposed Engineering Model


5.1. Introduction

There has been considerable research conducted in rock fragmentation prediction.
However, not all the published data can be used to verify the proposed development.
Most of the published data does not have enough input information about bench face
angle, rock properties, explosives types, and, in the same cases, drilling pattern. Of
course, the missing input data can be substituted by using average values. But such
substitutions will cause some prediction errors. Also, there are no published literature that
have the complete information including: 1) angle of bench face; 2) angle of breakage;
and 3) size distribution.
As a result, new field data have been used to verify the proposed development. The
effect of varying blast designs was analyzed with respect to the predicted and actual size
distributions in a limestone surface mine. Image analysis method has been used to
determine the actual size distribution by using the digital camera and imageprocessing
program SplitDesktop. The SplitDesktop is the result of over nine years of
research and development at the University of Arizona and the Julius Kruttschnitt
Mineral Research Center of Brisbane, Queensland, Australia (Higgins et al., 1999).
Visual observations of muckpiles immediately following the blasting are widely
used by mine operators to arrive at an approximation. These observations are qualitative.
In conjunction with the subsequent experience in loading and crushing, they may form an
important factor in the operators decision in the design ofblasting practice. However, for
normal everyday purposes image analysis method has the following advantages over

72
other forms of visual evaluation methods such as sieving and boulder counting: 1) it is
simple to use; 2) it gives a good approximation of the size distribution for a given blast;
3) measurements in the field are quick and less intrusive in the production process; 4) the
images obtained form a good record of the blast; and 5) the cost of equipment is
affordable.

5.2 Image analysis technique and sampling

The use of image analysis techniques for fragmentation analysis requires careful
consideration of the three stages in the process: sampling, image acquisition and image
analysis itself. Sampling concerns the taking of images that represent the blasted material
being analyzed. Image acquisition concerns taking of images which are of sufficient
quality for the intended analysis process. Image analysis refers to the measurement of
size distribution of fragments identified in the image. First the image is captured by the
analysis computer and stored as an array of picture points (pixels) of varying brightness.
Then image processing may be used to modify it to enable the computer to identify each
individual fragment. The results are then converted from a two- dimensional to a three-
dimensional parameter by empirical or stereological techniques.
Considering the analysis process as a whole, some errors will be introduced by the
use of two- dimensional images to represent three- dimensional blasted material, but the
magnitude of it can be minimized if each of the three stages is controlled. The three
parameters affecting the sampling errors are type, scale and the number of images. The
type of images refers to the location and state of the material being sampled and is the

73
most important factor in capturing a representative image. Operational constraints dictate
what methods can be used. In surface mining operations, photographs can be taken on the
muckpile surface (Nie and Rustan, 1987), at the face or of material in the rack of haul
trucks ( Maerz et al., 1986; McDermott et al., 1989 ). The specific blasting conditions can
dictate which of these is the most representative. Consideration should also be given to
the fact that random or systematic sampling is a requirement in order to minimize bias
(Gy, 1979).
At any given scale, image analysis can measure fragments within a size range
determined by the minimum resolvable size and the maximum visible size. The size
range is dependent on the image analysis technique. The minimum sizes are comparable
but for a large fragments the surface texture may cause automatic methods to detect false
edges to produce a group of small fragments. This often termed disintegration and its
occurrence depends primarily on rock texture and lighting conditions (McDermott et al.,
1989). A trade-off has to be found between taking close-up images in which fines can be
resolved and the sampling error introduced by analyzing a reduced area of blasted
material. It should also be kept in mind that to achieve an adequate sample size requires
the analysis of an increased number of images thus increasing the processing time.
Obviously, the greater the number of images, the nearer the result will be to the truth.
This concept was well presented by Nie and Rustan (1987). Empirical estimates have
been made in the past of the minimum sample size necessary to be analyzed in order to
achieve a given accuracy. However, consideration of sampling theory for particulate
materials shows that the maximum expected fragment size determines the proportion of
the material requiring to be analyzed for a given accuracy (Gy, 1979). The smaller

74
fragments size fractions require less material to be sampled for good accuracy. Most
published works in image analysis give good accuracy in the small to medium fragment
size fractions, even so it is recommended that sample sizes are maximized.

5.3. Constraints

Methods of fragmentation analysis offer quantitative measurement of blast
performance and thus opens the door to effective optimization of the blasting process. In
surface mining operations, blast optimization is of concern, but regular quality control of
the full size distribution of the product is of prime importance.
The alternatives to image analysis for fragmentation measurement are either
subjective or time consuming. Blast performance is commonly measured by visual
estimation after blasting and during loading, which can be subjective. For product quality
control, sampling and screening every day is usually carried out which cannot provide
rapid feedback and can introduce significant sampling errors.
Often the largest constraint in a surface mining operation is the taking of
photographs without disturbing production activities. This usually leads to photographic
sampling schemes that is less than ideal. Blasted material can be sampled either before
digging (the muckpile surface), during digging (at the face), or while in haul trucks. The
first two methods can lead to errors due to subjective judgement of the material to be
photographed, since more material can be seen than can be sampled. The use of haul
truck sampling is advantageous as the camera location can be fixed in a position and can
be automated, although the effect of material sorting during loading needs to be taking

75
into account. Another major constraint is the environmental conditions affecting the
quality of images. These conditions, such as poor lighting, shadows and dust, are difficult
to control in surface mines and may cause poor quality images.

5.4 Verification of rock blasting engineering model in the field


In this section, the field verification of FRAGM model is described. The field data
was obtained at a limestone surface mine in West Virginia, U. S. A. Comparison between
predicted and actual rock fragmentation was made for five different blasts from a
geologically similar area of the quarry. As all the input information required for the
model could not be obtained in the field, some of them, such as the explosive properties,
were taken from the literature and the other, such as the rock factor A, was obtained from
back calculations of fragmented data for the first blast. Once this parameter was adjusted
for the first blast, the same values were used in the prediction calculations of fragment
size distribution for the remaining four blasts.

a. Case 1:


The full size field test was performed at a mine site of limestone quarry in West
Virginia, U. S. A. The following technical data were obtained. The compressive strength
of the blasted rock mass had a mean value of 12,100 psi. The density was 2.68 g/cc. The
bench height ranged from 54.54 ft to 56.70 ft. Borehole diameter was 6.25 inches.
Columns were filled with a Heavy ANFO type of explosives. The designed burden value
was 16 ft. However, the bench profiles data were collected for each hole for the first row
of boreholes to find out the exact values of burden. The crest burden varied from 14.81 ft

76
to 18.35 ft with a calculated mean value of 16.01 ft. The fragmentation was measured by
photographic (image analysis) method. The Split-Desktop software was used for the
calculation of actual size distribution. A detailed description of this softwares application
is shown in the Appendix C of this thesis.

Figure 5.1.1 The quarry highwall before actual shot (blast # 1).

The missing input information is tensile strength of rock mass, explosive strength,
energy, detonation velocity, Youngs modulus, and rock factor. The rock factor value had
been taken as 6.67, which is reflective of the geological conditions existing at the quarry
site. The pictures of the quarry highwall before actual shot (Fig. 5.1.1) and muckpile
(Fig.5.1.2) are presented below. The input data are predicted with the average values and
adjusted for the borehole size influence.


77




Figure 5.1.2 Muckpile immediately after the shot (blast # 1)

The following input data were used:
Detonation velocity = 4,350 m/sec,
Explosive strength = 96,
Explosive energy = 3.9 MJ/ m
3
,
Youngs modulus = 8.0 * 10
6
psi,
Rock factor = 6.67,
Crest burden = 16.01 ft,
Toe burden = 28.59 ft,

78
Bench height = 55.79 ft,
Stemming = 8 ft,
Number of rows = 3,
Number of holes in row = 13,
Loading density = 1,300 kg/ m
3
,
Powder factor = 1.1 lb/yd
3
,
Rock density = 2.68 g/cc.

The comparisons between the field data and the predicted data by the new model
FRAGM and the traditional Kuz-Ram model are listed in Table 5.1.
From the comparison (Fig.5.1.3), the prediction of FRAGM is in good agreement
with the field observations. It can be concluded that the application of FRAGM to
case 1 was successful.











79




Table 5.1 Predicted and actual size distribution for the blast # 1

Screen Size,
cm

Kuz-Ram Method,
%

FRAGM,
%

Field data,
%
5 88.68 90.34 92.34
10 72.93 76.57 78.57
15 57.38 62.52 64.52
20 43.62 49.59 52.59
25 32.23 38.39 40.39
30 23.23 29.11 31.11
35 16.37 21.65 24.98
40 11.31 15.84 20.65
45 7.67 11.40 14.32
50 5.11 8.09 10.54
55 3.35 5.66 8.07
60 2.16 3.91 5.78
65 1.37 2.66 4.07
70 0.86 1.80 2.75
75 0.53 1.20 1.48
95 0.07 0.21 0.34
100 0.04 0.13 0.06
120 0.00 0.02 0.00
124 0.00 0.01 0.00
130 0.00 0.01 0.00
140 0.00 0.00 0.00














80






























Figure 5.1.3 Cumulative Size Distribution Blast # 1.



Cumulative Size Distribution
0
10
20
30
40
50
60
70
80
90
100
0 25 50 75 100
Screen size, cm
P
e
r
c
e
n
t
a
g
e

o
f

t
h
e

m
a
t
e
r
i
a
l

r
e
t
a
i
n
e
d

o
n

t
h
e

s
c
r
e
e
n
Kuz-Ram,
%
FRAGM,
%
Field data,
%

81
b. Case 2:


The full size field test was performed at a mine site of limestone quarry in West
Virginia, U. S. A. The following technical data were obtained. The compressive strength
of the blasted rock mass had a mean value of 12,100 psi. The density was 2.68 g/cc. The
bench height ranged from 33.18 ft to 42.68 ft. Borehole diameter was 5.0 inches.
Columns were filled with a Heavy ANFO type of explosives. The designed burden value
was 12 ft. However, the bench profiles data were collected for each hole for the first row
of boreholes to find out the exact values of burden. The crest burden varied from 12.03 ft
to 13.08 ft with a calculated mean value of 12.55 ft. The fragmentation was measured by
photographic (image analysis) method. The Split-Desktop software was used for the
calculation of actual size distribution. The images of the quarry highwall before actual
shot (Fig. 5.2.1) and muckpile (Fig. 5.2.2) are presented below.
The missing input information is tensile strength of rock mass, explosive strength,
energy, detonation velocity, Youngs modulus, and rock factor. The rock factor value had
been taken as 6.67, which is reflective of the geological conditions existing at the quarry
site. The input data are predicted with the average values and adjusted for the borehole
size influence.







82







Figure 5.2.1 The quarry highwall before actual shot (blast # 2)







83








Figure 5.2.2 Muckpile a few days after the shot (blast # 2)





84
The following input data were used:
Detonation velocity = 4,350 m/sec,
Explosive strength = 96,
Explosive energy = 3.9 MJ/ m
3
,
Youngs modulus = 8.0 * 10
6
psi,
Rock factor = 6.67,
Crest burden = 12.55 ft,
Toe burden = 35.30 ft,
Bench height = 38.31 ft,
Stemming = 7 ft,
Number of rows = 6,
Number of holes in row = 13,
Loading density = 1,300 kg/ m
3
,
Powder factor = 1.9 lb/yd
3
,
Rock density = 2.68 g/cc.

The comparisons between the field data and the predicted data by the FRAGM and
the traditional Kuz-Ram model are listed in Table 5.2.
From the comparison (Fig. 5.2.3), the prediction of FRAGM is in good agreement
with the field observations. It can be concluded that the application of FRAGM to
case 2 was successful.




85







Table 5.2 Predicted and actual size distribution for the blast # 2


Screen Size,
cm

Kuz-Ram Method,
%

FRAGM,
%

Field data,
%
5 85.92 87.78 87.57
10 70.05 73.65 69.92
15 55.65 60.44 57.24
20 43.40 48.81 46.49
25 33.34 38.92 38.36
30 25.30 30.70 31.96
35 18.99 24.00 26.93
40 14.12 18.60 22.37
45 10.40 14.31 18.30
50 7.61 10.94 14.69
55 5.52 8.31 11.67
60 3.98 6.27 9.03
65 2.85 4.71 6.67
70 2.03 3.52 4.87
75 1.44 2.62 3.45
95 0.34 0.77 0.60
100 0.24 0.56 0.30
120 0.05 0.15 0.00
130 0.02 0.08 0.00
140 0.01 0.04 0.00











86



















Figure 5.2.3 Cumulative Size Distribution Blast # 2.





Cumulative Size Distribution
0
10
20
30
40
50
60
70
80
90
100
0 20 40 60 80 100
Screen Size,cm
P
e
r
c
e
n
t
a
g
e

R
e
t
a
i
n
e
d

o
n

t
h
e

S
c
r
e
e
n
Kuz-Ram,
%
FRAGM,
%
Field data, %

87
c. Case 3:


The full size field test was performed at a mine site of limestone quarry in West
Virginia, U. S. A. The following technical data were obtained. The compressive strength
of the blasted rock mass had a mean value of 12,100 psi. The density was 2.68 g/cc. The
bench height ranged from 33.18 ft to 42.68 ft. Borehole diameter was 5.5 inches.
Columns were filled with an ANFO type of explosives. The designed burden value was
12 ft. However, the bench profiles data were collected for each hole for the first row of
boreholes to find out the exact values of burden. The crest burden varied from 12.00 ft to
13.88 ft with an actual field mean value of 12.73 ft. The fragmentation was measured by
photographic (image analysis) method. The Split-Desktop software was used for the
calculation of actual size distribution. The image of the quarry highwall before actual
shot has been damaged. The image of the quarry highwall approximately one week after
the shot (Fig.5.3.1) and the picture of muckpile, that has been taken immediately after the
shot (Fig.5.3.2), are presented below.
The missing input information is tensile strength of rock mass, explosive strength,
energy, detonation velocity, Youngs modulus, and rock factor. The rock factor value had
been taken as 6.67, which is reflective of the geological conditions existing at the quarry
site. The input data are predicted with the average values and adjusted for the borehole
size influence.





88






Figure 5.3.1 Quarry highwall and muckpile approximately week after actual shot
(blast # 3)








89








Figure 5.3.2 Muckpile after the shot (blast # 3)





90
The following input data were used:
Detonation velocity = 4,500 m/sec,
Explosive strength = 100,
Explosive energy = 4.2 MJ/ m
3
,
Youngs modulus = 8.0 * 10
6
psi,
Rock factor = 6.67,
Crest burden = 12.73 ft,
Toe burden = 27.84 ft,
Bench height = 63.01 ft,
Stemming = 7 ft,
Number of rows = 2,
Number of holes in row = 11,
Loading density = 810 kg/ m
3
,
Powder factor = 1.4 lb/yd
3
,
Rock density = 2.68 g/cc.

The comparisons between the field data and the predicted data by the FRAGM and
the traditional Kuz-Ram model are listed in Table 5.3.
From the comparison (Fig. 5.3.3), the prediction of FRAGM is in good agreement
with the field observations. It can be concluded that the application of FRAGM to
case 3 was successful.




91









Table 5.3 Predicted and actual size distribution for the blast # 3

Screen Size,
cm

Kuz-Ram Method,
%

FRAGM,
%

Field data,
%
5 93.55 95.37 93.61
10 83.56 88.00 82.78
15 72.57 79.59 76.59
20 61.64 70.86 65.52
25 51.39 62.26 58.77
30 42.14 54.06 51.79
35 34.05 46.45 45.51
40 27.15 39.53 39.98
45 21.37 33.34 35.25
50 16.63 27.89 31.21
55 12.80 23.14 27.70
60 9.74 19.06 24.49
65 7.35 15.59 21.60
70 5.49 12.67 19.05
75 4.06 10.23 16.72
95 1.12 4.09 9.43
100 0.80 3.20 8.11
120 0.19 1.15 4.21
124 0.14 0.93 3.55
130 0.09 0.67 2.92
140 0.04 0.38 1.86









92




















Figure 5.3.3 Cumulative Size Distribution Blast # 3.



Cumulative Size Distribution
0
10
20
30
40
50
60
70
80
90
100
0 25 50 75 10
0
12
5
Screen Size, cm
P
e
r
c
e
n
t
a
g
e

o
f

t
h
e

m
a
t
e
r
i
a
l

r
e
t
a
i
n
e
d

o
n

t
h
e

s
c
r
e
e
n
Kuz-Ram,
%
FRAGM,
%
Field data,
%

93
d. Case 4:


The full size field test was performed at a mine site of limestone quarry in West
Virginia, U. S. A. The following technical data were obtained. The compressive strength
of the blasted rock mass had a mean value of 12,100 psi. The density was 2.68 g/cc. The
bench height ranged from 60.61 ft to 63.16 ft. Borehole diameter was 5.5 inches.
Columns were filled with a Heavy ANFO type of explosives. The designed burden value
was 13 ft. However, the bench profiles data were collected for each hole for the first row
of boreholes to find out the exact values of burden. The crest burden varied from 11.06 ft
to 15.68 ft with an actual field mean value of 13.08 ft. The fragmentation was measured
by photographic (image analysis) method. The Split-Desktop software was used for the
calculation of actual size distribution. The author is not able to present the image of
quarry highwall that had been taken before actual blast. The images of muckpile have
been taken immediately after the shot and one of them is presented below (Fig.5.4.1).
The missing input information is tensile strength of rock mass, explosive strength,
energy, detonation velocity, Youngs modulus, and rock factor. The rock factor value had
been taken as 6.67, which is reflective of the geological conditions existing at the quarry
site. The input data are predicted with the average values and adjusted for the borehole
size influence.






94






Figure 5.4.1 Muckpile immediately after the shot (blast # 4)








95
The following input data were used:
Detonation velocity = 4,350 m/sec,
Explosive strength = 96,
Explosive energy = 3.9 MJ/ m
3
,
Youngs modulus = 8.0 * 10
6
psi,
Rock factor = 6.67,
Crest burden = 13.08 ft,
Toe burden = 23.14 ft,
Bench height = 61.78 ft,
Stemming = 7 ft,
Number of rows = 2,
Number of holes in row = 14,
Loading density = 1,150 kg/ m
3
,
Powder factor = 1.71 lb/yd
3
,
Rock density = 2.68 g/cc.

The comparisons between the field data and the predicted data by the FRAGM and
the traditional Kuz-Ram model are listed in Table 5.4.
From the comparison (Fig.5.4.2), the prediction of FRAGM is in good agreement
with the field observations. It can be concluded that the application of FRAGM to
case 4 was successful.




96









Table 5.4 Predicted and actual size distribution for the blast # 4

Screen Size,
cm

Kuz-Ram Method,
%

FRAGM,
%

Field data,
%
5 89.04 91.31 93.31
10 74.08 79.06 81.07
15 59.29 66.41 68.41
20 46.07 54.50 56.49
25 34.92 43.87 45.86
30 25.91 34.72 38.72
35 18.86 27.08 32.92
40 13.50 20.83 25.83
45 9.51 15.83 20.83
50 6.60 11.89 17.89
55 4.52 8.84 14.83
60 3.05 6.50 12.49
65 2.04 4.74 8.60
70 1.34 3.42 7.18
75 0.88 2.45 3.45
95 0.14 0.59 0.00
100 0.09 0.41 0.00
120 0.01 0.09 0.00
124 0.01 0.06 0.00
130 0.00 0.04 0.00
140 0.00 0.02 0.00









97














































Figure 5.4.2 Cumulative Size Distribution Blast # 4.








Cumulative Size Distribution
0
10
20
30
40
50
60
70
80
90
100
0 25 50 75 100
Screen Size, cm
P
e
r
c
e
n
t
a
g
e

o
f

t
h
e

m
a
t
e
r
i
a
l

r
e
t
a
i
n
e
d

o
n

t
h
e

s
c
r
e
e
n
Kuz-
Ram,
%
FRAGM,
%
Field
data,
%

98
e. Case 5:


The full size field test was performed at a mine site of limestone quarry in West
Virginia, U. S. A. The following technical data were obtained. The compressive strength
of the blasted rock mass had a mean value of 12,100 psi. The density was 2.68 g/cc. The
bench height ranged from 60.61 ft to 63.16 ft. Borehole diameter was 6.125 inch.
Columns were filled with a Heavy ANFO type of explosives. The designed burden value
was 14 ft. However, the bench profiles data were collected for each hole for the first row
of boreholes to find out the exact values of burden. The crest burden varied from 11.06 ft
to 15.68 ft with an actual field mean value of 14.06 ft. The fragmentation was measured
by photographic (image analysis) method. The Split-Desktop software was used for the
calculation of actual size distribution. The images of the quarry highwall before actual
shot (Fig.5.5.1) and muckpile after the shot (Fig. 5.5.2) are presented below.

The missing input information is tensile strength of rock mass, explosive strength,
energy, detonation velocity, Youngs modulus and rock factor. The rock factor value had
been taken as 6.67, which is reflective of the geological conditions existing at the quarry
site. The input data are predicted with the average values and adjusted for the borehole
size influence.






99







Figure 5.5.1 The quarry highwall before actual shot (blast # 5)







100








Figure 5.5.2 Muckpile immediately after the shot (blast # 5)






101
The following input data were used:
Detonation velocity = 4,350 m/sec,
Explosive strength = 96,
Explosive energy = 3.9 MJ/ m
3
,
Youngs modulus = 8.0 * 10
6
psi,
Rock factor = 6.67,
Crest burden = 14.06 ft,
Toe burden = 23.30 ft,
Bench height = 62.51 ft,
Stemming = 7 ft,
Number of rows = 2,
Number of holes in row = 11,
Loading density = 1,150 kg/ m
3
,
Powder factor = 1.19 lb/yd
3
,
Rock density = 2.68 g/cc.

The comparisons between the field data and the predicted data by FRAGM and the
traditional Kuz-Ram model are listed in Table 5.5.
From the comparison (Fig. 5.5.3), the prediction of FRAGM is in good agreement
with the field observations. It can be concluded that the application of FRAGM to
case 5 was successful.




102











Table 5.5 Predicted and actual size distribution for the blast # 5


Screen Size,
cm

Kuz-Ram Method,
%

FRAGM,
%

Field data,
%
5 92.21 93.66 96.66
10 79.84 83.38 87.38
15 66.43 71.87 76.87
20 53.54 60.38 64.38
25 41.99 49.62 51.62
30 32.14 39.99 43.67
35 24.07 31.66 35.66
40 17.66 24.66 29.75
45 12.72 18.91 23.94
50 8.90 14.30 17.30
55 6.26 10.67 15.87
60 4.29 7.86 12.76
65 2.89 5.72 10.34
70 1.92 4.11 6.56
75 1.26 2.92 4.72
95 0.20 0.67 2.25
100 0.13 0.45 0.68
120 0.02 0.09 0.31
124 0.01 0.06 0.28
130 0.01 0.04 0.20
140 0.00 0.01 0.11







103









































Figure 5.5.3 Cumulative Size Distribution Blast # 5.





Cumulative Size Distribution
0
10
20
30
40
50
60
70
80
90
100
110
0 25 50 75 100
Screen Size, cm
P
e
r
c
e
n
t
a
g
e

o
f

t
h
e

m
a
t
e
r
i
a
l

r
e
t
a
i
n
e
d

o
n

t
h
e

s
c
r
e
e
n
Kuz-Ram,
%
FRAGM,
%
Field data,
%

104
Chapter 6 Quality Control


6.1 Introduction
Regardless of the care taken in setting up optimum blast patterns unless there is
sufficient attention given to the quality of the explosives being loaded the produced
fragmentation will be poor. Thus a regular program for analyzing the quality of the
explosive products should be adopted. Some of the more important characteristics of the
explosives that should be watched out for are outlined in the following sections.

6.2 Ammonium Nitrate Based Dry Explosives
For an ANFO explosives it is critical to determine the fuel oil percentage being
added to the prilled ammonium nitrate. For an ANFO mixture the reaction equation is

3NH
4
NO
3
+ CH
2
3N
2
+ 7H
2
O + CO
2
(6.2.1)

for the optimal concentration of 5.6 % fuel oil in the ANFO mixture. Failure to obtain
this optimal value decreases the energy output of the ANFO. Since the energy loss is
greater for a decrease in the fuel oil content than for a corresponding increase it is better
for operators to steer towards a slightly fuel oil rich value of around 6.0 % rather than the
optimal 5.6 %.
In addition to the basic ANFO mixes some mixtures use aluminum additions to boost
their strength. It is known that the energy output relative to ANFO increases as an
increasing amount of aluminum is added. Bauer (1972) pointed out that in certain point

105
the increase in energy per percentage increase in aluminum would start to decrease, so 13
to 15 % should generally be chosen as the practical limit of aluminum addition.

6.3 Slurries
As with ANFO explosives it is essential that individual components of the slurry be
of the correct concentration. However, by virtue of the increasing number of constituents
making up the slurry as compared to an ANFO the number of constraints is greater. Two
factors that influence the energy output for the specific TNT slurry are the water content
and the aluminum content. The loss of energy is associated with higher water content in
the slurry (Bauer, 1972). Hence the question becomes one of minimizing this value
between practical limits. However, just as with dry mixes there is an optimal value
beyond which the percentage increase in energy per percentage increase in aluminum
decreases and the economics of adding this relatively expensive energizer preclude
further additions.
Other factors, which can also influence the performance of the slurry, particularly
with respect to the critical diameter of the explosive, include pressure, temperature,
density and TNT content.









106
Chapter 7 Conclusions and Recommendations


The analysis of bench blasting fragmentation and the critical evaluation of the events
that make-up the fragmentation process of the rock mass under the action of explosives
resulted in the development of a simulation model, FRAGM, for predicting fragmentation
curve from a given blast.
The engineering model, called FRAGM, has the ability to predict the size
distribution for a given blast before actual shot and takes into account irregularity of the
bench face. The fragmentation model uses the hole-by-hole analysis method for the first
row of blastholes. There is no method which can solve the effect of bench face
irregularities on rock blasting fragmentation without analyzing it hole by hole. The ability
of FRAGM model to fit known distribution was tested on several case studies. In every
case, the comparisons were favorable. Powder factor dependence of fragmentation results
was also well described by the model. It is therefore concluded that after careful
comparison the new engineering model has been successful in its applications to field
data.
The model is designed to provide means for testing variations in blasting patterns
without the need to conduct an extensive series of test blasts, and has the following
characteristics:
1. It runs on a personal computer which is usually available in an engineers
office;
2. It is fast, not time consuming;
3. It is simple.

107
Analysis of known fragmentation results showed that FRAGM model is a promising
tool for predicting fragmentation distribution from blasting, provided the correct values
for the parameters incorporated in the model are used.
The model can be used to:
1. Predict the effect of blasthole size, explosives selection, and many other
parameters on the resulting fragmentation distribution;
2. Predict the effect of a bench face irregularities on the fragmentation distribution;
3. Design a blasting technique to provide a specified results, for example, a
reduced fragmentation size range;
4. Determine the effect of discontinuities in the rock mass on blast fragmentation
distribution.
The main advantage of using this model as compared to other models available in the
literature is that the effect of bench face irregularities in rock fragmentation can be
studied for improving the efficiency of the blasting process. Furthermore, it is
recommended that additional research in the following areas may establish a more
comprehensive model for predicting fragmentation distribution.
FRAGM is certainly a restricted model. Modern blasting techniques make use of
variable loading density of explosive along the blasthole for better usage of explosives. In
these cases a three dimensional simulation of the region around the blasthole is required.
Under certain favorable geologic conditions (relatively short distance of muckpiles
from the faces, and fairly high efficiency of explosive excavation) the cheapest method of
moving the overburden is the blast casting method. In this mining method the degree of
fragmentation achieved is an important economic factor. If the degree of fragmentation

108
obtained from blast casting is to be predicted, a special crack analysis model, which takes
into account further enhancement of fragmentation due to collision of fragments and their
falling on a hard surface, must be included.
A high-speed camera should be used to monitor the full scale blasts in order to study
the mechanisms of crack initiation, propagation, and interaction in bench blasting. A
high-speed camera can also be used to determine the throwing direction and the velocity
distribution in each region. It will help to relate the theoretical burden velocity with the
velocity distribution inside the burden, and the throwing direction of the broken mass to
determine the muckpile profile.
A series of full-scale blast tests should be conducted to study the effect of the density
of discontinuities on the mean fragment size and uniformity of fragmentation.
Furthermore, field oriented studies may be taken up with high-speed photography
techniques for a better understanding of the effect of discontinuities on rock
fragmentation by blasting. By this way the understanding of fragmentation can be
improved in the near future, for example effective selection of the face orientation,
burden, spacing, etc. required with respect to weakness planes present in the rock mass.
Since the size distribution is the major determining factor in loading and hauling
operations, the predicted average fragment size can be used to combine the entire
operation together. The improved computer program will provide more help to blast
designers and mine operators. Also, it can be used for cost analysis of changing the
blasting pattern or explosive type.



109
References




Adushkin, V. V., Kryndushkin, S. K., and Tikhomirov, A. M., Compressional Wave
Behaviour with Explosion of a Cylindrical Charge in a Solid Medium,
Soviet Mining Science, Plenum Publishing Corporation, 1987,
pp. 253 261.

Anonymous, Open Cut Workings, Proceedings of the AIME (in
Russian), Nedra, Moskow, 1971, pp. 128 135.

Ash, R.L, The Influence of Geological Discontinuities on Rock Blasting, Ph.D.
Dissertation, University of Minnesota, Minneapolis, 1973, 289p.


Atchison, T. C., and Duvall, W. I., Mobile Laboratory for Recording Blasting and Other
Transient Phenomena, USBM, RI 5197, 1956, 22p.


Atchison, T. C., and Duvall, W. I., Rock Breakage by Explosives, USBM, RI
5356, 1957, 28p.

Azarkovich, A. E., The Radius of Fracture of an Elongated Charge, Vzryvnoe
Delo (in Russian), Volume 14, No. 57, Nedra, Moskow, 1965,
11pp.

Azarkovich, A. E., Influence of Natural Jointing of Ledge Rock on the Radius of Crack
Formation During an Explosion, Soviet Mining Science,
Volume 17, No. 1, Moskow, 1981, pp. 29 38.


110
Bauer, A., Current Drilling and Blasting Practices in Open Pit Mining, Mining
Congress Journal, March 1972, pp. 20- 26.

Bhandari, S., Burden and Spacing Relationship in the Design of Blasting Pattern,
Proceedings, 16th Symposium on Rock
Mechanics, University of Minnesota, 1975a, pp. 333 343.

Bhandari, S., Improved Fragmentation by Reduced Burden and More Spacing in
Blasting, Mining Magazine, 1975b, pp. 183 195.

Bhandari, S., Changes in Fragmentation Process with Blasting Conditions,
Proceedings, 5th Symposium on Rock
Fragmentation by Blasting, Montreal, 1996, pp. 301
309.

Bhandari, S., Engineering Rock Blasting Operations, A. A. Balkema, Rotterdam, 1997,
pp. 191 195.

Bjarnholt, G., and Skalare, H., Instrumental Model Scale Blasting in Concrete,
Proceedings, First International Symposium
on Rock Fragmentation by Blasting, Lulea, Sweden,
August 1983, pp. 799 814.

Blair, B. E., Use of HighSpeed Camera in Blasting Studies, USBM, RI 5584,
1960, 32p.

Cook, M. A., The Science of High Explosives, Robert E. Rieger Publishing Co., New
York, 1971, pp. 426.


111
Cook, M. A., The Science of Industrial Explosives, Graphic Services and Supply
Incorporation, New York, 1976, pp. 394.

Cunningham, C. V. B., The KuzRam Model for Prediction of Fragmentation from
Blasting, Proceedings, First International
Symposium on Rock Fragmentation by
Blasting, Lulea, Sweden, August 1983, pp. 439 453.

Cunningham, C. V. B., Fragmentation Estimations and the KuzRam ModelFour
Years on, Second International Symposium on
Rock Fragmentation by Blasting, Keystone, Colorado,
U.S.A., August 1987, pp. 475 487.

Da Gama, D., Use of Communution Theory to Predict Fragmentation of Jointed Rock
Masses Subjected to Blasting, Proceedings, First
International Symposium on Rock
Fragmentation by Blasting, Lulea, Sweden, August 1983,
pp. 565 579.

Drukovanyi, M. F., Komir, V. M., Myachina, N. I., Rodak, S. N., Semenyuk, E. A.,
Effect of the Charge Diameter and Type of Explosive on the Size
Overcrushing Zone During Explosion, Soviet Mining

112
Science (in Russian), Volume 9, No. 5, Moskow, 1973, pp.
500 506.

Drukovanyi, M. F., Kravtsov, V. S., Chernyavskyi, Y. E., Shelenok, V. V., Reva, N.P.,
Zverkov, S. N., Calculation of Fracture Zones Created by Exploding
Cylindrical Charges in Ledge Rocks, Soviet Mining Science
(in Russian), Volume 12, No.3, Moskow, 1976, pp. 292 295.

Fadeenkov, N. M., Applicability of the RosinRammler Law to the Analysis of the
GrainSize Composition of a Heap of Blasted Rock, Soviet
Mining Science, Volume 10, No.2, Moskow, 1975, pp. 685 688.

Favreau, R. F., Rock Displacement Velocity During a Bench Blasting,
Proceedings, First International Symposium
on Rock Fragmentation by Blasting, Lulea, Sweden,
August 1983, pp. 753 776.

Favreau, R. F., The Prediction of BlastInduced Swell by Means of Computer
Simulations, CIM Bulletin, 86 (967), 1993, pp. 69 72.

Fourney, L. W., Barker, B. D., and Holloway, C. D., Fragmentation in Jointed Rock
Material, Proceedings, First International

113
Symposium on Rock Fragmentation by
Blasting, Lulea, Sweden, August 1983, pp. 505 531.

Goldsmith, W., Pulse Propagation in Rocks, Proceedings, 8th
Symposium on Rock Mechanics, New York, 1967,
pp.528 537.

Gy, P.M., Sampling of Particulate Materials: Theory and Practice, Elsevier,
Amsterdam, 1979, 243p.

Harries, G., and Hengst, B., The Use of Computer to Describe Blasting,
Proceedings, 15th APCOM Symposium, Brisbane,
Austalia, 1977, pp. 317 324.

Harries, G., A Mathematical Model of Cratering and Blasting, Proceedings,
National Symposium on Rock Fragmentation,
Adelaide, 1983, pp. 41 54.

Higgings, M., BoBo, T., Girder, K., Kemeny, J., and Seppala, V., Integrated Software
Tools and Methodology for Optimization of Blast Fragmentation,
International Society of Explosives Engineers
Annual Conference, 1999, 7 p.
Hjelmberg, H., Some Ideas on How to Improve Calculations of the Size Fragment Size
Distribution in Bench Blasting, Proceedings, First

114
Symposium on Rock Fragmentation By
Blasting, Lulea, Sweden, August 1990, pp. 469 494.

Hustrulid, W., and Kuchta, M., Open Pit Mine Planning and Design, Volume 1
Fundamentals, Balkema, Rotterdam, 1998, pp. 254 255.

Hustrulid, W., Blasting Principles for Open Pit Mining: Theoretical Foundations,
Volume 2, Balkema, Rotterdam, 1999, pp. 980-992.

Jimeno, C. L., Jimeno, E. L., Carcero, F. J. A., Drilling and Blasting of Rock,
Geomining Technological Institute of Spain
under Contract with the E. P. M., S. A.
Company, Balkema, Rotterdam, 1995, p. 385.
Konya. C. J., Walter, E. J., Rock Blasting, Seminar on Blasting and
Overbreak Control, Montville, Ohio, May 1985, pp.36-42.

Kuznetsov, V. M., The Mean Diameter of Fragments Formed by Blasting Rock,
Soviet Mining Science (in Russian), Vol. 9,
Moskow, 1973, pp. 144 148.

Langefors, U., and Kihlstrom, B., The Modern Technique of Rock Blasting, Wiley,
New York, 1963, pp. 405 411.


115
Lilly, P. A., An Empirical Method of Assessing Rock Mass Blastability,
Proceedings, Large Open Pit Mining
Conference (J. R. Davidson, ed.), AIMM, Parkville,
Victoria, Canada, October 1986, pp. 89 92.

Lownds, C. M., and Seligmann, C. P., Primary Fracture from an Array of Shotholes,
Journal of the South African Institute of
Mining and Metallurgy, Volume 76, 1976, pp. 307 310.

Lownds, C. M., Computer Modelling of Fragmentation from an Array of Shotholes,
Proceedings, First International Symposium
on Rock Fragmentation By Blasting, Lulea, Sweden,
August 1983, pp. 407 418.

Mendenhall, W., Beaver, R. J., and Beaver, B. M., Introduction to Probability and
Statistics, Eleventh Edition, Brooks/Cole-Thomson Learning, Pacific
Grove, California, U. S. A., 2002, pp. 63 67.

MacKenzie, A. S., Cost of Explosives- Do You Evaluate It Properly?, Mining
Congress Journal, Volume 52, No. 5, 1966, pp. 32 41.

116
MacKenzie, A. S., Optimum Blasting, Proceedings of the 28
th

Annual Minnesota Mining Symposium, Duluth,
1967, pp. 181 188.

Maerz, N. H., Franklin, J. A., Rothenburg, L., and Coursen, D. L., Measurement of
Rock Fragmentation by Digital Photoanalysis, Proceedings,
Fifth International Congress International
Society of Rock Mechanics, 1986, pp. 687- 682.

Margolin, L. G., and Adams, T. F., Numerical Simulation of Fracture,
Proceedings, First International Symposium
on Rock Fragmentation by Blasting, Lulea, Sweden,
August 1983, pp. 418 425.

McDermott, C., Hunter, G. L., and Miles, N. J., The Application of Image Analysis to
the Measurement of Blast Fragmentation, Proceedings,
Symposium Mining-Future Concepts, Nottingham
University, Marylebone Press, Manchester, 1989, pp. 103- 108.

McHugh, S., Computational Simulations of Dynamically Induced Fracture and
Fragmentation, Proceedings, First International
Symposium on Rock Fragmentation by
Blasting, Lulea, Sweden, August 1983, pp. 407 418.

117
Mohanty, B., Strength of Rock under Strain Rate Loading Conditions Applicable to
Blasting, Proceedings, Second Symposium on
Rock Fragmentation by Blasting, Keystone, Colorado,
U. S .A., 1987, pp. 72 79.

Nie, S. L., and Rustan, A., Technique and Procedures in Analysing Fragmentation After
Blasting by Photographic Method, Proceedings, Second
International Symposium on Rock
Fragmentation by Blasting, Keystone, Colorado, U. S. A.,
1987, pp. 102- 113.

Obert, L., Duvall, W. I., Generation and Agitation of Strain Waves in Rock Part I,
USBM,RI 4583, 1950, 31 p.

Olofsson, S. O., Explosives Technology for Construction and Mining, Second Edition,
Applex, Arla, Sweden, 1990, pp. 106 107.

Petkof, B., Atchison, T.C., Duvall, W. I., Photographic Observation of Quarry
Blasting, USBM, RI 5849, 1961, 14 p.

Porter, D. D., Use of Rock Fragmentation to Evaluate Explosive in Blasting,
Mining Congress Journal, Volume 60, No. 1, 1974, pp. 41
43.

118

Preece, D. S., and Khudsen, S. D., Blasting Induced Rock Motion Modelling Including
Gas Pressure Effect, 24
th
U.S. Oil Shale Symposium
(J. H. Bary, ed.), Quarterly Colorado School Of
Mines, Volume 83, No. 4, Golden, Colorado, U. S. A., 1991, pp. 13-
19.

Preece, D.S., Tidman, J. P., Chung, S. H., Expanded Rock Blasting Modelling
Capabilities Of DMC_BLAST, Including Buffer Blasting,
Proceedings, 13
th
Annual Symposium on
Explosives and Blasting Research, ISEE, U.S.A.,
1997, pp.125 134.

Preece, D. S., Burchell, S. L., Scovira, D. S., Coupled Explosive Gas Flow and Rock
Motion Modelling with Comparison to Bench Blast Field Data,
Proceedings of the 4
th
International
Symposium on Rock Fragmentation by
Blasting, Fragblast-4 ( H .P Rossmanith, ed.),
Vienna, Austria, July 1993, pp. 239 245.

Roberts, D. K., Wells, A. A., Velocity of Brittle Fracture, Engineering, Volume
17, No.8, December 1954, pp.220-224.

Shapurin, A. V., and Kutuzov, B., Blast Round Design for Open Pit Mines", Manual
for the Explosives and Mining Engineering

119
Students, Department of Explosives Engineering, Moskow State
Mining Institute, Moskow, 1990, p.55.
Split Engineering LLC, Split-Desktop Software Manual, Tuscon, Arizona, U. S. A.,
2001, p.145.

Vovk, A. A., Mikhalyuk, A. V., Belinskii, I. V., Development of Fracture Zones in Rock
During Camouflet Blasting, Soviet Mining Science (in
Russian), Volume 9, No. 4, Moskow, 1973, pp. 383 387.

Worsey, P., Rustan, A., Line, N. S., New Method to Test the Rock Breaking Properties
of Explosives in Full Scale, Second Symposium on Rock
Fragmentation by Blasting, Keystone, Colorado, U. S. A.,
1987, pp. 485 487.













120
Appendix A Verification of the Proposed Methodology by
Hand Calculations


The calculations involved in the estimation of fragment size distribution, described
in Chapter 3, are done by hand here for the case # 1 to illustrate the details of the
proposed algorithm. This exercise will also demonstrate the necessity of a computer to
deal with the laborious calculations involved in the process.

Case # 1

Hand verification of the proposed methodology was done using the field data
obtained from a limestone quarry in West Virginia, U. S. A. The following technical data
required for the estimation of fragment size distribution were collected at the site. The
compressive strength of the rock mass had a mean value of 12,100 psi. The density was
2.68 g/cc. The bench height ranged from 54.54 ft to 56.70 ft. Borehole diameter was 6.25
inches. Columns were filled with a Heavy ANFO type of explosives. The designed
burden value was 16 ft.
However, the bench profiles data were collected for each hole for the first row of
boreholes to find out the exact values of burden. This data is presented in Table B.1.1 and
Table B.1.1.1 in Appendix B of this thesis. The crest burden varied from 14.81 ft to 18.35
ft with a calculated mean value of 16.01 ft. The fragmentation was measured by
photographic (image analysis) method. Details of the Split-Desktop software used for
the image analysis and determination of the actual fragment size distribution in the field
are provided in Appendix C of this thesis.

121
The missing input information is the tensile strength of rock mass, explosive
strength, energy, detonation velocity, Youngs modulus, and rock factor. However, for
the present calculations, rock factor value had been taken as 6.67, which is reflective of
the geological conditions existing at the quarry site.
The following input data were used:
Detonation velocity = 4,350 m/sec,
Explosive strength = 96,
Explosive energy = 3.9 MJ/ m
3
,
Youngs modulus = 8.0 * 10
6
psi,
Rock factor = 6.67,
Crest burden = 16.01 ft,
Toe burden = 28.59 ft,
Bench height = 55.79 ft,
Stemming = 8 ft,
Number of rows = 3,
Number of holes in row = 13,
Loading density = 1,300 kg/ m
3
,
Powder factor = 1.1 lb/yd
3
,
Rock density = 2.68 g/cc.

Given below is a detailed and step-by-step description of different aspects of the
estimation of blast fragment size distribution using methodology developed in Chapter 3
of this thesis.

122
Step # 1

One of the most important parameters for the execution of efficient blasting is the
true burden. Measuring the true burden has, until recently, been a difficult operation.
However, as discussed in Chapter 3, technique like laser profiling could now be used to
estimate the true burden. As a matter of fact, the laser profiling technique has been used
for the estimation of the true burden in the field for the verification of the proposed rock
blasting fragmentation model. More details of the technique of the bench face profiling
have been presented in Appendix B of this thesis.
Once the values of true burden are obtained for each borehole in the first row,
trapezoidal area method is used for estimation of the rock volume in this row. The
trapezoidal area method has been chosen for its simplicity and fairly accurate results of
estimation. For illustration purpose, Figure A-1 presents a bench face profile for a given
borehole in the first line of boreholes. Lines BC and AD represent the axis of borehole
and bench face, respectively. The perpendiculars beginning at the axis of borehole and
ending at the bench face are the true burdens for a given borehole. Bench face
configuration AD was obtained from bench face profiling. One can see that any segment
lying between two perpendiculars has an approximate surface area defined by the rule
for determining the area of a trapezoid. For a given segment the area can be defined by:
A
i
=
i
i i
H
B B
|
.
|

\
| +
+
2
1
, sq. ft (sq. m) (A.1)
where : B
i
, B
i+1
upper and lower perpendiculars (burdens) in a given segment, ft (m)
respectively;
H
i
the height of a given segment, ft (m).


123
























Figure A-1 Typical bench face profile of borehole in the first row




A B1 B
D C
B2 H1
B3 H2
B4 H3
B5 H4
Bi-1 Hi-1
Bi Hi
B6 H6
B7 H7

124

Total surface area of a profile for a given hole in the first row can be estimated by:
A =

=
n
i
i
A
1
, sq. ft. (sq. m.) (A.2)

where: A
i
the segment area, sq. ft. (sq. m.);
n- number of segments in a given profile.

For hole # 1 of the case study, the total profile surface area is calculated and
presented below as an example. The true burden data has been taken from Table B.1.1 in
the Appendix B of this thesis.
( ) ( )
( ) ( )
( ) ( )
( ) ( )
.) . ( 3 . 1157
2961 . 1157 94 . 55 13 . 56
2
34 . 31 32 . 31
17 . 47 94 . 55
*
2
32 . 31 29 . 30
44 . 39 17 . 47
2
29 . 30 71 . 20
29 . 31 44 . 39
*
2
71 . 20 63 . 18
97 . 22 29 . 31
2
63 . 18 66 . 17
6 . 9 97 . 22
*
2
66 . 17 48 . 15
01 . 0 6 . 9
2
48 . 15 03 . 16
0 01 . 0
2
03 . 16 03 . 16
ft sq
=
|
.
|

\
| +
+
|
.
|

\
| +
+
|
.
|

\
| +
+
|
.
|

\
| +
+
|
.
|

\
| +
+
|
.
|

\
| +
+
|
.
|

\
| +
+
|
.
|

\
| +


The results of estimation of the total profile surface area for each hole in the first row
for the case # 1 are presented in Appendix B (Table B.1.1 and Table B.1.1.1).
After the total surface area for the first row of boreholes is estimated, the total
volume of rock in this row of boreholes can be found by using the following equation:


125

1
1
1
1
2
+
=
+

|
.
|

\
| +
=
i holei
n
i
holei holei
S
A A
V , cu. ft. (cu.m.) (A.3)

where A
holei
, A
holei+1
- the total face profile surface areas for the two adjacent
boreholes in the first row, sg. ft. (sq. m.);
S
holei-i+1
spacing between holes, ft (m).
n- number of boreholes in the first row.

The results of the estimation of rock volume in the first line for the case # 1 are
presented in the Appendix B (Table B.1.1.1).

Step # 2
For simplicity in establishing the boundaries of damage zones in the first row of
blastholes, an average profile of a bench face is determined. This is done by defining the
average bench height, average crest, and toe burdens. However, estimation of these
parameters should be done in relation to the total rock volume in the first row of
blastholes estimated in step # 1. Only then the powder factor taken in further calculations
will have the same numerical value as that obtained using the true volume estimated in
step # 1.
The average crest burden can be calculated as follows:
B
crest average =

=
n
i
i crest
n
B
1
,
ft (m) (A.4)
where: B
crest i
- the crest burden for a given borehole in the first row, ft (m);
n- the number of boreholes in the first row.

126
For the case #1, the average crest burden can be estimated by taking individual crest
burdens from Table B.1.1 and plugging them into equation (A.4):

B crest = (16.03 + 15.57 + 15.78 + 15.61 + 14.81 + 15.9 + 15.72 + 16.8 + 15.77 +
+ 18.35 +18.35 + 15.91 + 15.94)/13 = 16.01ft = 4.88 (m.)

It must, however, be pointed out that the method of establishing an average bench profile
for a given blast should not be used if the bench has a highly irregular shape in the plan
view. When this is the case, the individual crest burden values for different holes in the
first row will be highly variable. As a rule of thumb (Mendenhall et al., 2002), it is
suggested that if the arithmetic difference between the individual crest burden exceeds by
three standard deviations, then those areas must be treated separately. Also those crest
burdens should not be used for the estimation of the average crest burden for the bench.

Step # 3
The results of the step 1 and step 2 provide information required to estimate the
average toe burden. The average toe burden can be calculated as follows:
B
toe average
=
average crest
average
B
L H
V

1
2
, ft (m) (A.5)
where: V
1
- the total rock volume in the first row of blasholes, cu. ft. (cu. m.);
H
average
- the average bench height, ft (m);
B
crest average
the average crest burden, ft (m);
L- the total length of a block being blasted, ft (m).


127
For the case # 1, these values are taken from Table B.1.1.1 and the average toe
burden is found:
B
toe average
= 59 . 28 01 . 16
208 * 79 . 55
40 . 258774 * 2
= (ft.) = 8.72 (m.)
So the average bench face profile for the case # 1 has the following parameters:
Average bench height = 55.79 ft = 17.01 m;
Average toe burden = 28.59 ft = 8.72 m;
Average crest burden = 16.01 ft = 4.88 m.

Step # 4
After the average bench parameters are determined, the next step is to divide the total
rock volume among different holes in the first row involved in the blast. The boundary of
these areas will be the face of a bench and the axis of the borehole rows. For instance, if
there are three rows of blastholes, three areas should be established. They are: a) the area
of the first row of blastholes: b) the area of the second row of blastholes: c) the area of
the third row of blastholes.

Step # 5
The rock factor A must be calculated by using empirical equation (2.5.1.6). For case
# 1, the rock factor has been taken as 6.67, which is reflective of the geological
conditions existing at the quarry site. This factor may also be found for a given part of a
quarry by back analysis of fragmentation data.



128
Step # 6
The next step is to calculate the location of breaking points on each side of a
borehole. The purpose of this step is to define the main breakage and secondary breakage
areas, and also to find out the area of overlap between main breakage areas. As can be
seen from Figure 3.5, if there is no overlap of main breakage areas between two adjacent
boreholes, some part of the rock between them will remain intact. By defining the
boundaries of breakage areas and, hence, the overlap, the blast designer can predict
whether or not a zone of unbroken rock exists and make necessary adjustments to the
blast design parameters to avoid this situation. In order to determine the breakage areas,
the following procedure is used:
1. Find the linear density of the explosive by using the following formula:
q
e
=
10 * 4
2
d
, kg/m (A.6)

where: - the explosive density, g/cc;
d- the diameter of borehole, cm.
For the case # 1:
q
e
=
10 * 4
875 . 15 * 1419526 . 3 * 68 . 2
2
= 25.73 kg/m
2. Find the location of the breaking points:
=
( )
2 . 1
3
2
exp
10 * 62 . 6 *
*

|
|
.
|

\
|
e
q
R
c
E

(A.7)
where: E
exp
- the concentration of the explosive energy in charge, J/m
3
;
- the material density, kg/m
3
;

129
c- the elastic longitudinal wave velocity in the material, m/sec;
R-the distance between breaking points, m;
q
e
- the linear density of the explosive, kg/m;
- the strain of blast wave.
For the case #1:
E
exp
= 3.9*10
-9
J/m
3
; = 2680 kg/m
3
; c = 4350 m/sec; q
e
= 25.73 kg/m,
Youngs modulus (E) = 8 * 10
6
psi = 5.6241*10
9
kg/m
2

=
2 . 1
3
2 3
9
73 . 25
10 * 62 . 6 *
4350 * 10 * 68 . 2
10 * 9 . 3

|
|
.
|

\
| R
;
= 5.0911* 10
-4
*
2 . 1
73 . 25

|
|
.
|

\
| R


The relationship between stress ( ) and strain ( ) is defined as:
= * E (A.8)
where: , , E- the stress, strain and Youngs modulus, respectively.
Equation (B.8) can be written as:
=
( )
2 . 1
3
2
exp
10 * 62 . 6 *
*

|
|
.
|

\
|
qe
R
c
E

*E (A.9)
where E- Youngs modulus, kg/m
2
;
- the stress, kg/m
2
.

For the case # 1: = 5.0911* 10
-4
*
2 . 1
73 . 25

|
|
.
|

\
| R
* 5.6241*10
9
;

130
R = 73 . 25 *
10 * 35973 . 863 , 2
51 . 271 , 301 , 1
12 / 10
3

|
.
|

\
|
= 9.9793 9.98 m
To estimate the distance between breaking points, R, from equation (A.9), the value
of the critical stress, , required to cause breakage in the rock mass is required. A major
portion of the rock, in a typical blast, is broken when the tensile strain ( ) of the
reflected shock wave exceeds the strain (
0
) corresponding to the tensile strength of the
rock. Cook (1971) pointed out that when the average strain ) ( of the reflected wave
exceeds the limiting tensile strain ( )
o
by a factor K, then rock breakage occurs. That is,
when

o
K * = (A.10)
Cook (1971) also found that the values of K generally lies between 1.4 and 1.6. For
the case # 1, the value of K has been taken as 1.53. From Equations (A.10) and (A.8), the
critical stress for the case # 1 was estimated as 1,851 psi. This value was used for the
estimation of distance between breakage points, R, for each hole in the first row by using
equation (A.9).

Step # 7
In this step the blasted volume for each hole in the first row is calculated using the
boundary locations determined in step 4 and step 6. Connecting the breaking points on
each side with borehole forms the triangle of the main breakage area (Fig. 3.5). Using this
procedure, the main breakage area can be assigned for each hole in the first line. From the
data (case # 1) presented in Table A-1 it is obvious that there are overlaps of the main

131
breakage areas for each hole in the first row. It indicates that every point of the rock mass
within the breakage areas will be affected by blasting.
For simplicity in the following calculations, it is assumed that the boundaries of
damage zones are: 1) the perpendiculars (beginning with the axis of borehole in the first
row of holes and ending with the bench face) which include the point of intersection of
the main breakage areas; 2) the axis of boreholes in the first row; and 3) the bench face.
The area of rock enclosed by the above boundaries must be defined in a quantitative
way to estimate the volume of rock broken by each hole in the first row. In order to
complete this task, the distance between the points of intersection of the main breakage
areas should be found.

Table A-1 Relative coordinates of borehole in the first row and breaking points
(Blast # 1)

Hole ID Relative Coordinates
of Borehole
Relative Coordinates of
Breaking
Points to the left of a
Charge
Relative Coordinates of
Breaking
Points to the right of a
Charge

X (m) Y (m) X (m) Y (m) X (m) Y (m)
1 8 50 3.0104 54.88 12.9896 54.88
2 12.88 50 7.8884 54.88 17.8677 54.88
3 17.76 50 12.7665 54.88 22.7457 54.88
4 22.63 50 17.6465 54.88 27.6238 54.88
5 27.51 50 22.5226 54.88 32.5018 54.88
6 32.39 50 27.4006 54.88 37.3799 54.88
7 37.27 50 32.2787 54.88 42.2579 54.88
8 42.15 50 37.1567 54.88 47.136 54.88
9 47.02 50 42.0348 54.88 52.014 54.88
10 51.02 50 42.9128 54.88 56.8921 54.88
11 56.78 50 51.7909 54.88 61.7701 54.88
12 61.66 50 56.6689 54.88 66.6482 54.88
13 66.54 50 61.5469 54.88 71.5262 54.88


132
The intersection of two straight lines can be found if the slopes of the lines are known.
The slopes of the lines can be calculated as follows:

m =
1 2
1 2
x x
y y

(A.11)
where: x
2
, y
2
- relative coordinates of borehole in the first row;
x
1
, y
1
-relative coordinates of breakage point.
To find the point of intersection of two lines, both equations of the lines may be
written in the form y = mx + b and then solved for the unknown x and y.
The results of the calculation of the X coordinate of breaking points, that are needed for
the calculations in the next step, are presented in Table A-2.

Table A-2 Relative coordinates X of intersection
of the main breakage areas for each hole
in the first row for the case # 1



Hole ID
Relative Coordinate X of the
Intersection of the Breakage Areas
1-2 10.439
2-3 15.3171
3-4 20.1951
4-5 25.0732
5-6 29.9512
6-7 34.8293
7-8 39.7073
8-9 44.5854
9-10 49.4634
10-11 54.3415
11-12 59.2195
12-13 64.0976




133
Step # 8
In this step, the rock volume within the boundaries for each hole in the first row is
estimated by using the following formula:
V
i
= A
i
S
i
, cu. ft. (cu.m.) (A.12)

where: A
i
- the profile surface area, sq. ft. (sq. m.);
S
i
- the distance between the points of intersection of the main
breakage areas, ft (m).
The results of the rock volume estimation for the case # 1 are presented in Table A.1.1.1
in the appendix A of this thesis.

Step # 9
Once the volume of rock broken by each hole is determined, the average fragment
size for each hole in the first row could be predicted using equation (2.5.1.2). Then the
average fragment size and size distribution for the first row of blastholes and a given
blast are calculated by using Equations (2.5.1.2) and (2.5.1.12), respectively. The result
of the size distribution for the case # 1 (blast # 1) are presented in Table 5.1 and Figure
5.1.






134
Appendix B Profiling the Quarry Bench Face


In order to properly lay out a drill pattern for an open pit or quarry, accurate
measurements of bench dimensions are a necessity. Too often this is not adhered to by
drilling and blasting crews. What may appear ideal on paper can become a nightmare
muckpiles due to poor measuring techniques. There are a number of methods ranging
from excellent to poor.
Using markers behind the previous shot is the simplest method to determine the first
row burden on the succeeding shot. This may be the only means if a shot must be laid out
before the preceding muckpile is cleaned away. Mining operators should be discouraged
from doing this since surprises often occur in the unseen face once it is fully
uncovered. Also the bank may be changing, undetectable to the naked eye.
Another fairly accurate way is a pole with a hook or eye at one end for a tape to run
through. This is generally not applicable where there is an excess burden. This method
leaves a large margin for error.
One of the best methods is the use of an inclinometer. With it and a basic knowledge
of geometry, relatively precise values of the bench height and burden can be determined
at any selected point on the bank. The information necessary to do the mathematical
operations is easily attained from a measuring tape and a table of natural functions of
angles and the readings obtained from the inclinometer.
Some of the most important parameters for execution of efficient blasting are the true
burden on the blastholes and the real bench height. Measuring the true bench height and
burden has, until recently, been a difficult and tedious operation. However, a relatively
new development such as laser surveying equipment has been available for some years

135
now. It can receive a beam reflected directly off the rock (as opposed to using specially
designed reflectors).


FIRST ROW OF DRILLHOLES CREST

TOE BENCH FACE

Figure B-1 Laser surveying equipment used for profiling the face (courtesy Measurement Devices Ltd.)

Figure B-1 shows how the system is set and the shots taken. The rock face is covered
by a grid of points, which are stored in the instrument on board memory. Once down
loaded into a computer the special software calculates the coordinates of each point and
the values of true burden are obtained. It is possible to locate collars of the holes already
drilled or to be drilled.
Where the blast holes have been already drilled the software gives the true profile for
each blasthole. This enables the blast design is changed details to be planned before

136
actual blast. This technique has been used in this thesis for the estimation of the true
bench height and burden in a field verification of the proposed rock blasting
fragmentation prediction model. The bench face profiles data is presented below.





















137

Table B.1.1 Results of the bench face profiling (Blast # 1)
Hole ID Burden, ft

Depth
from the
top to the
bottom of
borehole,
ft

Bench
Height,
ft
Profile
Surface
Area,
sq. ft.
Total Profile
Surface Area
Per Hole, sq. ft.

1 16.03 0 56.13 1157.2961
16.03 0.01 0.1603
15.48 9.6 151.0905
17.66 22.97 221.5409
18.63 31.29 150.9664
20.71 39.44 160.3105
30.29 47.17 197.115
31.32 55.94 270.1599
31.34 56.13 5.9527
2 15.57 0 54.65 1463.55715
15.57 0.12 1.8684
22.82 11.83 224.7735
25.36 19.43 183.084
25.7 21.19 44.9328
29.05 35.19 383.25
28.51 42.92 222.4694
37.28 50.7 255.9231
37.28 54.65 147.256
3 15.78 0 57.55 1476.03195
15.79 0.19 2.99915
17.94 7.1 116.5372
24.94 19.39 263.4976
28.89 31.2 317.8662
28.63 42.29 318.9484
30.37 55.14 379.075
33.38 57.36 70.7625
33.42 57.55 6.346
4 15.61 0 56.25 1569.1691
15.62 0.69 10.77435
27.51 17.22 356.4695
31.61 36.98 584.1056
31.51 51.55 459.8292
35.72 56.25 157.9905
5 14.81 0 54.54 1089.93735
14.84 0.21 3.11325

138
Hole ID Burden, ft

Depth
from the
top to the
bottom of
borehole,
ft

Bench
Height,
ft
Profile
Surface
Area,
sq. ft.
Total Profile
Surface Area
Per Hole, sq. ft.

5 18.33 7.32 117.9194
19.38 17.88 199.1088
21.49 29.65 240.52
19.35 40.94 230.5418
22.3 49.86 185.759
25.98 54.54 112.9752
6 15.9 0 54.87 1282.95855
15.91 0.23 3.65815
20.89 7.8 139.288
23.45 19 248.304
24.3 32.42 320.4025
24.72 45.64 324.0222
28.77 54.66 241.2399
28.79 54.87 6.0438
7 15.72 0 55.69 1144.63605
15.72 0.15 2.358
19.35 10.17 175.7007
19.81 19.62 185.031
22.1 40.59 439.4264
22.03 50.64 221.7533
25.64 55.69 120.3668
8 16.8 0 56.7 1253.9973
16.9 0.9 15.165
19.39 12.57 211.7522
23.66 30.78 391.9703
24.99 40.57 238.1418
23.12 51.55 264.1239
28.47 56.7 132.8443
9 15.77 0 56.59 1166.2503
15.78 0.69 10.88475
16.96 17.84 280.7455
22.3 30.6 250.4788
25.04 39.89 219.8943
22.4 51.12 266.3756
28.01 56.59 137.8714
10 18.35 0 55.81 1385.43965
18.36 0.96 17.6208
23.4 11.82 226.7568
23.59 22.09 241.2937
26.27 34.94 320.3505
26.07 46.31 297.5529
33.27 55.81 281.865

139
Hole ID Burden,
ft

Depth
from the
top to the
bottom of
borehole,
ft

Bench
Height,
ft
Profile
Surface
Area,
sq. ft.
Total Profile
Surface Area
Per Hole, sq. ft.

11 18.35 0 55.81 1385.43965
18.36 0.96 17.6208
23.4 11.82 226.7568
23.59 22.09 241.2937
26.27 34.94 320.3505
26.07 46.31 297.5529
33.27 55.81 281.865
12 15.91 0 54.96 917.8852
15.91 0.04 0.6364
15.19 10.5 162.653
17.17 27.7 278.296
16.91 39.84 206.8656
16.53 49.36 159.1744
22.4 54.54 100.8287
22.51 54.96 9.4311
13 15.94 0 55.72 880.8016
15.94 0.03 0.4782
15.22 11.44 177.7678
15.12 24.4 196.6032
13.69 38.82 207.7201
15.55 46.05 105.7026
24.27 55.72 192.5297











140

Table B.1.1.1 Results of the estimation of rock volume in the first row of blastholes
(Blast # 1)

Hole ID Bench Height,
ft
Total
Profile
Surface
Area, sq.
ft.

Spacing
Between
Holes, ft

Volume Per
Hole, cu. ft.

Crest
Burden, ft
Toe
Burden,
ft
1 56.13 1157.296 16 18516.7376 16.03 31.34
2 54.65 1463.557 16 23416.9144 15.57 37.28
3 57.55 1476.032 16 23616.5112 15.78 33.42
4 56.25 1569.169 16 25106.7056 15.61 35.72
5 54.54 1089.937 16 17438.9976 14.81 25.98
6 54.87 1282.959 16 20527.3368 15.9 28.79
7 55.69 1144.636 16 18314.1768 15.72 25.64
8 56.7 1253.997 16 20063.9568 16.8 28.47
9 56.59 1166.25 16 18660.0048 15.77 28.01
10 55.81 1385.44 16 22167.0344 18.35 33.27
11 55.81 1385.44 16 22167.0344 15.91 33.27
12 54.96 917.8852 16 14686.1632 15.91 22.51
13 55.72 880.8016 16 14092.8256 15.94 24.27




U. S. System 55.79 208.00 258774.40 16.01 28.59
Average bench height,
ft (m)
Total
Distance,
ft (m)

Total
volume in
first row of
holes, cu. ft.
(cu. m.)

Average
crest
burden,
ft (m)



Average
toe
burden,
ft (m)
Metric System 17.01 63.41 7327.66 4.88 8.72







141
Table B.1.2 Results of the bench face profiling (Blast # 2)

Hole ID
Burden, ft

Depth
from the
top to the
bottom of
borehole,
ft

Bench
Height,
ft
Profile
Surface
Area,
sq. ft.
Total Profile
Surface Area
Per Hole, sq. ft.

1 12.03 0 42.42 1242.263
12.04 0.32 3.8512
19.68 7.68 116.7296
34.29 15.59 213.45135
25.26 22.81 214.9755
26.88 27.64 125.9181
37.21 33.98 203.1653
43.19 38.64 187.332
48.49 41.45 128.8104
50.54 42.42 48.02955
2 12.08 0 42.68 1284.2236
12.09 0.67 8.09695
15.7 4.69 55.8579
26.03 16.16 239.32155
31.4 24.16 229.72
35.07 27.6 114.3284
38.91 32.65 186.7995
40.77 36.78 164.5392
43.3 37.82 43.7164
53.1 41.29 167.254
53.66 41.33 2.1352
53.68 42.68 72.4545
3 12.46 0 39.51 1085.47015
12.47 0.32 3.9888
19.95 8.49 132.4357
19.83 13.48 99.2511
29.22 20.03 160.63875
35.53 25.54 178.38625
35.31 29.06 124.6784
35.44 29.21 5.30625
36.7 34.37 186.1212
38.36 37.98 135.4833
39 39.51 59.1804

142
Hole ID Burden, ft

Depth
from the
top to the
bottom of
borehole,
ft

Bench
Height,
ft

Profile
Surface
Area,
sq. ft.
Total Profile
Surface Area
Per Hole, sq. ft.

4 12.08 0 42.68 1284.9625
12.09 0.67 8.09695
15.7 4.69 55.8579
26.03 16.16 239.32155
31.4 24.16 229.72
35.07 27.6 114.3284
38.91 32.65 186.7995
40.77 36.78 164.5392
43.3 37.82 43.7164
53.65 41.33 170.14725
53.66 42.46 60.63015
53.66 42.68 11.8052
5 12.74 0 33.41 770.9718
12.75 0.43 5.48035
17.89 6.16 87.7836
19.06 12.12 110.111
22.58 16.86 98.6868
26.74 21.76 120.834
29.32 25.08 93.0596
30.96 29.96 147.0832
31.61 33.41 107.93325
6 12.29 0 36.68 871.50135
12.3 0.3 3.6885
16.96 6.44 89.8282
17.92 9.2 48.1344
19.98 14.52 100.814
29.58 25.24 265.6416
31.96 29.21 122.1569
31.99 32.14 93.68675
32.7 35.81 118.70615
33.61 36.68 28.84485
7 12.22 0 36.73 802.93235
12.23 0.22 2.6895
16.44 6.77 93.89425
23.63 18.26 230.20215
26.02 28.26 248.25
26.05 29.35 28.37815
28.02 36.73 199.5183
8 13.8 0 37.34 877.6898
13.9 1.76 24.376
18.91 9.71 130.41975
22.54 16.62 143.20975

143
Hole ID Burden, ft

Depth
from the
top to the
bottom of
borehole,
ft

Bench
Height,
ft

Profile
Surface
Area,
sq. ft.
Total Profile
Surface Area
Per Hole, sq. ft.

8 25.53 22.78 148.0556
29.01 28 142.3494
30.05 32.37 129.0461
30.81 32.63 7.9118
33.87 37.34 152.3214
9 13.8 0 37.34 881.7011
18.91 9.71 158.80705
22.54 16.62 143.20975
25.53 22.78 148.0556
29.01 28 142.3494
30.05 32.37 129.0461
30.81 32.63 7.9118
33.87 37.34 152.3214
10 13.06 0 40.49 706.50035
13.07 1.25 16.33125
15.19 7.3 85.4865
16.79 14.85 120.7245
17.96 20.83 103.9025
17.6 27.09 111.3028
17.99 31.8 83.81445
20.93 36.3 87.57
24.78 39.96 83.6493
26.99 40.49 13.71905
11 12.24 0 33.18 649.7955
17.82 4.33 65.0799
17.82 6 29.7594
19.49 12.74 125.7347
20.75 18.46 115.0864
21.3 22.69 88.93575
20.95 26.44 79.21875
20.62 28.58 44.4799
22.25 32.11 75.66555
26.04 33.18 25.83515
12 12.07 0 37.9 823.08985
19.85 8.08 128.9568
24.02 19.76 256.2008
23.65 22.5 65.3079
23.34 26.12 85.0519
23.16 29.51 78.8175
23.69 31.88 55.51725
24.05 35.59 88.5577
31.95 37.9 64.68

144
Hole ID Burden, ft

Depth
from the
top to the
bottom of
borehole,
ft

Bench
Height,
ft
Profile
Surface
Area,
sq. ft.
Total Profile
Surface Area
Per Hole, sq. ft.

13 12.21 0 37.62 631.62005
15.33 5.06 69.6762
16.23 11.02 94.0488
15.15 15.55 71.0757
16.13 21.44 92.1196
16.61 26.26 78.9034
17.03 30.84 77.0356
20.25 32.34 27.96
22.89 34.85 54.1407
25.24 37.62 66.66005

















145
Table B.1.2.1 Results of the estimation of rock volume in the first row of blastholes
(Blast#2)

Hole
ID
Profile
Surface
Area, sq. ft.
Bench
Height,
ft
Spacing
Between
Holes, ft
Volume
Per Hole,
cu. ft.
Crest
Burden, ft
Toe
Burden,
ft
1 1242.263 42.42 12 14907.156 12.03 50.54
2 1284.2236 42.68 12 15410.6832 12.08 53.68
3 1085.47015 39.51 12 13025.6418 12.46 39
4 1284.9625 42.68 12 15419.55 12.08 53.66
5 770.9718 33.41 12 9251.6616 12.74 31.61
6 871.50135 36.68 12 10458.0162 12.29 33.61
7 802.93235 36.73 12 9635.1882 12.22 28.02
8 877.6898 37.34 12 10532.2776 13.8 33.87
9 881.7011 37.34 12 10580.4132 13.8 33.87
10 706.50035 40.49 12 8478.0042 13.06 26.99
11 649.7955 33.18 12 7797.546 12.24 26.04
12 823.08985 37.9 12 9877.0782 12.07 31.95
13 631.62005 37.62 12 7579.4406 12.21 25.24





U. S. System 38.31 156 142952.66 12.55 35.30
Average
height, ft
(m)
Total
Distance,
ft (m)
Volume
per hole,
cu. ft.
(cu. m.)

Average
crest, ft
(m)
Average
toe, ft
(m)
Metric system 11.68 47.56 4047.96 3.83 10.76









146

Table B.1.3 Results of the bench face profiling (Blast # 3)

Hole ID Burden,
ft

Depth from
the top to the
bottom of
borehole, ft

Bench
Height,
ft
Profile
Surface
Area,
sq. ft.

Total Profile
Surface Area
Per Hole, sq. ft.
1 12.64 0 62.64 1134.40225
12.65 0.83 10.49535
13.09 3.94 40.0257
13.2 15.96 158.0029
17.74 26.65 165.3743
19.21 38.15 212.4625
21.41 56.64 375.5319
34.88 62.1 153.6717
34.89 62.64 18.8379
2 12.88 0 61.52 1214.8878
12.89 0.86 11.0811
12.13 0.97 1.3761
14.89 7.23 84.5726
18.76 15.45 138.3015
20.84 24.52 179.586
21.05 29.07 95.29975
21.72 35.29 133.0147
21.85 58.38 503.01565
21.87 61.52 68.6404
3 12.73 0 61.75 1390.1969
12.76 0.6 7.647
16.35 6.26 82.3813
20.46 19.83 249.75585
22.67 28.34 183.51815
24.01 33.54 121.368
24.01 39.27 137.5773
24.66 43.64 106.34395
25.18 49.99 158.242
25.71 55.27 134.3496
38.35 61.52 200.1875
38.4 61.75 8.82625
4 13.88 0 63.23 1426.13365
13.89 0.93 12.91305
18.92 8.31 121.0689
19.98 14.87 127.592
21.42 24.06 190.233
23.96 29.05 113.2231

147
Hole ID Burden,
ft

Depth from
the top to the
bottom of
borehole, ft

Bench
Height,
ft
Profile
Surface
Area,
sq. ft.

Total Profile
Surface Area
Per Hole, sq. ft.
4 24.5 34.49 131.8112
24.57 43.29 215.908
24.57 47.27 97.7886
24.62 50.29 74.2769
24.67 53.31 74.4279
24.72 57 91.12455
31.52 63.06 170.4072
31.53 63.23 5.35925
5 12 0 62.5 1482.70525
14.87 2.15 28.88525
22.42 16.21 262.1487
24.54 23.33 167.1776
25.02 28.73 133.812
25.11 33.27 113.7951
25.95 56.48 592.5513
35.26 62.48 183.63
35.27 62.5 0.7053
6 12.82 0 63.47 1071.93755
12.83 0.59 7.56675
14.29 23.21 306.7272
16.52 30.63 114.3051
17.66 34.89 72.8034
18.65 40.36 99.30785
19.36 56.69 310.35165
22.05 59.93 67.0842
23.55 60.05 2.736
29.49 63.35 87.516
29.5 63.47 3.5394
7 12.44 0 63.29 1154.32505
12.45 0.28 3.4846
12.53 5.55 65.8223
14.4 13.7 109.73975
17 21.11 116.337
17.24 26.08 85.0864
17.41 30.65 79.17525
19.04 36.05 98.415
20.08 45.39 182.6904
21.22 54.38 185.6435
26.62 58.38 95.68
27.25 63.29 132.25085
8 12.92 0 63.37 1050.73725
12.93 0.29 3.74825
14.29 23.21 311.9412
16.43 31.73 130.8672

148
Hole ID Burden,
ft

Depth from
the top to the
bottom of
borehole, ft

Bench
Height,
ft
Profile
Surface
Area,
sq. ft.

Total Profile
Surface Area
Per Hole, sq. ft.
8 17.96 34.89 54.3362
18.45 40.5 102.13005
20.66 58 342.2125
20.66 58 0.0000
22.05 59.93 41.21515
23.55 60.05 2.736
29.46 62.35 60.9615
29.46 62.37 0.5892
9 12.45 0 62.97 1491.94845
12.46 0.34 4.2347
14.97 2.23 25.92135
21.42 15.91 248.9076
24.44 23.29 169.2234
25.12 28.73 134.8032
25.21 34.47 144.4471
26.12 56.45 564.1167
35.32 62.97 200.2944
10 12.18 0 63.27 1247.5289
12.19 0.36 4.3866
14.61 4.7 58.156
16.82 12.12 116.6053
17.89 18.16 104.8242
18.68 30.72 229.6596
19.15 38.32 143.754
20.28 45.57 142.93375
21.89 52.01 135.7874
21.91 57.18 113.223
32.51 57.18 0
32.58 63.27 198.19905
11 13.08 0 65.06 1393.553
13.09 1.67 21.85195
18.17 7.5 91.1229
19.47 13.35 110.097
20.14 20.34 138.43695
20.9 24.71 89.6724
21.35 34.8 213.15125
21.4 42.93 173.77875
21.83 46.57 78.6786
22.21 53.61 155.0208
24.88 59.55 139.8573
41.14 65.06 181.8851



149
Table B.1.3.1 Results of the estimation of rock volume in the first row of blastholes
(Blast # 3)

Hole ID Profile
Surface
Area,
sq. ft.
Bench
Height,
ft
Spacing
Between
Holes, ft

Volume
Per
Hole, cu.
ft.

Crest
Burden, ft
Toe
Burden,
ft
1 1134.40225 62.64 16 18150.436 12.64 34.89
2 1214.8878 61.52 16 19438.2048 12.88 21.87
3 1390.1969 61.75 16 22243.1504 12.73 38.4
4 1426.13365 63.23 16 22818.1384 13.88 31.53
5 1482.70525 62.5 16 23723.284 12 35.27
6 1071.93755 63.47 16 17151.0008 12.82 29.5
7 1154.32505 63.29 16 18469.2008 12.44 27.25
8 1050.73725 63.37 16 16811.796 12.92 29.46
9 1491.94845 62.97 16 23871.1752 12.45 35.32
10 1247.5289 63.27 16 19960.4624 12.18 32.58
11 1393.553 65.06 16 22296.848 13.08 41.14




U. S.
System
14058.36 63.01 176.00 224933.70 12.73 27.84
Total
profiles area
in first row,
sq. ft. ( sq.
m.)
Average
Height,
ft ( m )
Total
Distance,
ft (m)

Total volume
in first row of
holes, cu. ft
(cu. m.)

Average
crest burden,
ft
(m)

Average
toe
burden, ft
(m)
Metric
System
1306.06 19.21 53.66 6369.40 3.88 8.49









150
Table B.1.4 Results of the bench face profiling (Blast # 4)

Hole ID Burden,
ft
Depth
from the top
to the bottom of
borehole, ft
Bench
Height,
ft
Profile
Surface
Area, sq.
ft.
Total Profile
Surface Area
Per Hole, sq. ft.
1 13.14 0 62.55 1504.6353
13.16 2.75 36.1625
16.28 7.73 73.3056
16.78 11.42 60.9957
17.95 17.71 109.2259
18.65 23.22 100.833
18.71 29.22 112.08
19.23 39.07 186.8545
19.15 44.02 94.9905
20.11 48.1 80.0904
24.89 62.54 324.9
24.9 62.55 325.1973
2 12.78 0 61.83 925.26285
12.76 0.72 9.1944
13.15 15.28 188.6248
15.92 26.72 166.2804
15.05 33.08 98.4846
15 39.79 100.8178
15.65 46.72 106.2023
14.66 51.13 66.83355
15.4 55.69 68.5368
23.39 61.53 113.2668
23.42 61.83 7.0215
3 12.59 0 63.03 1144.9504
12.62 0.22 2.7731
14.37 5.32 68.8245
16.02 11.31 91.01805
16.52 16.53 84.9294
18.48 23.93 129.5
17.93 28.93 91.025
18.01 35.46 117.3441
18.16 38.79 60.22305
18.53 42.26 63.65715
17.76 46.56 78.0235
19.25 51.27 87.15855
20.17 55.44 82.1907
25.43 60.53 116.052
32.14 62.95 69.6597

151
Hole ID Burden,
ft
Depth
from the top
to the bottom of
borehole, ft
Bench
Height,
ft
Profile
Surface
Area, sq.
ft.
Total Profile
Surface Area
Per Hole, sq. ft.
3 32.15 63.03 2.5716
4 13.5 0 61.62 1070.61825
13.53 0.33 4.45995
12.33 4.97 59.9952
15.76 10.88 83.00595
15.71 16.61 90.16155
16.55 21.52 79.1983
16.94 25.49 66.47765
18.26 30.31 84.832
18.76 34.99 86.6268
18.41 39.63 86.2344
17.92 44.18 82.65075
17.85 52.96 157.0303
23.89 58.59 117.4981
23.93 61.62 72.4473
5 15.68 0 59.32 1157.6083
16.59 4.42 71.3167
16.42 10.6 102.0009
18.35 18.41 135.7769
20.53 26.04 148.3272
20.67 31.41 110.622
21.16 35.51 85.7515
20.57 40.06 94.93575
20.39 46.6 133.9392
20.22 49.18 52.3869
20.57 53.23 82.59975
22.7 57.12 84.16015
27.7 59.18 51.912
27.72 59.32 3.8794
6 12.84 0 60.69 1224.60395
12.84 0.14 1.7976
16.3 4.27 60.1741
17.46 10.99 113.4336
18.99 21.5 191.5448
22.18 30.69 189.1762
21.55 36.67 130.7527
21.79 42.37 123.519
20.85 47.1 100.8436
21.49 52.01 103.9447
22.58 56.34 95.41155
29.82 60.68 113.708
29.82 60.69 0.2982
7 13.13 0 61.83 1237.98845
16.17 5.74 84.091

152
Hole ID Burden,
ft
Depth
from the top
to the bottom of
borehole, ft
Bench
Height,
ft
Profile
Surface
Area, sq.
ft.
Total Profile
Surface Area
Per Hole, sq. ft.
7 18.58 12.72 121.2775
17.77 17.86 93.4195
19.42 23.47 104.318
20.14 28.92 107.801
20.77 32.57 74.66075
20.88 37.61 104.958
21.53 41.65 85.6682
20.13 45.86 87.6943
21.8 49.87 84.06965
22.59 53.63 83.4532
23.43 56.85 74.0922
28.73 60.85 104.32
28.75 61.83 28.1652
8 13.13 0 60.61 1158.77955
17.72 2.91 44.88675
16.64 8 87.4462
15.67 11.58 57.8349
17.93 17.92 106.512
19.19 25 131.4048
19.22 27.11 40.52255
19.36 34.59 144.2892
20.19 39.97 106.3895
20.53 44.28 87.7516
22.01 56.14 252.2622
22.5 60.61 99.47985
9 12.9 0 63.44 1294.9396
12.84 0.89 11.4543
17.19 6.45 83.4834
20.45 19.16 239.2022
20.92 24.07 101.5634
21.01 29.03 103.9864
21.51 36.42 157.1114
21.33 39.27 61.047
21.66 43.25 85.5501
21.99 45.93 58.491
22.36 50.3 96.90475
21.6 52.98 58.9064
21.98 55.92 64.0626
22.02 59.33 75.02
25.63 63.31 94.8235
25.65 63.44 3.3332
10 13.02 0 61.3 957.08055
13.03 0.02 0.2605
13.27 5.23 68.5115

153
Hole ID Burden,
ft
Depth
from the top
to the bottom, ft
Bench
Height,
ft
Profile
Surface
Area, sq.
ft.
Total Profile
Surface Area
Per Hole, sq. ft.
10 14.23 13.34 111.5125
14.51 17.61 61.3599
15.63 21.46 58.0195
15.69 25.94 70.1568
16.15 30.04 65.272
16.05 33.69 58.765
16.98 36.8 51.36165
16.56 39.55 46.1175
16.05 41.75 35.871
16.78 45.54 62.21285
16.76 49.66 69.0924
16.99 52.92 55.0125
16.88 56.14 54.5307
17.43 59.53 58.15545
17.45 61.3 30.8688
11 13.2 0 62.01 1247.16875
13.21 0.36 4.7538
15.6 5.11 68.42375
17.96 11.06 99.841
18.26 15.3 76.7864
18.87 19.13 71.10395
20.02 23.39 82.8357
20.18 30.04 133.665
20.4 34.69 94.3485
20.47 37.83 64.1659
21.65 41.28 72.657
22.62 46.06 105.8053
22.17 52.82 151.3902
22.94 56.87 91.34775
27.35 61.65 120.1931
27.38 62.01 9.8514
12 11.6 0 63.16 989.3604
11.64 1.21 14.0602
13.72 3.85 33.4752
13.84 11.07 99.4916
14.84 14.24 45.4578
16.34 21.2 108.5064
4.93 25.72 48.0702
18.02 29.34 41.5395
18.46 32.11 50.5248
17.91 38.29 112.3833
16.96 43.13 84.3854
15.79 46.29 51.745
15.54 52.73 100.8826
19.37 59.95 126.0251

154
Hole ID Burden,
ft
Depth
from the top
to the bottom of
borehole, ft
Bench
Height,
ft
Profile
Surface
Area, sq.
ft.
Total Profile
Surface Area
Per Hole, sq. ft.
12 24.59 62.3 51.653
24.62 63.16 21.1603
13 12.86 0 62.07 886.16375
13.75 4.43 58.94115
12.87 8.39 52.7076
13.85 20.7 164.4616
14.67 24.16 49.3396
13.75 28.14 56.5558
14.4 31.67 49.68475
14.74 35.51 55.9488
14.19 42.87 106.4624
14.03 52.25 132.3518
14.66 54.36 30.26795
15 58.01 54.1295
21.52 61.71 67.562
21.54 62.07 7.7508
14 12.72 0 61.42 864.26455
15.58 4.29 60.7035
15.14 8.17 59.5968
15.37 11.45 50.0364
15.58 16.03 70.8755
15.29 24.04 123.6344
15.01 28.03 60.4485
15.73 37.14 140.0207
15.92 40.61 54.91275
15.88 46.66 96.195
16.25 48.87 35.50365
17.6 52.59 62.961
18.18 55.35 49.3764









155
Table B.1.4.1 Results of the estimation of rock volume in the first row of blastholes
(Blast # 4)


Hole ID Profile
Surface
Area,
sq. ft.
Bench
Height,
ft.
Spacing
Between
Holes,
ft

Volume
Per
Hole,
cu. ft.
Crest
Burden,
ft
Toe
Burden,
ft
1 1504.6353 62.55 16 24074.1648 13.14 24.9
2 925.26285 61.83 16 14804.2056 12.78 32.15
3 1144.9504 63.03 16 18319.2064 12.59 32.15
4 1070.61825 61.62 16 17129.892 13.5 23.93
5 1157.6083 59.32 16 18521.7328 15.68 27.72
6 1224.60395 60.69 16 19593.6632 12.84 29.82
7 1237.98845 61.83 16 19807.8152 13.13 28.75
8 1158.77955 60.61 16 18540.4728 13.13 22.5
9 1294.9396 63.44 16 20719.0336 12.9 25.65
10 957.08055 61.3 16 15313.2888 13.02 17.45
11 1247.16875 62.01 16 19954.7 13.2 27.38
12 989.3604 63.16 16 15829.7664 11.6 24.62
13 886.16375 62.07 16 14178.62 12.86 21.54
14 864.26455 61.42 16 13828.2328 12.72 18.18





U. S.
System
15663.42 61.78 224.00 250614.79 13.08 23.14
Total surface
area in first
row,
sq. ft. ( sq. m.)
Average
Bench.
Height,
ft (m)
Total
Distance, ft
(m)
Total volume
in first row of
holes, cu. ft.
(cu. m.)
Average
crest
burden, ft
(m)

Average
toe
burden,
ft (m)
Metric
System
1455.19 18.83 68.29 7096.61 3.99 7.06








156
Table B.1.5 Results of the bench face profiling (Blast # 5)

Hole ID Burden,
ft
Depth from the
top to the bottom
of borehole, ft
Bench
Height,
ft
Profile
Surface
Area,
sq. ft.
Total Profile
Surface Area
Per Hole,
sq. ft.
1 14.69 0 62.46 1338.27535
14.7 0.3 4.4085
18.23 6.13 95.99095
18.24 11.38 95.73375
18.87 16.45 94.07385
21.27 31.16 295.2297
21.93 37.3 132.624
20.4 43.72 135.8793
31.29 62.46 484.3353
2 14.79 0 63.34 1277.9968
15.9 1.13 17.33985
14.64 6.64 84.1377
15.81 12.16 84.042
18.59 18.74 113.176
20.86 31.23 246.3653
20.8 35.65 92.0686
21.27 39.41 79.0916
22.47 55.67 355.6062
31.29 63.34 206.1696
3 13.78 0 62.02 916.104
14.48 23.62 333.7506
14.01 31.69 114.9572
16.08 59.26 414.7907
22.04 62.02 52.6056
4 13.94 0 61.94 966.59725
14.55 3.58 50.9971
13.7 9.92 89.5525
14.66 16.39 91.7446
15.89 22.99 100.815
16.35 28.96 96.2364
15.49 33.21 67.66
15.92 37.65 69.7302
15.28 43.27 87.672
14.64 47.27 59.84
15.63 52.98 86.42085
15.59 58.59 87.5721
31.19 61.94 78.3565
5 12.85 0 64.11 944.81605
13.47 1.95 25.662
12.91 15.21 174.8994

157
Hole ID Burden,
ft
Depth from the
top to the bottom
of borehole, ft
Bench
Height,
ft
Profile
Surface
Area,
sq. ft.
Total Profile
Surface Area
Per Hole,
sq. ft .
5 13.48 21.42 81.94095
14.21 25.27 53.30325
14.74 30.4 74.25675
15.06 33.48 45.892
14.16 39.09 81.9621
14.36 44 70.0166
15.34 47.05 45.2925
15.23 52.65 85.596
16.34 59.31 105.1281
20.36 62.35 55.784
30.87 64.11 45.0824
6 14.3 0 63.62 1426.9459
14.31 0.11 1.57355
17.3 3.4 51.99845
20.15 8.32 92.127
21.13 12.47 85.656
20.27 15.71 67.068
21.74 20.69 104.6049
23.28 25.26 102.8707
23.29 30.48 121.5477
23.51 34.86 102.492
23.13 40.18 124.0624
22.83 43.88 85.026
22.95 49.47 127.9551
24.35 52.71 76.626
24.67 57.42 115.4421
29.49 63.62 167.896
7 14.25 0 63.81 1380.7088
14.25 0.09 1.2825
23.61 7.78 145.5717
18.77 9.33 32.8445
21.27 15.34 120.3202
20.49 19.37 84.1464
20.51 26.07 137.35
21.6 30.4 91.16815
22.29 34.57 91.51065
22.27 39.62 112.514
21.88 43.79 92.05275
22.38 47.51 82.3236
21.9 52.43 108.9288
22.11 56.35 86.2596
30.93 63.67 194.1264
30.98 63.68 0.30955
8 13.21 0 64.14 1248.5701

158
Hole ID Burden,
ft
Depth from the
top to the bottom
of borehole, ft
Bench
Height,
ft
Profile
Surface
Area,
sq. ft.
Total Profile
Surface Area
Per Hole,
sq. ft.
8 15.6 5.25 75.62625
18.25 10.53 89.364
17.49 15.17 82.9168
16.69 20.33 88.1844
19.29 27.06 121.0727
19.33 31.12 78.3986
19.59 37.29 120.0682
19.75 40.47 62.5506
19.6 46.2 112.7378
20.04 50.38 82.8476
28.27 63.65 320.5369
29.96 64.14 14.26635
9 13.66 0 63 1205.88155
13.41 5.7 77.1495
16.43 12.22 97.2784
18.26 19.02 117.946
18.92 23.62 85.514
19.83 28.7 98.425
19.87 34.29 110.9615
20.63 36.78 50.4225
20.39 42.46 116.4968
20.82 46.87 90.86805
19.69 52.42 112.4153
21.13 57.07 94.9065
30.64 63 153.4981
10 13.48 0 62.13 1007.4326
13.51 0.79 10.66105
12.96 7.73 91.8509
12.17 18.7 137.8381
13.81 24.28 72.4842
14.88 29.28 71.725
16.25 32.15 44.67155
17.18 37.67 92.2668
17.67 41.76 71.26825
18.36 45.81 72.96075
18.37 48.77 54.3604
18.14 54.42 103.1408
19.64 57.62 60.448
31.8 60.86 83.3328
31.86 62.13 40.4241
11 15.04 0 61.02 1239.5488
15.05 0.72 10.8324
14.07 3.36 38.4384
16.32 6.86 53.1825

159
Hole ID Burden,
ft
Depth from the
top of the bottom
of borehole, ft
Bench
Height,
ft
Profile
Surface
Area,
sq. ft.
Total Profile
Surface Area
Per Hole,
sq. ft.
11 17.48 12.73 99.203
17.8 18.72 105.6636
18.07 23.19 80.16945
21.52 30.12 137.1794
21.97 33.07 64.14775
22.44 38.54 121.4614
22.87 42.9 98.7758
23.1 46.94 92.8594
21.38 52.58 125.4336
24.94 57.91 123.4428
32.14 61.02 88.7594
12 14.21 0 60.66 1135.4712
14.91 5.39 78.4784
16.12 12.4 108.7602
17.52 19.46 118.7492
17.62 27.25 136.8703
19.48 30.53 60.844
20.87 37.93 149.295
19.65 42.28 88.131
19.84 46.8 89.2474
20.07 52.28 109.3534
21.52 57.01 98.36035
31.84 60.66 97.382
13 14.45 0 61.54 1192.17055
14.45 0.19 2.7455
15.82 5.13 74.7669
17.64 8.95 63.9086
17.65 14.77 102.6939
19.25 16.1 24.5385
19.28 20.66 87.8484
19.05 25.33 89.50055
17.67 30.46 94.1868
18.76 35.79 97.08595
19.38 40.17 83.5266
21.2 44.37 85.218
21.02 49.48 107.8721
21.1 54.33 102.141
22.55 57.99 79.8795
31.68 61.54 96.25825
14 14.23 0 61.31 1065.6868
15.46 2.11 31.32295
14.8 5.96 58.2505
14.86 10.88 72.9636
12.58 16.93 83.006

160
Hole ID Burden,
ft
Depth from the
top to the bottom
of borehole, ft
Bench
Height,
ft
Profile
Surface
Area,
sq. ft.
Total Profile
Surface Area
Per Hole,
sq. ft.
14 15.14 21.84 68.0526
18.14 29.01 119.3088
19.53 33.71 88.5245
19.33 37.99 83.1604
18.8 42.2 80.26365
18.52 48.67 120.7302
18.16 51.38 49.7014
19.87 57.89 123.7877
21.14 58.42 10.86765
31.28 61.31 75.7469

















161
Table B.1.5.1 Results of the estimation of rock volume in the first row of blastholes
(Blast # 5)

Hole ID Profile
Surface
Area,
sq. ft.
Bench
Height,
ft.
Spacing
Between
Holes,
ft

Volume
Per
Hole,
cu. ft.
Crest
Burden,
ft
Toe
Burden,
ft
1 1338.27535 62.46 16 21412.4056 14.69 31.29
2 1277.9968 63.34 16 20447.9488 14.79 31.29
3 916.104 62.02 16 14657.664 13.78 22.04
4 966.59725 61.94 16 15465.556 13.94 31.19
5 944.81605 64.11 16 15117.0568 12.85 30.87
6 1426.9459 63.62 16 22831.1344 14.3 29.49
7 1380.7088 63.81 16 22091.3408 14.25 30.98
8 1248.5701 64.14 16 19977.1216 13.21 29.96
9 1205.88155 63 16 19294.1048 13.66 30.64
10 1007.4326 62.13 16 16118.9216 13.48 31.86
11 1239.5488 61.02 16 19832.7808 15.04 32.14
12 1135.4712 60.66 16 18167.5392 14.21 31.84
13 1192.17055 61.54 16 19074.7288 14.45 31.68
14 1065.6868 61.31 16 17050.9888 14.23 14.23





U. S.
System
16346.21 62.51 224.00 261539.29 14.06 23.30
Total surface
area in first
row,
sq. ft. ( sq. m.)
Average
Bench
Height,
ft (m)
Total
Distance, ft
(m)
Total volume
in first row of
holes, cu. ft
(cu. m.)
Average
crest
burden, ft
(m)

Average
toe
burden, ft
(m)
Metric
System
1518.62 19.06 68.29 7405.96 4.29 7.10









162
Appendix C Using the Image-Processing Program
Split-Desktop

Split-Desktop is an image-processing program designed to calculate the size
distribution of rock fragments through analyzing digital grayscale images. Digital
grayscale images can be acquired manually through use of a digital camera. The Split-
Desktop software is the result of over nine years of research and development at the
University of Arizona and Julius Kruttschnitt Mineral Research Center of Brisbane,
Queensland, Australia (Higgings et al., 1999).
There are five steps for using Split-Desktop as listed below:
1. Ready the Image: crop and scale image;
2. Find Particles: produce binary images from grayscale images;
3. Edit the Image;
4. Compute Sizes;
5. Output Results.
The first step is to crop and scale the grayscale image that one wants to process. The
second step is to produce the binary or segmented image that shows the outlines of the
particles visible in the image. The third step is to edit the grayscale image. The fourth
step is to calculate the size distribution from the segmented image. The final step is to
plot and save results.
Numerous images can be processed together and the results from each image will be
weighted together according to the scale and areas of the particles in each image for one
combined size distribution output file. The actual number of images Split-Desktop can

163
process as a group depends on the memory availability in the computer, and in some
cases one may be able to process up to 100 images (Split Engineering LLC, 2001).
Split-Desktop manipulates, displays and analyzes only digital images. It can
open/import TIFF (.tif) and JPEG (.jpg) formatted images and uses TIFF formatted
images for analysis. JPEG images opened in Split-Desktop will be converted to TIFF
format automatically by the software during the processing steps. Digital images are two
dimensional array of pixels (picture elements). Pixels are represented by 8-bit unsigned
integers, ranging in value from 0 to 255. Split-Desktop displays pixels with the value of
zero (0) as white and those with a value of 255 as black.
As can be seen from the picture (Fig. C-1) the objects of known size must be in the
picture in order to set the scale and slope for the entire image that is to be analyzed. The
change in apparent size of objects due to the top of a pile being further away from the
observer than the bottom is also corrected by using the scaling information. If no slope is
present in the image, one object can be placed in the image to set the scale. Otherwise,
one must place two scaling objects on the pile so that both are in the field of the view of
the image that is to be taken. The scaling objects should be at different heights within the
image to correct the effect of slope on the scale. The best scaling tools for this method are
rubber balls with handles on them so that a rope can be tied between them in order to
retrieve the balls.
The step by step Split-desktop procedures are briefly described below. The first step
is to crop and scale the grayscale image that must be analyzed. At this step unnecessary
portions should be cut such as the sky, ground in the front of the pile or machinery in the
picture. Then the new file name must be given for a new image.

164



Figure C-1 Typical JPEG image is taken in the field for the calculation of
size distribution.

Scaling of the image is a critical step. The accuracy of size distribution output
depends on proper scaling of the image. In this step one has to select the button Scaling
Tools on the vertical toolbar of the screen and click on edge of the objects of known
size. The left mouse button should be depressed and one must drag the line to the other
edge of the object.
Then the Scale Image menu item must be selected. This will open the scale dialog
box where scale must be specified. It automatically fills the respective image with this
value and the image is scaled. It will also automatically saved as a grayscale TIFF image.
Once the image is scaled, it should not be resized or cropped.


165


Figure C-2 Typical grayscale image.

The second step, so called Find Particles, is to produce the binary image that
shows the outlines of the particles visible in the image. In this step Split-Desktop
performs the automatic delineation of the particles. For the program to identify the
particles for size calculation, it converts the grayscale image to a binary image. As can be
seen from the picture above (Fig. C-2), the binary image actually allows three color
levels: white for particles, black for boundary and fines, and gray for areas that do not
count in the size distribution.
Next step, so called Edit the Images, is to edit the grayscale image. In this step the
particle edges should be clearly outlined in a continuous line, patches of fine material

166
should be noted by filling in, and the objects within the image that are not to be sized
should be removed. They are the sky, loading equipment and the floor in front of the pile.


Figure C-3 Typical binary image.

When it is done, editing menu Done editing must be selected. The grayscale image
is discarded, the particles are changed to white and the areas that have been edited out
will be changed to gray and the areas that are black will remain black. The binary image
is automatically renamed and saved with the binary prefix (bi-filename). Typical bynary
image is presented above(Figure C-3).
Usually, groups of binary images are to be processed. Unless only one image is to be
analyzed, one should return to Step One and begin to work on the next grayscale image.
Once the binary images have been completely edited, they can be processed to
calculate the size distribution. It is the fourth step of the Split-Desktop so called
Compute Sizes.

167
To process a group of images together one must open all images in Split-Desktop
and check the box Include All Open Images before processing. Computing the size
distribution as a group is the only way to make sure that the results of images taken at
different scales are combined correctly. The results from single images cannot be simply
averaged. Then Split-Desktop begins to fill in each individual particle within the image
and calculates the size of each particle. The method of size calculation is well described
in the Chapter 6 of the software manual (Split Engineering LLC, 2001). When Split-
Desktop completed with the size calculation step, the binary images are reverted back to
normal (black, white and gray color) and will remain open on the computer desktop. All
that is left to be completed is to output the size distribution file and save it. The Split-
Desktop has been used in the field verification of a proposed rock blasting fragmentation
engineering model FRAGM developed in this thesis. The Split-Desktop output files of
the fragment size distributions for the blasts at a limestome surface mine are presented
below.

Blast # 1 Split-Desktop output file

Date: Wed May 14 09:42:08 2003

Split Desktop Academic License
Sieve series: ASTM Units: (mm) Number of Images 16

Cumulative Percent Passing Data

Size All Images

1450.00 100.00
1400.00 100.00
1350.00 100.00
1300.00 100.00
1250.00 100.00
1200.00 100.00
1150.00 100.00
1100.00 100.00
1050.00 100.00

168
1000.00 99.66
950.00 99.94
750.00 98.52
700.00 97.25
650.00 95.93
600.00 94.22
550.00 91.93
500.00 89.46
450.00 85.68
400.00 79.35
350.00 75.02
300.00 68.89
250.00 59.61
200.00 47.41
150.00 35.48
100.00 21.43
50.00 7.66


Following Data in (mm)

F10 58.51
F20 98.07
F30 130.5
F40 168.94
F50 239.30
F60 251.10
F70 309.05
F80 405.13
F90 521.05
Topsize 888.03

Finescutoff: 48.00
Fines factor: 50.00

RosRam uniformity: 1.39
RosRam X50: 23.93
R-squared: 0.9833

Schuhmann Slope: 0.97
Schuhmann X50: 26.42
R-squared: 0.9710















169
Blast # 2 Split-Desktop output file


Date: Sat May 17 11:11:39 2003

Split Desktop Academic License
Sieve series: ASTM Units: (mm) Number of Images 15

Cumulative Percent Passing Data

Size All Images

2000.00 100.00
1475.00 100.00
1450.00 100.00
1400.00 100.00
1350.00 100.00
1300.00 100.00
1250.00 100.00
1200.00 100.00
1150.00 100.00
1100.00 100.00
1050.00 99.91
1000.00 99.70
950.00 99.40
900.00 98.98
850.00 98.42
800.00 97.64
750.00 96.55
700.00 95.13
650.00 93.33
600.00 90.97
550.00 88.33
500.00 85.31
450.00 81.70
400.00 77.63
350.00 73.07
300.00 68.04
250.00 61.64
200.00 53.51
150.00 42.76
100.00 30.08
50.00 12.43



Following Data in (mm)

F10 42.60
F20 71.79
F30 99.77
F40 139.06
F50 182.04
F60 238.91
F70 318.19
F80 428.43

170
F90 580.93
Topsize 962.65

Finescutoff: 100.00
Fines factor: 50.00

RosRam uniformity: 1.13
RosRam X50: 18.20
R-squared: 0.9947

Schuhmann Slope: 0.74
Schuhmann X50: 22.88
R-squared: 0.9384


Blast # 3 Split-Desktop output file
Date: Wed May 21 16:17:48 2003

Split Desktop Academic License
Sieve series: ASTM Units: (mm) Number of Images 16

Cumulative Percent Passing Data

Size All Images

2000.00 100.00
1475.00 98.66
1450.00 98.51
1400.00 98.14
1350.00 97.66
1300.00 97.08
1250.00 96.45
1200.00 95.79
1150.00 94.99
1100.00 94.08
1050.00 93.03
1000.00 91.89
950.00 90.57
900.00 89.09
850.00 87.40
800.00 85.46
750.00 83.28
700.00 80.95
650.00 78.40
600.00 75.51
550.00 72.30
500.00 68.79
450.00 64.75
400.00 60.02
350.00 54.49
300.00 48.21
250.00 41.23

171
200.00 34.48
150.00 23.41
100.00 17.22
50.00 6.39



Following Data in (mm)

F10 66.67
F20 122.46
F30 142.96
F40 241.38
F50 313.61
F60 399.83
F70 516.75
F80 680.75
F90 930.04
Topsize 1804.10

Finescutoff: 24.00
Fines factor: 50.00

RosRam uniformity: 1.27
RosRam X50: 31.36
R-squared: 0.9662

Schuhmann Slope: 0.85
Schuhmann X50: 33.14
R-squared: 0.9453




Blast # 4 Split-Desktop output file


Date: Mon May 19 10:13:12 2003

Split Desktop Academic License
Sieve series: ASTM Units: (mm) Number of Images 18

Cumulative Percent Passing Data

Size All Images


1475.00 100.00
1450.00 100.00
1400.00 100.00
1350.00 100.00
1300.00 100.00
1250.00 100.00
1200.00 100.00
1150.00 100.00
1100.00 100.00

172
1050.00 100.00
1000.00 100.00
950.00 100.00
900.00 99.14
850.00 98.28
800.00 97.41
750.00 96.55
700.00 92.82
650.00 91.40
600.00 87.51
550.00 85.17
500.00 82.11
450.00 79.17
400.00 74.17
350.00 67.08
300.00 61.28
250.00 54.14
200.00 43.51
150.00 31.59
100.00 18.93
50.00 6.69



Following Data in (mm)

F10 63.52
F20 104.23
F30 143.72
F40 185.28
F50 230.53
F60 291.04
F70 370.59
F80 464.12
F90 632.01
Topsize 941.86

Finescutoff: 39.00
Fines factor: 50.00

RosRam uniformity: 1.36
RosRam X50: 23.05
R-squared: 0.9889

Schuhmann Slope: 1.13
Schuhmann X50: 28.06
R-squared: 0.9087








173
Blast # 5 Split-Desktop output file
Date: Tue May 20 15:18:44 2003

Split Desktop Academic License
Sieve series: ASTM Units: (mm) Number of Images 12

Cumulative Percent Passing Data

Size All Images

2000.00 100.00
1475.00 100.00
1450.00 100.00
1400.00 99.89
1350.00 99.85
1300.00 99.80
1250.00 99.72
1200.00 99.69
1150.00 99.60
1100.00 99.51
1050.00 99.41
1000.00 99.32
950.00 97.76
900.00 97.14
850.00 96.52
800.00 95.90
750.00 95.28
700.00 93.44
650.00 89.66
600.00 87.24
550.00 84.13
500.00 82.70
450.00 76.06
400.00 70.25
350.00 64.34
300.00 56.33
250.00 48.38
200.00 35.62
150.00 23.13
100.00 12.62
50.00 3.34



Following Data in (mm)

F10 35.88
F20 135.11
F30 177.50
F40 217.16
F50 260.13
F60 322.74
F70 397.89
F80 479.67
F90 654.50

174
Topsize 1095.56
Finescutoff: 63.00
Fines factor: 50.00

RosRam uniformity: 1.46
RosRam X50: 26.01
R-squared: 0.9851

Schuhmann Slope: 1.25
Schuhmann X50: 28.78
R-squared: 0.9079























175
Vita


Sergey Victorovych Zagreba was born in Krivoi Rog, Ukraine on July 7, 1975. He is
the son of Zagreba Victor Grigoryevich and Lyudmila Yakovlevna . He attended a school
in Krivoi Rog and graduated from Kiev Suvorov Military School (Kiev, Ukraine) in
1992. He later joined the Technical University of Krivoi Rog (former Krivoi Rog Ore
Mining Institute) where he obtained his bachelors and masters degrees in mining
engineering in 1996 and 1998, respectively. He was employed as an assistant to a surface
mine foreman and a surface mine foreman in the quarry of the Mining and Enriching
Combine Southern (OMCW Yuzhnyy), Krivoi Rog, Ukraine.
In June 2000, he moved to the U.S. and lived with his wife, Mary Beth, in
Charleston, South Carolina, where he worked in real estate business. In August 2002,
Sergey joined the Department of Mining Engineering at the West Virginia University for
a masters program in mining engineering. He is expected to graduate in December 2003
and would like to pursue a career in the mining industry.

You might also like