An Engineering Approach To Predict The Proportion of Fines PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

An engineering approach to predict the proportion of fines

generated by blasting

I. Onederra1, S. Esen1& H.A. Bilgin2


1
Julius Kruttschnitt Mineral Research Centre, The University of Queensland, Brisbane, Qld, Australia
2
Department of Mining Engineering, Middle East Technical University, Ankara, Turkey

Italo Onederra BE(Hons) MEngSc.


Senior Research Engineer - Mining Research
Julius Kruttschnitt Mineral Research Centre (JKMRC)
Isles Rd. Indooroopilly
QLD. Australia 4068
Phone: +61 7 3365 5888
Fax: +61 7 3365 5999
e-mail: I.Onederra@mailbox.uq.edu.au

1
An engineering approach to predict the proportion of fines
generated by blasting

I. Onederra1, S. Esen1& H.A. Bilgin2


1
Julius Kruttschnitt Mineral Research Centre, The University of Queensland, Brisbane, Qld, Australia
2
Department of Mining Engineering, Middle East Technical University, Ankara, Turkey

ABSTRACT

A new model to predict the proportion of fines generated during blasting is presented
in this paper. Within the context of this work, fines are defined as the proportion of
material less than 1.18 mm. The model is based on the back-analysis of a
comprehensive experimental program that included the fragmentation assessment and
direct measurement of the zone of crushing from 92 blasting tests on concrete blocks
using two commercial explosives. The properties of the concrete blocks varied from
low, medium to high strength and measured 1.5 m in length, 1.0 m in width and 1.0 m
in height. Sieve analysis showed that the percentage of fines is directly proportional to
the percentage of the mass of crushed material given by the cylindrical volume of
crushing around a blasthole. The proposed model combines two specific
developments arising from the experimental program, first is the relationship between
the assumed 1.18 mm cut off point and the volume of crushing; and secondly but
equally important is the development of a model to predict the size of the crushing
zone around a blasthole.

1 INTRODUCTION

For a number of years, both researchers and practising engineers have been aware of
the importance of being able to tailor blast fragmentation to optimise the overall
mineral extraction and recovery cycle. It is widely acknowledged that in production
blasting a significant proportion of fines present in a muckpile originates from the
zone of crushing produced during blasting. Predicting the proportion of fines
generated from crushing and the overall breakage process is important to practitioners
interested in the modelling of the complete size distribution of fragments in blasting.

Fines can have a negative or positive impact on the efficiency of downstream


processes. For example, the generation of excessive fines in operations adopting in
situ leaching as their main ore processing method, may hinder recovery as certain
fines tend to affect the permeability of leaching pads. Leaching performance may be
affected if the proportion of material that is less than 150µm exceeds 12 percent in the
feed to the agglomerators (Scott et al. 1998). Similarly, the efficiency of coal
processing is strongly related to the generation of fines of less than 0.5 mm. Increased
fines content in run of mine (ROM) feed leads to higher handling and processing
costs, low yields, increased product moisture content, and in many cases a reduced
product value (Djordjevic et al. 1998). There is also evidence (e.g., Grundstrom et al.
2001) to suggest that by providing an appropriate size distribution to crushing and
grinding circuits, a measurable increased throughput and/or reduced power draw can

2
be obtained. This may entail a requirement to increase the proportion of finer material
in production blasting.

The need to be able to predict the amount of fines from blasting has driven the
development of this new engineering model. The model developed in this study is
based on a comprehensive experimental program conducted between 1996 and 1999
as part of a collaborative research project between the Middle East Technical
University and the BARUTSAN explosives company in Ankara, Turkey (Bilgin et al.
1999). One of the principal aims of this research project was to investigate the
influence of the explosive, rock and blast design parameters on the efficiency of
blasting. Tests were not limited to measuring the extent of crushing but also included
measurements of breakage angle, breakage width, size distribution, throw, back-
break, face velocity, minimum response time and peak particle velocity. In this study,
the data set consisted of 92 model scale blasting tests.

Model scale blasting experiments using synthetic materials (e.g. Plexiglass, Homolite
100, Epon 815 epoxy and concrete) have been used to understand the mechanisms of
rock breakage by explosives and have provided very useful insights into the blasting
process. This includes work conducted by Rustan & Vutukuri (1983), Dick et al.
(1993), Fourney (1993), Aimone-Martin et al. (1998) and Raina et al. (2000). Studies
by Dick et al. (1993) and Stimpson (1970) suggest that cement-based materials such
as concrete can be used to simulate rocks. Applications largely arise from the
cheapness, ease of fabrication and reproducibility of samples (Stimpson 1970).

Work by Stagg et al. (1992) demonstrated that the proportion of fines generated by
blasting can be considered to be scale independent. This means that the proportion of
fine material generated in a small scale blast can indicate the tendency to generate
fines in full production scale blasting (Scott et al. 1998). However, model scale blast
tests conducted in this study do not take into account the influence of rock mass
discontinuities and multiple hole blasts in the generation and liberation of fines.

There are only a handful of models that deal explicitly with the prediction of fines
generated in blasting, these include the approaches proposed by Djordjevic (1999) and
Kanchibotla et al. (1999). Both of these approaches are mechanistic in nature as they
combine some predictions of the physical response of the rock mass (i.e.
crushing/plastic deformation) to the explosive detonation and link this to observed
fragmentation analysis conducted in a limited number of large scale production blasts
and model scale blast chamber tests.

The model proposed here is also mechanistic in nature and follows the concept of
relating the volume of crushed material around a blasthole to the very fine material
present in a muckpile (i.e. less than 1.18 mm). This model differs from previously
documented methods in that it provides a new and improved engineering model to
predict the extent of crushing around a blasthole and through direct sieve analysis it
demonstrates and provides an empirical relationship between the volume of crushing
and the proportion of fines generated during blasting.

3
2 EXPERIMENTAL WORK

2.1 Sample preparation

As shown in Figure 1, concrete blocks were rectangular in shape and measured 1.5 m
in length, 1.0 m in width and 1.0 m in height.

Figure 1. Concrete block samples (Bilgin et al. 1999).

Three concrete mix designs were prepared to obtain low, medium and high strength
concrete types. Table 1 summarises the components used in 1 m3 of concrete for each
mixture. Sica FF was used in order to increase the workability of high strength
concrete. Mix designs were adjusted after determining the moisture content of the
aggregates.

Table 1. Material quantity for 1 m3 of concrete (Bilgin et al. 1999).

Material Amount, kg
Low strength Medium High
concrete strength strength
concrete concrete
Cement 200 425 500
Water 126 192 95
0/3 mm Aggregate 1393 807 897
5/15 mm Aggregate 587 859 956
Sica FF - - 10

Specially prepared mixes were poured into a steel mould arrangement which included
a greased cylindrical steel pipe of a specified diameter. The steel pipe was placed at
the centre of the mould and fixed at the desired burden distance from the front side of
the mould and later removed to create the blasthole. All concrete blocks were left to
cure for at least 28 days before testing.

4
Cylindrical concrete samples of 15x30 cm in size were tested to obtain physical and
mechanical properties, including unit weight, unconfined compressive strength,
splitting tensile strength, P and S-wave velocity, all in accordance with ASTM
Standards (ASTM 1992). In addition, non-destructive tests such as Rebound Hardness
and Ultrasonic Pulse Velocity were carried out. A summary of the range of measured
physical and mechanical properties is given in Table 2.

Table 2. Physical and mechanical properties of concretes (Bilgin et al. 1999).

Concrete R σc (MPa) T ρ Vp Vs Ed νd
(MPa) (kg/m3) (m/s) (m/s) (GPa)
Low Min 15.9 6.7 0.3 2255 3372 1871 20.2 0.278
strength Max 25.1 10.5 0.8 2271 3752 2064 24.8 0.283
concrete
Medium Min 29.6 16.3 1.2 2286 3935 2157 27.3 0.285
strength Max 44.7 24.6 2.9 2379 4553 2471 37.5 0.291
concrete
High Min 39.5 42.1 2.2 2340 4341 2363 33.7 0.290
strength Max 52.9 56.5 4.3 2456 4891 2642 44.4 0.294
concrete
R: Hammer rebound; σc: Uniaxial compressive strength; T: Splitting tensile strength; ρ: Density; Vp:
P- wave velocity; Vs: S- wave velocity; Ed: Dynamic Young’s modulus; νd: Poisson’s ratio.

2.2 Explosive properties

Properties of the commercial explosives used in the experimental work are


summarised in Table 3.

Table 3. Properties of explosives (Bilgin et al. 1999).

Gelatin Dynamite Elbar 1 Dynamite


Density, g/cm3 1.5 1.0
Heat of reaction, MJ/kg 4.70 3.76
Ideal VOD, m/s 7527 5070
Ideal Pressure, GPa 23.71 8.29
Unconfined VOD of 16 mm 1292 1985
charge, m/s

Because charge lengths were small in these particular tests, the confined velocity of
detonation (VOD) was not measured directly. However, independent unconfined and
confined VOD tests were conducted for all the commercial explosives (dynamites and
ANFO type explosives) used in the research project. These measurements were used
to determine the confined VOD for each test by adopting the following relationship
developed by Esen (2001):

ω
β ϕ  Ed 
Dconfined = αq D
n unconfined
  (1)
 1 +ν d 

5
where α, β, ϕ and ω are constants; Dconfined is the confined VOD (m/s); qn is the heat
of reaction for non-ideal detonation (MJ/kg); Dunconfined is the unconfined VOD of an
explosive at a given charge diameter (m/s); Ed is the dynamic Young’s Modulus (GPa)
and νd is the dynamic Poisson’s ratio.

2.3 Test parameters and data collection procedures

During the tests, factors such as confinement, explosive type, specific charge, burden
distance, blasthole diameter and decoupling ratio were varied one at a time. A
summary of the range of parameters used in the experimental work is given in Table
4.

Table 4. Blast design parameters for fully coupled and decoupled model scale tests.

Parameter Fully Coupled Tests Decoupled Tests


Explosive Gelatin dynamite, Elbar 1 Elbar 1 dynamite
dynamite
Decoupling Ratio* 1 1.25, 1.50, 1.75, 2.00
Blasthole Diameter 16-20 20, 24, 28, 32
Burden, cm 22.7-46.2 18.2-31.3
Hole Depth, cm 40.4-45.4 39.8-45.0
Specific Charge, kg/m3 0.110-0.250 0.150-0.175
Explosive Amount, g 8.0-22.8 7.8-16.1
Stemming Material 1.18-3 mm aggregate 1.18-3 mm aggregate
Stemming Length, cm 26.5-40.3 21.0-39.6
Stemming Length/Burden 0.67-1.47 0.69-2.18
Burden/ Blasthole Diameter 14.2-28.9 6.5-15.4
Initiation System Electric detonator Electric detonator
Confined VOD**, m/s 1901-2600 -
Borehole Pressure**, GPa 1.002-1.469 0.470-0.940
* Decoupling ratio = borehole diameter / charge diameter where charge diameter is 16 mm for
decoupled blast tests.
** Detonation velocity and borehole pressure are computed by the non-ideal detonation model
developed by Esen (2001).

Each test blast was instrumented with a triaxial arrangement of high frequency
geophones and monitored with the use of a high-speed video camera. After each test,
the overall fragment size distribution, the extent of crushing, breakage angle, breakage
width, throw, backbreak, face velocity, minimum response time and peak particle
velocity were measured. Details of the monitoring procedures have been previously
discussed by Bilgin et al. (1999).

Fragmentation assessment was conducted using the following methodology:


Fragments larger than 75mm were measured using steel tape in their three axes and
weighed on site. The rest of the broken material (less than 75mm) was taken to the
laboratory for sieve analysis. The sieve sizes used were 1.18mm, 4.75mm, 9.5mm,
19mm, 25mm, 37.5mm, 50mm and 75mm. A sample sieve analysis is shown in
Figure 2.

6
100

10
% Passing

0.1

0.01
1 10 100 1000
Size (mm)

Figure 2. A sample size distribution curve for a typical model scale test.

Of interest to this particular study was the proportion of fragments less than the 1.18
mm size fraction. This particular size fraction is consistent with the 1mm cut off point
discussed by Comeau (1991) and assumed by Kanchibotla et al. (1999).

3 DATA ANALYSIS AND MODEL DEVELOPMENT

In this paper, the analysis focuses on the development of a model that would enable
the prediction of the proportion of fines (material less than 1.18mm) generated during
blasting.

The mechanistic approach combines two specific developments. First is the


development of a relationship between the measured proportion of fines (percent less
than 1.18mm) and the volume of crushing around a blasthole. Secondly is the
development of a model to predict the size of the crushing zone around a blasthole
(Esen et al. 2002).

The concept of the proposed model is similar to that adopted by Kanchibotla et al.
(1999). It is assumed that the proportion of material less than 1.18mm will mainly be
generated by the crushing around a blasthole. The main difference between this
method and others is that the relationship between fines and volume of crushing is
explicitly defined from model scale blasts and it is not based on full scale run of mine
fragmentation calibration. ROM fragmentation has generally gone through many
other forms of degradation which are very difficult if not impossible to quantify.

It is recognised that from a fundamental point of view, fines will also come from other
sources such as the creation of new surfaces through breakage and fragmentation,
liberation of fines within pre-existing discontinuities, particle collisions etc. However,
the contribution to the overall concentration of fines prior to loading and hauling is
very difficult if not impossible to quantify. These factors are therefore ignored in this

7
particular approach as we attempt to predict only the proportion of fines that would be
generated during the crushing stage of the blasting process.

3.1 Correlation between generated fines and the crushing zone volume

From experimental observations, the shape of the crushing zone generally follows a
pear shape form and can approach a cylindrical shape depending on the length of
charge, explosive type and material properties. A cylindrical approach is commonly
used and this shape is adopted here (Figure 3). Given this assumption, the volume of
the crushing zone (Vc) can be simply estimated by,

Vc = V f − Vb (2)

where Vf is the volume of the cylinder of crushed material for the maximum radius of
crushing (rc) and charge length and Vb is the blasthole volume.

Blasthole radius (ro)

Maximum crushing zone


radius (rc)
Explosive
column (L)

Figure 3. Approximation of crushing zone volume.

Given the above approximation, a relationship between the measured proportion of


material less than 1.18mm and the volumetric proportion of crushed material has been
developed (Figure 4).

8
1

0.9 % less than 1.18mm = 0.79 (%Crushed) + 0.026


R2 = 0.909
0.8
% less than 1.18mm

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
0.001 0.01 0.1 1 10
% Crushed

Figure 4. Relationship between % crushed and measured proportion of fines.

In Figure 4, the percent of crushed material is calculated by,

mc
%Crushed = × 100 (3)
M

where mc is the mass of crushed material and M the total mass of broken material. mc
is calculated by,

mc = Vc ρ (4)

where Vc is the volume of crushing around the blasthole and ρ is the material density.

A good correlation exists between the measured proportion of material less than
1.18mm and the volume of the crushed zone. It is important to highlight that a certain
contribution of finer particles could come from the breakage and fragmentation
process itself, however due to the experimental conditions (i.e. model scale,
homogeneous environment) this contribution was considered negligible. The
following equation allows the estimation of the proportion of the fines contribution
from the blasting process only:

% less than 1.18mm = 0.79 (%Crushed) + 0.026 (5)

The above relationship appears to validate findings by Kanchibotla et al (1999) and


Comeau (1991) in describing the 1mm cut off point as a critical value observed in
production blasting. In general, literature acknowledges the contribution of other
sources to the final ROM fragmentation being fed to processing equipment
(Djordjevic et al. 1998, Scott et al. 1998). Therefore, the authors of this paper

9
recognise that in order to predict the total proportion of fines (i.e. material less than
1.18 mm) being fed downstream, a site specific adjustment factor should be
incorporated. This factor would be a function of the following:

• Rock mass characteristics such as degree of fracturing, infilling type and


weathering characteristics,
• The creation of new fractures during the blast fragmentation process,
• Properties of rock material after the blast, that is, degree of pre-conditioning
making the rock material more susceptible to further degradation,
• Impact of digging/loading equipment on the further generation of fines,
• Specifically to underground blasting, the impact of attrition before material
reaches the drawpoint and through ore passes prior to feeding primary crushers.

The above is an extensive list of factors that can only be assessed on a site by site
basis. Back-analysis of ROM fragmentation data would allow for the development of
empirically derived site specific degradation factors.

As discussed earlier, one of the key components of the proposed approach is the
ability to predict the size of the crushing zone around a blasthole for a given explosive
and rock type. This model is discussed in the following section.

3.2 Modelling the size of the crushing zone around a blasthole

As discussed by Whittaker et al. (1992), the compressive strength and stiffness


characteristics of the rock material play a major role in the development of the zone of
crushing or zone of plastic deformation. As the process of crushing is dynamic, a
crushing zone model should include dynamic rock material properties. However,
many of these dynamic properties cannot be directly measured and are not readily
available, resulting in the use of assumed multipliers. For example, Szuladzinski
(1993) assumes the dynamic confined compressive strength to be 8 times the value of
unconfined compressive strength. Mohanty & Prasad (2001) give an insight into the
dynamic strength characteristics of rock materials at strain rates similar to those
experienced during blasting. Using the Split Hopkinson Bar they suggest that the ratio
of the unconfined dynamic strength to static values of unconfined compressive
strength ranged between 2.5 and 4.6 for twelve rock types.

As indicated earlier, the approach adopted in this study was to include both static and
dynamic properties that are readily available. Hence the extent or radius of crushing
denoted as rc shown in Figure 5 was assumed to be a function of explosive type,
material properties and borehole diameter.

10
rc = f (ro,Pb,K,σc)

rc
ro

Figure 5. Parameters influencing the extent of crushing.

As shown by Figure 5, ro is the original borehole radius (mm), Pb is the borehole


pressure (Pa) calculated using non-ideal detonation theory, K is the rock stiffness (Pa)
and σc is the uniaxial compressive strength (Pa). Rock stiffness K is defined assuming
that the material within the crushing zone is homogeneous and isotropic and is given
by,

Ed
K= (6)
1 + υd

where Ed is the dynamic Young’s modulus and νd is the dynamic Poisson’s ratio. If
the dynamic Young’s modulus Ed is not available, it may be estimated from
knowledge of the static value by the following relationship (Eissa & Kazi, 1988),

log10 Est = 0.02 + 0.77 log10 (γEd ) (7)

where Est is the static Young’s modulus (GPa) and γ is the density (g/cm3).

Borehole pressure is computed from the non-ideal detonation model developed by


Esen (2001).

Given the above measurements and calculations and by applying dimensional


analysis, two dimensionless indices (π1 and π2) were derived:

ro
π1 = (8)
rc
and

π2 =
(Pb )
3
or crushing zone index (CZI) (9)
(K )× σ c2

11
where as noted earlier, rc is the crushing zone radius (mm), ro is the borehole radius
(mm), Pb is the borehole pressure (Pa), K is the rock stiffness (Pa) and σc is the
uniaxial compressive strength (Pa).

The relationship between the two indices given above is shown in Figure 6. The
function obtained by non-linear regression is given by:

ro
= 1.231(CZI )
− 0.219
(10)
rc

where CZI is defined as the crushing zone index. This is a dimensionless index that
identifies the crushing potential of a charged blasthole. The correlation coefficient of
the relationship given by Equation 10 is R2=0.83. The crushing zone index (CZI)
appears to capture the dynamic process taking place in the crushing zone by taking
into account both explosive and rock properties.

1
Fully coupled Decoupled
0.9

0.8

0.7

0.6
ro/rc

0.5

0.4

0.3

0.2

0.1

0
0 500 1000 1500 2000 2500 3000
CZI

Figure 6. Relationship between CZI and ro/rc for 92 model scale test blasts.

Because it is physically impossible for the ratio between ro and rc to be greater than 1,
the relationship (Equation 10) is constrained to 1 for very small values of CZI, in this
case for values of CZI of less than 2.6. These small values of CZI will generally
correspond to small borehole pressures (i.e. decoupled charges). The 92 data points
shown in Figure 6 clearly cover a wide range of conditions for which ratios of ro/rc
can be obtained.

The new proposed approach is compared against predictions given by a selection of


models found in the open literature, namely Szuladzinski's (1993), Kanchibotla et al.
(1999), Djordjevic (1999), Il'yushin (1971) and Vovk et al. (1973). This comparison
is conducted for the results obtained in the model scale blasting experiments.
Measured versus predicted values are shown in Figure 7. The new model developed in
this study predicts the size of the crushing zone with reasonable accuracy. However,
other models could not approximate the conditions of the experimental work.

12
Szuladzinski's (1993) predictions are closer to the new model than the approaches
proposed by , Kanchibotla et al. (1999), Djordjevic (1999), Il'yushin (1971) and Vovk
et al. (1973). One of the possible reasons is that these models assume ideal detonation
which is not valid for the model scale tests. Explosives show a more pronounced non-
ideal detonation behaviour under these conditions.

Further details of the proposed model are discussed by Esen et al (2002). Their
analysis has indicated that this new model follows the expected trends; for example
for a specific rock environment and explosive type, as the blasthole diameter
increases the crushing zone radius increases. Similarly, the model shows that an
explosive with the capacity to generate higher borehole pressures has the potential to
increase crushing for the same blasthole diameter and rock environment.

500
Predicted crushing zone radius (mm)

450
400
350
300
250
200
150
100
50
0
0 5 10 15 20 25 30 35 40 45 50
Measured crushing zone radius (mm)
New model Szuladzinski (1993)
Kanchibotla et al. (1999) Djordjevic (1999)
Il'yushin (1971) and Vovk et al. (1973) Perfect fit

Figure 7. Comparison of other models against model scale blasting experiments.

4 APPLICATION OF THE PROPOSED MODEL

The proposed model can be used to predict the likely proportion of fine material
generated during the blasting process. This method allows engineers to analyse
different drilling and blasting strategies to better match explosives to blasting
domains, this is particularly useful where domains are highly susceptible to crushing
and continual degradation (e.g., low strength rocks).

To demonstrate the application of the proposed approach, a number of simulations


were conducted for the geotechnical conditions described in Table 5, these
simulations are based on actual case studies (Esen et al. 2000, Thornton et al. 2001).
The analysis included, three different explosive types, namely ANFO1 (density=0.8

13
g/cm3), ANFO2 (density=0.928 g/cm3) and Emulsion (density=1.25 g/cm3). Results
are summarised in Table 6.

Table 5. Physical and mechanical properties of rock types used in the simulations.

Rock σc, MPa T, ρ, Ed, GPa νd


MPa kg/m3
Coal (Thornton et al. 2001) 20.0 2.0 1440 7.61 0.144
Limestone (Esen et al., 2000) 99.0 8.0 2707 71.48 0.325

Table 6. Relative comparison of % fines generated during blasting.

Blasting Explosive Hole Borehole Crushed Charge Burden x Total Mass %


Domain diameter, Pressure radius length, Spacing x Mass, crushed, fines
mm Pb, r c, m Bench height, kg kg -1.18mm
GPa mm m

Coal Emulsion 150 7.189 812 5.4 7x7x9.5 670320 15970 1.91
Coal ANFO1 150 2.514 408 5.4 7x7x9.5 670320 3929 0.50
Limestone ANFO1 89 1.617 57 6.7 2.6x3.6x7.5 190031 72 0.06
Limestone ANFO2 89 2.308 72 6.7 2.6x3.6x7.5 190031 183 0.10

The above analysis clearly shows the influence of explosive type and rock properties
on the amount fines likely to be generated by blasting. The proportion of fines is
given by a cut off point of 1.18 mm.

In general, the proposed model follows the expected trends, for example for a given
rock type and borehole diameter, an explosive with the capacity to generate higher
borehole pressures, has the potential to form a larger zone of crushing and produce a
higher proportion of fines.

As shown in Table 6, the proportion of fines generated during coal blasting can be
significantly affected by the type of explosive used. The crushing radius is decreased
with the use of a low density explosive (ANFO1) by almost one half which translates
to a reduction in the mass of crushed material per hole of approximately four times. In
addition, the percent of fines (less than 1.18 mm) is decreased by a factor of 3.8. This
type of simulations give a good insight into the coal loss problem during the blasting
process.

The above analysis indicating a proportion of fines (-1.18mm) of 1.91% excludes the
well documented degradation occurring during both excavation and handling
(Djordjevic et al. 1998, Thornton et al. 2001). Sieved measurements discussed by
Thornton et al. (2001) indicate that the proportion of fines (-1mm) is approximately
17%, combining both blasting, excavation and handling processes. This suggests that
coal is susceptible to significant downstream degradation after blasting.

A study conducted by Esen et al. (2000) in a limestone quarry, indicated that by the
introduction of a higher density ANFO product (i.e. ANFO2), the overall
fragmentation was significantly improved, based on comparisons of the coarser

14
fragments of muckpiles (i.e. 50% and 80% passing sizes). However, based on the
simulations (Table 6) the generation of fines (-1.18mm) produced during blasting
would have been increased by a factor of 1.7, which translates to a loss of 5.6 tonnes
of material per blast, this is assuming a blast consisting of 50 blastholes and excludes
the influence of degradation from excavation and handling processes.

5 CONCLUSIONS

An engineering approach to predict the proportion of fines generated during blasting


has been presented. The model is based on the back-analysis of a comprehensive
experimental program that included the fragmentation assessment and direct
measurement of the zone of crushing from 92 blasting tests on concrete blocks using
two commercial explosives.

The proposed model follows the expected trends, for example for a given rock type
and borehole diameter, an explosive with the capacity to generate higher borehole
pressures, has the potential to form a larger zone of crushing and produce a higher
proportion of fines.

Simulations conducted with this model have confirmed the susceptibility of low
strength rock masses (e.g., coal) to further degradation after blasting. In addition, in
quarry operations it was clear that blast optimisation to improve fragmentation must
consider the whole size distribution for a particular application.

A significant contribution to the proposed approach has been the development of a


new model to predict the radius of crushing around a blasthole. This particular
development was required as a number of previously proposed models could not
approximate the conditions measured in the experimental work and there were noted
discrepancies between the reviewed approaches, particularly in smaller diameter holes
and low strength rock conditions. This new model has confirmed that the ratio
between the crushing zone radius and borehole radius (rc/ro) is a function of
explosive, rock properties and blasthole diameter.

This engineering approach provides a good insight into the source of fines from the
blasting (explosive/rock interaction) process, enabling the further development of
fragmentation optimisation tools. The proposed approach is currently being
incorporated in the further development of a mechanistic fragmentation model for
open pit and underground blasting applications.

ACKNOWLEDGEMENTS

The authors would like to acknowledge the support of the BARUTSAN Explosives
Company, Turkey. Many thanks go to Muharrem Kilic, Nedim Yesil and
BARUTSAN Co. staff and technicians. The authors would also like to thank Prof. Bill
Whiten, Dr. R. Trueman and Dr. G. Chitombo of the Julius Kruttschnitt Mineral
Research Centre (JKMRC) for their suggestions.

15
REFERENCES

Aimone-Martin, C.T., Dick, R.D., Weaver, T.A. & Edwards, C.L. 1998. Small Scale
cratering experiments I: concrete. FRAGBLAST – International Journal of
Blasting and Fragmentation 2: 143-180.

ASTM. 1992. Annual Book of ASTM Standards. Vol 04.02.

Bilgin, H.A., Esen, S. & Kilic, M. 1999. Patarge Project, Internal Report, Barutsan
A.S., Elmadag, Ankara, Turkey. in Turkish.

Comeau, W. 1991. General Report: Blasting Technology. Proceedings of the


International Congress on Rock Mechanics, Aachen, Germany: 1605-1609.

Dick, R.D., Fourney, W.L., Wang, X.J. & Young, C. 1993. Results from instrumented
small scale tests. Proceedings of the Fourth International Symposium on Rock
Fragmentation by Blasting-Fragblast-4, Vienna, Austria: 47-54.

Djordjevic, N., Esterle, J., Thornton, D. & La Rosa, D. 1998. A new approach for
prediction of blast induced coal fragmentation. Mine to Mill 1998 Conference.
The Australasian Institute of Mining and Metallurgy, Brisbane, Australia: 175-
181.

Djordjevic, N. 1999. Two-component of blast fragmentation. Fragblast 1999, South


African Institute of Mining and Metallurgy, Johannesburg, South Africa: 213-
219.

Eissa, E.A. & Kazi, A. 1988. Relation between static and dynamic Young's Moduli of
rocks. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 25(6): 479-482.

Esen, S., Bilgin, H.A. & BoBo, T. 2000. Effect of explosive on fragmentation. The
4th Drilling and Blasting Symposium, Ankara, Turkey: 63-72. In Turkish.

Esen, S. 2001. Modelling non-ideal detonation behaviour of commercial explosives.


Internal Report, JKMRC, Australia.

Esen, S., Onederra, I. & Bilgin, H. 2002. Modelling the size of the crushing zone
around a blasthole. Paper submitted to the Int. J. Rock Mech. Min. Sci.
Geomech. Abstr. under review.

Fourney, W.L. 1993. Mechanisms of rock fragmentation by blasting. Comprehensive


Rock Engineering Principles, Practice and Projects. Pergamon Press, Oxford.
Vol. 4, p. 39-69.

Grundstrom, C, Kanchibotla, S.S., Jankovich, A. & Thornton, D. 2001. Blast


fragmentation for maximising the sag mill throughput at Porgera Gold Mine.
Proceedings of the Twenty-Seventh Annual Conference on Explosives and
Blasting Technique, ISEE, Orlando, Florida, USA: 383-399.

16
Hustrulid, W. 1999. Blasting principles for open pit blasting. A.A. Balkema,
Rotterdam. Vol. II, Chapter 21, p. 964-1009.

II'yushin, A.A. 1971. The mechanics of a continuous medium. Izd-vo MGU. Moscow.
In Russian. Translated in Hustrulid (1999).

Kanchibotla, S.S., Valery, W. & Morrell, S. 1999. Modelling fines in blast


fragmentation and its impact on crushing and grinding. Explo’99 – A conference
on rock breaking. The Australasian Institute of Mining and Metallurgy,
Kalgoorlie, Australia: 137-144.

Mohanty, B. & Prasad, U. 2001. Degree of rock fragmentation under high strain rates.
Proceedings of the Twenty-Seventh Annual Conference on Explosives and
Blasting Technique, ISEE, Orlando, Florida, USA: 89-95.

Raina, A.K., Chakraborty, A.K., Ramulu, M. & Jethwa, J.L. 2000. Rock mass damage
from underground blasting, a literature review, and lab- and full scale tests to
estimate crack depth by ultrasonic method. FRAGBLAST – International
Journal of Blasting and Fragmentation 4: 103-125.

Rustan, A. & Vutukuri V.S. 1983. The influence from specific charge, geometric
scale and physical properties of homogeneous rock on fragmentation. First
International Symposium on Rock Fragmentation by Blasting, Lulea, Sweden:
115-142.

Scott, A., David, D., Alvarez, O. & Veloso, L. 1998. Managing fines generation in the
blasting and crushing operations at Cerro Colorado Mine. Mine to Mill 1998
Conference. The Australasian Institute of Mining and Metallurgy, Brisbane,
Australia: 141-148.

Stagg, M.S., Otterness, R.E. & Siskind, D.E. 1992. Effects of blasting practices on
fragmentation, Proceedings of the 33rd US Symposium on Rock Mechanics,
Santa Fe, NM, USA: 313-322,

Stimpson, B. 1970. Modelling materials for engineering rock mechanics. Int. J. Rock
Mech. Min. Sci. Geomech. Abstr. 7: 77-121.

Szuladzinski, G. 1993. Response of rock medium to explosive borehole pressure.


Proceedings of the Fourth International Symposium on Rock Fragmentation by
Blasting-Fragblast-4, Vienna, Austria: 17-23.

Thornton, D., Kanchibotla, S. & Esterle, J. 2001. A fragmentation model to estimate


ROM distribution of soft rock types. Proceedings of the Twenty-Seventh Annual
Conference on Explosives and Blasting Technique, ISEE, Orlando, Florida,
USA: 41-53.

Vovk, A., Mikhalyuk, A. & Belinskii, I. 1973. Development of fracture zones in rocks
during camouflet blasting. Soviet Mining Science 9(4): 383-387.

17
Whittaker, B.N., Singh, R.N. & Sun, G. 1992. Rock fracture mechanics principles,
design and applications, Amsterdam: Elsevier. Chapter 13, p. 444-445.

18

You might also like