You are on page 1of 198

Blast Design and Assessment

for Surface Mines and Quarries


Areas of Study: Mining
Qualifies for CMS
Qualifies for Certification
This is a practical course that provides a review of blasting theory and blasting products, and
emphasizes the design, assessment, optimization and safety of blasting practices for open cast
mining and quarrying. Topics are presented in an applied manner and address the impact of
blasting on mine design and mining efficiency.
Authors: Alan Cameron P.Eng (Author), Bill Forsyth P.Eng (Author), Dr. Tom Kleine P.Eng.
(Author)

Blast Design and Assessment for Surface Mines and Quarries (Text Level)
Part 1: Introduction

Introduction
Neither Edumine nor the authors of this course accept responsibility for the results of
application or usage of theory, practice, technique or reference material presented in this
course. No warrantee is made with respect to the accuracy of information presented.

General

Surface mining and quarrying operations usually require the use of blasting for the preparation
of rock for excavation. As urban areas spread and as the mines become deeper, the constraints
on blasting increases. This has resulted in an increased requirement for well trained and
experienced personnel for the design and implementation of the blasts.
When appropriate precautions and safety procedures are implemented, blasting can be safe.
The cost efficiency of blasting is often greater than that of mechanical methods such as ripping
by a dozer or removing rock with hydraulic breakers. Blasts need to be designed and
implemented so as to minimize:

• the risk of damage to nearby structures


• disturbance of people who live and/or work near the blast area

This course describes methods of rock blasting and how they may be safely and cost efficiently
employed. Included are the following topics:

• Types of Explosives
• Initiators and Initiating Systems
• Mechanisms of Rock Fracture by Explosives
• Fundamentals of Production Bench Blasting
• Blasting Techniques for Producing Stable Rock Slopes
• Preventing Damage From Ground Vibration, Air Blast and Flyrock
• Safety Procedures

Also included is a list of references (from which more details of the products and techniques
described in this manual can be obtained), and a glossary of blasting and excavation terms.

Blasting Applications in Surface Mining and Quarrying

The following is a brief description of some of the types of blasts commonly used in surface
mines and quarries:

Production Bench Blasts

This is the most common type of blasting. The goal is to fragment and loosen the rock in
preparation for excavation by front end loaders, shovels, draglines or dozers. The amount of
preparation or conditioning of the rock, to be done by the blasting, depends on the rock mass
characteristics as well as the type, size and mode of operation of the excavation equipment.

Figure 1. Photograph Of Production Bench Blast


Wall Control Blasts
These blasts are designed to break the rock near or up to the final pit or quarry limit while
causing minimal damage to the rock beyond this limit. A number of techniques are used to
achieve this including: line drilling (?), presplitting (?), trim blasting (?), cushion blasting (?) and
buffer blasting (?).

Figure 2. Photograph of presplit wall.


Throw/Cast Blasts

In many surface coal mines there is an economic incentive to move a significant portion of the
overburden material to its final position by blasting. The throw or cast blasts are designed in
such a way as to maximize the horizontal movement of the rock in the desired direction.

Figure 3. Photograph of throw / cast blast.


Sinking Cut Blasts

These blasts are usually the first blast in a new bench where a vertical free face is not available.
They are designed to break and loosen the rock volume using the surface as the only free face.
A ramp is dug through the muckpile to the floor of the new bench and a free face established.

Non-Explosive Methods of Rock Breakage

There are some circumstances where non-blasting methods of rock breakage should be
employed. For example, in locations where there is nearby equipment or structures which are
extremely sensitive to vibrations, or in areas of soft rock where it is more economical to rip than
to blast. The following is a brief description of some non-explosive methods of rock breakage.

Ripping

Weak rock can often be broken by ripping using an excavator or, in harder rock, a bulldozer. The
most commonly used criterion for determining whether ripping is possible, and for selecting the
appropriate equipment, is the seismic velocity of the rock. The Caterpillar Equipment Handbook
provides information on seismic velocities and the required ripping equipment.

The advantage of ripping over blasting in weak rock is that ripping can be carried out to close
tolerances, and there is virtually no damage to the rock in the sides of the excavation. Ripping
can also be done after dusk.

Hydraulic Splitting

For small excavations in brittle rock, an hydraulically powered splitter or wedge can be used to
create a tensile fracture between closely spaced drillholes. This is a slow but highly controlled
method of rock breakage. It is most often used for breaking concrete and in dimensional stone
quarries, or heavily loaded foundation areas where overbreak from blasting would reduce the
bearing capacity of the rock.

Hydraulic Breakers

Hydraulically powered breakers, often mounted on excavator boom, can be used to break both
strong and weak rock. This method of rock breakage can be carefully controlled, and is
sometimes used in conjunction with ripping where lenses of harder rock are encountered.
Disadvantages of this equipment are that it is noisy, and not effective in very hard bedrock with
few cracks or discontinuities.

Chemicals

Another product on the market is a cement-like compound that expands when mixed with
water. The expansion pressure is sufficient to break rock when the material is placed in a row of
closely spaced holes. While this is a highly controlled method of breaking rock, it has the
disadvantages that the material is costly and the expansion process takes 10 to 20 hours.

EXPLOSIVE / ROCK INTERACTION

Background

The interaction of the explosive and the surrounding rock mass during and immediately after
detonation is a function of the detonation (?) properties of the explosive and the dynamic
physical properties of the adjacent rock mass. The theories of rock breakage and the
mechanisms of muckpile formation are based on the interaction of the detonating explosive and
the surrounding rock. An understanding of the mechanism of rock breakage by explosives
enables the blast designer to fragment the rock mass economically, while minimizing the
damage caused by the blast beyond the excavation perimeter.
The mechanisms by which rock is fractured by explosives are fundamental to the design of
blasting patterns. They also relate to the damage that can be suffered by surrounding rock and
structures and to the reactions of people living in the vicinity of a blast.

Theories of Rock Breakage

There are many theories and models that attempt to describe the process that occurs during
and after the detonation of an explosive charge in a rock mass. In general terms this process
involves the rapid release of energy by the explosive, the application of the energy to the rock
and the subsequent response of the rock to the application of the energy. It is complicated by
such things as the rate, type and amount of energy released by the explosive, the design of the
blast and the properties of the rock mass.

The mechanisms of rock breakage that have been identified (Hagan (1967), Hagan (1973) and
Mercer (1980)) are:

• crushing
• relative radial motion
• release of load
• spalling
• gas extension of strain wave-generated and/or natural cracks
• flexural rupture
• shear fracturing along natural and strain wave generated cracks
• in-flight collisions

These mechanisms cause varying amounts of breakage depending on the characteristics of the
explosive, rock properties and geometry of the rock mass and explosive charge. They can be split
into two categories, those caused by the shock component of the energy from the explosive and
those resulting from the gas energy (Brown (1956)).

Shock Energy Breakage Mechanisms

When an explosive is detonated, it is converted within a few thousandths of a second into high-
temperature gases. When confined in a blasthole, this very rapid reaction causes pressures that
usually exceed about 1800 GPa (approximately 260 kips/in2) to be exerted against the blasthole
wall. This energy is transmitted into the surrounding rock mass in the form of a compressive
strain wave, or shock wave, which travels at a velocity of 2000 - 6000 metres per second (6500
- 20000 ft/sec) through the rock.

The rock breakage mechanisms that can be attributed to the shock component of the energy
released by an explosive are crushing, relative radial motion, release of load and spalling.
Crushing occurs around a blasthole wall when the pressure in the detonation front exceeds the
dynamic compressive strength of the rock (Hagan (1973) and Bauer (1978)). The out-going strain
pulse generated by the high pressure detonation front disperses and loses energy rapidly.
Crushing ceases when the strain level in the pulse drops below the elastic limit of the rock. This
is usually very close to the blasthole wall.
The rock that forms the wall of the blasthole outside the crushed zone is subjected to very
sudden compression due to the dispersing strain pulse as illustrated in Figure 1. This
compression (i.e. relative radial motion) results in tangential stresses which can cause cracks to
develop radially from the blasthole (see Figure 1). The radial cracks initially develop in all
directions from the blasthole wall and are not influenced by a local free face.

Figure 1. First stages of explosive / rock interaction showing expanding strain wave (after
Mercer (1980))

The mechanism of fracturing caused by release of load occurs immediately after the strain or
compression pulse passes, resulting in a local decrease in density with subsequent tensile
stresses. These tensile stresses produce fractures aligned perpendicularly to the direction of
travel of the strain pulse (Clay et al. (1965)).

Spalling occurs when a compression or strain pulse is reflected by a free surface. At this point
two waves are generated, a tensile wave and a shear wave. The tensile wave may cause cracking
and the rock to spall in the region of the free surface. Both the tensile and shear waves may
extend pre-existing or new (i.e. formed by the initial out-going strain pulse) cracks.

Gas or Heave Energy Breakage Mechanisms

The mechanisms of fracturing described above are caused by the initial strain or compression
pulse from the detonating explosive charge. A zone of very high pressure and temperature gases
occupies the blasthole behind the detonation front. These gases penetrate the crushed zone
around the blasthole and flow into the radial or naturally occurring cracks. The gas pressure
tends to wedge open the cracks and cause them to extend.

The pressure on the blasthole walls caused by the explosive generated gases and the stress field
due to the pressurised cracks displaces the rock mass between the blasthole and the free face.
Because of the geometry of the explosive charge and the rock mass (see Figure 2) the rock at
the face bends causing fracturing by flexural rupture (Mercer (1980)). This has been observed in
high-speed films and still photographs of experimental and production blasts. An example is
shown in Figure 3.
Figure 2. Later stages of explosive/rock interaction showing gas penetration
and burden movement (after Mercer (1980)).

Figure 3. Photograph of a free face immediately after detonation of a long


cylindrical explosive charge.

Shear fracturing occurs when adjoining rock is displaced at different times or at different rates.
The displacement is caused by the high pressure gases.

Some fracturing occurs when rock particles which are in motion collide. The amount of fracturing
resulting from this mechanism depends on the geometry of the explosive charges, their order
and relative time of initiation plus the physical properties of the rock.

The previous four mechanisms of rock fracturing are a result of the high pressure and
temperature gases acting on the rock mass. This gas energy also plays an important role in the
displacement of the muckpile.

Mechanics of Muckpile Displacement

The movement of material in a blast is primarily a result of the high pressure gases produced by
the detonation of explosive charges (Edl (1983)). These high pressure gases flow into the cracks
surrounding the blasthole forming a hydrostatically stressed region. The shape of this region
depends on the geometry of the explosive charge and the blasthole, though in most mining
applications the explosive charge is in the form of a long cylinder creating a cylindrically shaped
hydrostatically stressed region. The high pressure gases in this region exert a force in all
directions with material movement occurring in the direction of least resistance.

The amount of displacement of material in a blast is a function of the physical properties of the
material, blasthole orientation, burden distance, spacing between blastholes, sequence and
relative time of initiation of charges, amount and distribution of the explosive and the properties
of the gas generated by the detonating explosive. These factors all influence the length of time
the high pressure gases are contained within the rock mass and therefore determine the amount
of work performed by them (Lownds (1986) and Brinkmann (1990)). When these gases vent to
the atmosphere they stop doing useful work.

Influence of Blast Design on Explosive/Rock Interaction

The fragmentation achieved by the process is highly dependent upon the degree of confinement
and coupling (?) of charges within the blastholes, the amount of burden (?) and the sequencing
of the blast. If confinement of the charges by stemming is inadequate, some energy will be lost
from the blastholes. Inadequate coupling results in poor transmission of the strain wave to the
rock mass. Excessive burdens result in choking and poor movement of the rock, whereas
inadequate burden results in a waste of explosive energy and excessive throw of blasted rock.
Effective delaying of individual blastholes ensures maximum development and utilization of free
surfaces by reducing effective burden, provides freedom for the rock to move toward the free
face, and reduces the extent of damage to surrounding rock.

ROCK PROPERTIES

Introduction

Blasting results are greatly influenced by properties of the intact rock and the rock mass.
Fragmentation, displacement and blasting damage can, in some instances, be more influenced
by the rock than by the blast design.

Intact Rock Properties

The primary intact rock properties commonly applied in blast design are:

• Elastic constants (Young's Modulus, E and Poisson's Ratio, ν)


• Strength (uniaxial compressive σc and tensile, σt)
• Density (ρ)
• P-Wave Velocity (Vp)

Definitions
• Young's Modulus is the ratio of axial stress to axial strain in uniaxial compression
(typically expressed in GPa).
• Poisson's Ratio is the ratio of lateral to axial strain magnitudes (dimensionless).
• Compressive strength is the ratio of peak load to the cross-sectional area of a test
sample in uniaxial compression (typically expressed in MPa).
• Tensile strength is the peak load per cross section area at the point of rupture in tension
(typically expressed in kPa).
• The density of a rock is the specific weight (typically expressed in g/cc).
• The P-wave velocity of a rock is a measure of the compressive wave transmission
velocity (typically expressed in m/s).

Rock Mass Properties

The primary rock mass properties commonly applied in blast design are:

• the number of fractures (density);


• the orientation of fracturing (absolute and relative to the free face).

Application of Rock Properties

Rock and rock mass properties can be used to assess the following:

• blastability
• fragmentation
• blast damage

Blastability Indices

Rock and rock mass properties can be used in the development of blastability indices. A popular
method was developed by Lilly (1986) for use in the northwest iron ore range in Australia. This
Blastability Index uses the following variables in the calculation of blasting energy requirements:

• rock mass description (fracture density);


• joint plane spacing and orientation;
• specific gravity; and,
• hardness.

Other indices have been developed that use similar data.

Fragmentation Prediction

The Kuz-Ram fragmentation (?) equations were developed by Cunningham (1983), Cunningham
(1987) to estimate fragmentation from a blast with given geologic and design variables. The Kuz-
Ram calculation uses a "Rock Factor" that is based on the blastability calculation described by
Lilly (1986).

This method is widely accepted and easily adaptable to calculation within a spreadsheet.
Blast Damage

The determination of strain (vibration) based damage criteria uses the relationship between
induced strain and peak vibration as shown in the following equation:

where:

ε = induced strain in the rock

PPV = peak particle velocity at the point of interest (mm/s)

Vp = P-wave velocity of rock (mm/s)

From Hooke's law and assuming a brittle failure mode for rock, the maximum particle velocity
can be calculated.

where:

PPVmax = maximum particle velocity before tensile failure (mm/s)

σt = uniaxial tensile strength of rock (Pa) - 1/10 to 1/15 the UCS

Vp = P-wave velocity of rock (mm/s)

E = Young's Modulus of the rock (Pa)

Many other damage predictions methods have been published that use essentially the same
information to assess damage potential from blasting.

Summary

It is important to remember that blasting is rarely undertaken in a homogeneous, isotropic rock


mass. In terms of achieving successful blasting, the properties of the intact rock and the rock
mass may be as important as the selection of an explosive and blasting pattern.

Examples of the influence of geology on blasting results is shown in the following figures:
Figure 1 - In-dipping structure controlling bench face angle.

Figure 2 - Horizontally bedded rock allows near vertical faces and the
opportunity for successful wall control blasting.

Figure 3 - Blocky ground makes wall control blasting difficult and


fragmentation highly variable.
Figure 4 - Complex, highly fractured rock mass, fragmentation controlled by
small in-situ block size.

Figure 5 - Poor quality rock mass impacting wall control.


Figure 6 - High quality rock mass allows for excellent wall control and blasting
control of fragmentation.

Figure 7 - Overbreak associated with dominant geologic structure.


Figure 8 - Blocky rock mass with in-dipping structure. Fragmentation and
bench face angle controlled by the rock mass.

Glossary of Blasting and Excavation Terms

Acoustical Impedance (?)- The mathematical expression for characterizing a material as to its
energy transfer properties (the product of its unit density and its sound velocity (pV)).

Adit (?) - A nearly horizontal passage from the surface by which an underground mine is entered,
as opposed to a tunnel.

Air Deck (?)- A blasting technique wherein a charge is suspended in a borehole, and the hole
tightly stemmed so as to allow a time lapse between detonation and ultimate failure of the rock
(no coupling realized).

ANFO (?)- Ammonium Nitrate - Fuel Oil Mixture. Used as a blasting agent.

Astrolite- A family of two-component explosives, usually liquid, with variable detonating


velocities.

Back (?)- The roof or top of an underground opening. Also, used to specify the ore between a
level and the surface. or that between two levels

Back Break (?)- Rock broken beyond the limits of the last row of holes.

Bench- The horizontal ledge in a face along which holes are drilled vertically. Benching is the
process of excavating whereby terraces or ledges are worked in a stepped shape.

Blast (?)- The operation of rending (breaking) rock by means of explosives. Shot is also used to
mean blast.

Blasting Agent (?) - Any material or mixture, consisting of a fuel and oxidizer, intended for
blasting, not otherwise classified as an explosive and in which none of the ingredients are
classified as an explosive, provided that the finished product, as mixed and packaged for use or
shipment, cannot be detonated by means of a No. 8 test blasting cap when unconfined.

Blast Hole - A hole drilled in rock or other material for the placement of explosives.

Block Hole (?) - A hole drilled into a boulder to allow the placement of a small charge to break
the boulder.

Booster (?) - A chemical compound used for intensifying an explosive reaction. A booster
does not contain an initiating device but must be cap sensitive.

Bootleg (?) - A situation in which the blast fails to cause total failure of the rock due to
insufficient explosives for the amount of burden, or caused by incomplete detonation of the
explosives. That portion of a borehole that remains relatively intact after having been charged
with explosive and fired.

Bridging - Where the continuity of a column of explosives in a borehole is broken, either


by improper placement, as in the case of slurries or poured blasting agents, or where some
foreign matter has plugged the hole.

Bulk Strength (?) - Refers to the strength of an explosive in relation to the same volume
of a standard explosive, usually ANFO.

Burden (?) - generally considered the distance from an explosive charge to the nearest free
or open face. Technically, there may be an apparent burden and a true burden, the latter being
measured always in the direction in which displacement of broken rock will occur following firing
of an explosive charge.

Centers - The distance measured between two or more adjacent blast holes without reference
to hole locations as to row. The term has no association with the blast hole burdens.

Chambering - More commonly termed Springing (?). The process of enlarging a portion of a
blast hole (usually the bottom) by firing a series of small explosive charges.

Collar - The mouth or opening of a borehole, drill steep or shaft. Also, to collar in drilling means
the act of starting a borehole.
Condenser - Discharge (?) - A blasting machine which uses batteries to energize a series of
condensers, whose stored energy is released into a blasting circuit.

Connecting Wire - Any wire used in a blasting circuit to extend the length of a leg wire (?)
or leading wire (?).

Connector - Refers to a device used to initiate a delay in a Primacord circuit, connecting one
hole in the circuit with another, or one row of holes to other rows of holes.

Coupling - The act of connecting or joining two or more distinct parts. In blasting the
reference concerns the transfer of energy from an explosive reaction into the surrounding rock
and is considered perfect when there are no losses due to absorption or cushioning.

Coyote Blasting (?) - The practice of drilling blast holes (tunnels), horizontally into a rock
face at the foot of the shot. Used where it is impractical to drill vertically.

Critical Diameter (?) - Minimum diameter of an explosive charge at which detonation will still
take place.

Cushion Blasting (?) - The technique of firing of a single row of holes along a neat excavation
line to shear the web between the closely drilled holes. Fired after production shooting has been
accomplished.

Cut - More strictly it is that portion of an excavation with more or less specific depth and
width, and continued in like manner along or through the extreme limits of the excavation. A
series of cuts are taken before complete removal of the excavated material is accomplished. The
specific dimensions of any cut is closely related to the material's properties and required
Production levels.

Cut-Off - Where a portion of a column of explosives has failed to detonate due to bridging, or to
a shifting of the rock formation due to an improper delay system.

Deck (?) - In blasting a smaller charge or portion of a blast hole loaded with explosives
that is separated from the main charge by stemming or air cushion.
Deflagration (?) - An explosive reaction that consists of a burning action at a high rate of speed
along which occur gaseous formation and pressure expansion.

Delay Element (?) - That portion of a blasting cap which causes a delay between the instant
of impressment of electrical energy on the cap and the time of detonation of the base charge of
the cap.

Detonating Cord (?) - A plastic covered core of high velocity explosives used to detonate
charges of explosives in boreholes and under water, e.g. Primacord.

Detonation (?) - An explosive reaction that consists of the propagation of a shock wave through
the explosive accompanied by a chemical reaction that furnishes energy to sustain the
shockwave propagation in a stable manner, with gaseous formation and pressure expansion
following shortly thereafter.

Detonation Pressure (?)- The pressure within the primary reaction zone bounded by the shock
front and the C-J Plane.

Dip (?) - The angle at which strata, beds, or veins are inclined from the horizontal.

Drop Ball (?) - Known also as a Headache Ball. An iron or steel weight held on a wire rope that
is dropped from a height onto large boulders for the purpose of breaking them into smaller
fragments.

Explosion - A thermochemical process whereby mixtures of gases, solids, or liquids react


with the almost instantaneous formation of gaseous pressures and near sudden heat release.
There must always be a source of ignition and the proper temperature limit reached to initiate
the reaction. Technically, a boiler can rupture but cannot explode.

Explosive - Any chemical mixture that reacts at high speed to liberate gas and heat and
thus cause tremendous pressures. The distinctions between High and Low Explosives are
twofold; the former are designed to detonate and contain at least one high explosive ingredient;
the latter always deflagrate and contain no ingredients which by themselves can be exploded.
Both High and Low Explosives can be initiated by a single No.8 blasting cap as opposed to
Blasting Agents which cannot be so initiated.

Face (?) - The end of an excavation toward which work is progressing or that which was last done.
It is also any rock surface exposed to air.
Fire - In blasting it is the act of initiating an explosive reaction.

Floor (?) - The bottom horizontal, or nearly so, part of an excavation upon which haulage
or walking is done.

Fragmentation (?) - The extent to which rock is broken into small pieces by Primary blasting.

Fracture - Literally, the breaking of rock without movement of the broken Pieces.

Fuel - In explosive calculations it is the chemical compound used for purposes of combining with
oxygen to form gaseous products and cause a release of heat.

Galvanic Action - Currents caused when dissimilar metals contact each other, or through a
conductive medium. This action may create sufficient voltage to cause premature firing of an
electric blasting circuit, particularly in the presence of salt water.

Galvanometer - A device containing a silver chloride cell which is used to measure resistance in
an electric blasting circuit.

Grade - In excavation, it specifies the elevation of a roadbed, rail, foundation, etc. When given a
value such as percent or degree grade it is the amount of fall or inclination compared to a unit
horizontal distance for a ditch, road, etc. To grade means to level ground irregularities to a
prescribed level.

Gram Atom (?) - The unit used in chemistry to express the atomic weight of an element in terms
of grams (weight).

Hardpan (?) - Boulder clay, or layers of gravel found usually a few feet below the surface and so
cemented together that it must be blasted or ripped in order to excavate.

Highwall (?) - The bench, bluff, or ledge on the edge of a surface excavation and most usually
used only in coal strip mining.
Initiation (?) - The act of detonating a high explosive by means of a mechanical device or other
means.

Joints - Planes within rock masses along which there is no resistance to separation and along
which there has been no relative movement of the material on each side of the break. They
occur in sets, the planes of which are generally mutually perpendicular. Joints, like stratification,
are often called partings.

Lead Wire (?) - The wires connecting the electrodes of an electric blasting machine with the final
leg wires of a blasting circuit.

LEDC - Low Energy Detonating Cord. Used to initiate non - electric caps at the bottom of
boreholes.

Leg Wires (?) - Wires, leading from the top end of an electric blasting cap; used to couple caps
into the circuit.

Mat (?) - Used to cover a shot to hold down flying material; usually made of woven wire cable,
tires or conveyor belt.

Millisecond Delay Caps - Delay electric caps which have a built - in delay element, usually
25/1000th of a second apart, consecutively. This timing may vary from manufacturer to
manufacturer.

Misfire (?) - A charge, or part of a charge, which for any reason has failed to fire as planned. All
misfires are to be considered extremely dangerous until the cause of the misfire has been
determined.

Mole (?) - A unit in chemical technology equal to the molecular weight of a substance expressed
in grams weight).

Muckpile - The pile of broken material or dirt in excavating that is to be loaded for removal.

Mud Cap - Referred to also as Adobe or Plaster Shot (?). A charge of explosive fired in contact
with the surface of a rock after being covered with a quantity of mud, wet earth, or similar
substance, no borehole being used.
Open Pit - A surface operation for the mining of metallic ores, coal, clay, etc.

Overbreak (?) - Excessive breakage of rock beyond the desired excavation limit.

Overburden (?) - The material lying on top of the rock to be shot; usually refers to dirt and gravel,
but can mean another type of rock; e.g. shale over limestone.

Oxidizer (?) - A supplier of oxygen.

Permissible (?) - Explosives having been approved by the U.S. Bureau of Mines for non - toxic
fumes, and allowed in underground work.

Powder (?) - Any of various solid explosives.

Powder Factor (?) - Term used to describe the mass of explosive used to break a unit volume or
weight of rock.

Premature (?) - A charge which detonates before it is intended to.

Presplitting (?) - Stress relief involving a single row of holes, drilled along a neat excavation line,
where detonation of explosives in the hole causes shearing of the web of rock between the
holes. Presplit holes are fired in advance of the production holes.

Primary Blast - The main blast executed to sustain production.

Primer (?) - An explosive unit containing a suitable firing device that is used for the initiation of
an entire explosive charge.

Quarry (?) - An open or surface mine used for the extraction of rock such as limestone, slate,
building stone, etc.

Riprap (?) - Coarse sized rocks used for river bank, dam, etc., stabilization to reduce erosion by
water flow.
Round (?) - A group or set of blast holes constituting a complete cut in underground headings,
tunnels, etc.

Seam (?) - A stratum or bed of mineral. Also, a stratification plane in a sedimentary rock deposit.

Secondary Blasting (?) - Using explosives to break up larger masses of rock resulting from the
primary blasts, the rocks of which are generally too large for easy handling.

Seismograph (?) - An instrument that measures and supplies a permanent record of earthborne
vibrations induced by earthquakes, blasting, etc.

Sensitizer (?) - The ingredient used in explosive compounds to promote greater ease in initiation
or propagation of the reactions.

Shot Firer (?) - Also referred to as the Shooter or Blaster. The person who actually fires a blast.
A Powderman, on the other hand, may charge or load blast holes with explosives but may not
fire the blast.

Shunt (?) - A piece of metal connecting two ends of leg wires to prevent stray currents from
causing accidental detonation of the cap. The act of deliberately shorting any portion of an
electrical blasting circuit.

Slope - Used to define the ratio of the vertical rise or height to horizontal distances in describing
the angle a bank or bench face makes with the horizontal. For example, a 1 - 1/2 to 1 slope
means there would be a 1 - 1/2 ft rise to each 1 ft or horizontal distance.

Snake Hole (?) - A hole drilled or bored under a rock or tree stump for the placement of
explosives.

Spacing - In blasting, the distance between boreholes or charges in a row

Stemming (?) - The inert material, such as drill cuttings, used in the co ar Portion (or elsewhere)
of a blast hole so as to confine the gaseous products formed on explosion. Also, the length of
blast hole left uncharged.
Strength - Refers to the energy content of an explosive in relation to an equal amount of
nitroglycerine dynamite.

Stratification - Planes within sedimentary rock deposits formed by interruptions in the


deposition of sediments.

Strike (?) - The course or bearing of the outcrop of an inclined bed or geologic structure on a
level surface.

Sub - Drill (?) - To drill blast holes beyond the planned grade lines or below floor level.

Swell Factor - The ratio of the volume of a material in its solid state to that when broken.

Tamping (?) - The process of compressing the stemming or explosive in a blast hole.

Toe (?) - The burden or distance between the bottom of a borehole to the vertical free face of a
bench in an excavation.

Velocity of Detonation (?) - The measure of the rate at which the detonation wave travels
through an explosive.

Review #1
The randomly selected multiple-choice questions below are designed to
review your understanding of the material covered in the preceding sessions.
Your selections are lost when you leave the review page. On return the
review will start afresh with a new selection of questions.

This review is currently set to practise mode. To optimize your learning


experience you need to register for certification before entering the course.
Certification tests more rigorously, keeps track of your answers to the
multiple choice review questions, and enables you to report and submit your
review scores to complete the certification process. If you have already
registered and been approved for certification then you should Exit and re-
enter before proceeding.
Each question below has one or more correct responses. Your selection of a
response is immediately marked correct or not.

Q1. The properties of intact rock that are relevant to blast design include ...

elastic constants?

compressive strength?

density or specific gravity?

fragmentation?

tensile strength?

P-wave velocity?

fracture density?

Q2. The objectives of blast design for surface mining and quarrying include
...

fragmentation and loosening of rock in


preparation for excavation?

minimizing the damage to rock beyond excavation


limits?

minimizing the risk of damage to nearby


structures?

minimizing the disturbance of people who live


and/or work near the blast area?

Q3. Which of the following rock breakage mechanisms are attributed to the
shock component of the energy released by an explosive?

shear fracturing along cracks?


flexural rupture?

relative radial motion?

crushing?

release of load?

gas pressure extension of cracks?

spalling?

Q4. The term "coyote blasting" refers to ...

a method of pest control practised in the western


US states?

the practice of drilling blast holes horizontally into


the foot of a rock face?

the process of enlarging the bottom of a blast hole


by firing a series of small explosive charges?

Q5. Prediction of blast damage to structures in the vicinity of a blast is


based on the induced strain in the rock, which is determined from ...

peak particle velocity at the point of interest?

the P-wave velocity of rock?

compressive strength of rock?

tensile strength of rock?

Young's modulus of rock?


Blast Design and Assessment for Surface Mines and Quarries (Text Level)
Part 2: Explosives and Charging Systems

INTRODUCTION
Neither Edumine nor the authors of this course accept responsibility for the results of
application or usage of theory, practice, technique or reference material presented in this
course. No warrantee is made with respect to the accuracy of information presented.

Introduction

The effectiveness of explosives employed in mine and quarry blasting is due to the magnitude
of their shock pulse as well as the peak gas pressures and the rates at which these pressures are
developed.

Some of the energy released in a blasthole:

• can be wastefully dissipated in crushing or plastically deforming weak/soft rocks, and/or


• tends to escape through any path of low resistance.

Each blasthole should be primed, charged and stemmed so that explosion gases are confined
for a reasonable period of time. The gases should break, displace and loosen the burden rock
satisfactorily without creating excessive throw, overbreak, ground vibrations or air vibrations.

Ideally, every explosive should have a low cost and a high energy yield per unit weight coupled
with:

• insensitivity to initiation by friction, mechanical impact and fire


• totally reliable sensitivity to initiation by the detonator or primer for which the explosive
has been designed
• unlimited resistance to water and low temperatures
• oxygen balance and, hence, minimal yield of poisonous explosion gases
• excellent handling characteristics
• unlimited shelf life

Common Terms Relating to Explosives

• detonation - an explosive decomposition or combustion reaction that moves through a


charge at greater than the speed of sound to produce shock waves and significant
overpressure;
• deflagration - rapid burning with a sudden evolution of flame and vapor;
• ideal explosive - a molecular explosive such as TNT (Tri-nitro-toluene), NG (Nitro-
glycerine) or PETN (Pentaerythritol tetranitrate), where the fuel and oxidizer are
contained in the molecule;
• high explosive - an explosive which can be detonated by a standard #8 detonator in
unconfined conditions;
• blasting agent - an explosive material that meets prescribed criteria for insensitivity to
initiation; a material or mixture consisting of fuel and oxidizer used in blasting, but not
otherwise defined as a high explosive;
• sympathetic detonation - where charges exhibit a relatively high level of initiation
sensitivity (exemplified by dynamites), the distance between blastholes is small and/or
the rock is an effective transmitter of strain waves, the shock created by an earlier-firing
charge can initiate an adjacent charge prematurely. Sympathetic detonation is
encouraged by the presence of groundwater, by clay seams and by structural features
such as open joints. Fortunately, the on-going replacement of dynamites by
compositions having lower initiation sensitivities (e.g. emulsion explosives) has greatly
reduced the probability of sympathetic detonation.

Materials Used in Explosives

• ammonium nitrate (AN) - NH4NO3 - classified as an oxidizing agent and supplies oxygen
to the detonation reaction; it is the primary ingredient in most commercial explosives
used for mine blasting;
• fuels (e.g. diesel) - supplies fuel to the detonation reaction;
• oxidizers (e.g. ammonium nitrate (AN), sodium nitrate (SN), calcium nitrate (CN)) -
compounds supplying oxygen to the detonation reaction;
• nitroglycerine (NG) - C3H5N3O9 - ideal explosive commonly used in dynamites;
• trinitrotoluene (TNT) - C7H5N3O6 - ideal explosive used in primers and as a sensitizer
for cartridge explosives;
• pentaerythritol tetranitrate (PETN) - C5H8N4O12 - ideal explosive used in primers and
detonating cord.

Categories of Explosives

Modern commercial explosives can be grouped into two principal categories:

• high explosives (i.e. detonator-sensitive explosives), and


• blasting agents (i.e. detonator-insensitive explosives).

During the last 25 years, the high degree of replacement of nitroglycerine (NG) -sensitized
explosives (e.g. dynamites) by ANFO, watergels, and more recently, emulsion-type explosives,
has been accompanied by a marked decrease in the sensitivity of charges to impact, friction and
fire. Consequently, the potential for accidental detonations has been reduced and, as a result,
explosives are now manufactured, transported, stored and used more safely. Unfortunately,
knowledge of the greater potential safety of the newer explosives has led to a general reduction
in the care and respect with which these explosives are handled and used. Whilst this attitude
(which is a natural one) detracts from the potential gain in safety, the resultant gain is still very
appreciable.

Development of Bulk Explosives

Early 1950's: ANFO = AN + Organic Fuel


• no water resistance

Mid 1950's: Watergels

Developed to increase water resistance and density of bulk explosive.

• primarily a mixture of oxidizers, fuels and water


• used TNT and aluminium as fuels and sensitizers
• contained approximately 20% water

Late 1970's: Emulsions

developed to provide a lower cost and more efficient explosive compared to Watergel
explosives

1980's: Emulsions Blends

• heavy ANFO's - ANFO with up to 30% emulsion


• used to increase energy concentration and water resistance compared to ANFO
• doped emulsion - emulsion explosive with up to 40% ANFO

1980'S: Low Density ANFO's

• special applications where reduced energy concentration is required

1990's: Water Resistant ANFO

• ANFO with geling agent to provide some water resistance

1990's: High Density ANFO's

• ANFO made with more dense AN prills to increase energy concentration in the blasthole

Properties of Explosives

Density

• units are g/cc


• relates weight of explosive to volume
• controls energy concentration in the blasthole

Sensitivity

Minimum condition required to start detonation of the explosive.


Ease of initiating the explosive shown by:

• minimum booster weight


• gap sensitivity
• impact test
• pressure tolerance
• critical diameter

Water Resistance

Estimate of how the detonation reaction of an explosive is affected by water in the blasthole.

Usually expressed in qualitative terms (good, fair, poor, etc.) or as maximum sleeping time in a
blasthole.

Chemical Stability

Length of time an explosive can sleep in a blasthole without change.

Watergels can have gel structure breakdown resulting in increased density, segregation and
reduced water resistance.

Figure 1. Bulk explosive that has crystallized.


Emulsions / Heavy ANFO's may experience crystalization of the emulsion phase resulting in loss
of water resistance (Figure 1 right).

Performance

Performance of explosives is covered in the next session (see Detonation Properties of


Explosives).

Detonation Properties of Explosives


Detonation Properties

Detonation properties of explosives are a function of the detonation reaction and the
subsequent energy released. They can be related to the ability of the explosive to break and/or
move a rock mass.

Detonation properties include:

Figure 1. Illustration of explosive detonation (Atlas Powder Company (1987)).


• available energy
• relative effective energy
• weight and bulk strengths
• gas yield
• velocity of detonation
• detonation pressure
• borehole pressure
• shock/heave energy split

Some of these properties are related to the interaction between a detonating explosive charge
and the surrounding medium. In this way they can be related to the mechanisms of rock
fragmentation and muckpile formation and also to the techniques developed for assessing the
performance of explosives.

Available Energy

The available energy (Ao) of an explosive is the energy released during detonation. The
calculation of Ao is based on the thermodynamic calculation of the detonation reaction (Meyer
(1987)). This includes the chemical breakdown of the initial components of the explosive and
the products after the detonation reaction. The heats of formation of the ingredients and the
products are compared and then a proportion of the thermal energy retained in the solid
products of detonation is subtracted to produce the value for available energy. The proportion
of thermal energy in the solid products of detonation which is assumed to be lost varies between
100% and 33%, depending on the assumptions applied during the analysis. The assumptions that
are usually made in this analysis are that:

• The detonation reaction is ideal and goes to completion.


• The gases from the detonation reaction do useful work until the pressure drops to one
atmosphere and their temperature reaches approximately 25°C (77°F).

The calculated available energy of a bulk explosive overstates the energy that is available to do
useful work in blasting rock because:

The detonation reaction that occurs in most commercial explosives is not ideal and does not go
to completion. This is due to the size of the ingredients being relatively large and therefore there
is not a truly homogeneous mix of the components.

The gases generated by the detonation reaction do work on the rock mass in creating fractures
and movement until they vent to atmosphere. The pressure at which the venting occurs is much
higher than one atmosphere but it varies depending on such things as the geometry of the blast,
properties of the rock mass, time between initiation of adjacent blastholes, the volume of gas
and the rate at which the gas is generated by the explosive (Lownds (1986) and Brinkmann
(1990)).

Available energy is usually expressed in units of Megajoules per kilogram (MJ/kg) (1 MJ/kg =
108.34 kcal/lb). This term does not give any indication of the rate of energy release or the
partition between shock and heave components of the energy.

The available energy of an explosive formulation is usually calculated using a thermo-


hydrodynamic detonation code or ideal detonation model such as Fortran-BKW, Ruby and Tiger
(Finger et al (1976)) and IDeX (Sheahan and Minchinton (1988)). These codes are very complex
and different codes will produce different estimates of the available energy for the same
explosive formulation because they are based on different assumptions. Often the available
energy is presented in a relative format compared with a standard explosive, usually ANFO, and
referred to as weight strength. ANFO is given a value of 100.

Relative Weight Strength and Relative Bulk Strength

The strength of an explosive is a measure of its ability to do work in the form of fragmenting and
displacing a rock mass. The calculated available energy (Ao) per unit mass of an explosive
compared with that of ANFO is the most commonly used definition of weight strength (?). This
term is a comparative one: ANFO is the common standard for comparing bulk commercial
explosives and is arbitrarily given a value of 100. Weight strength expresses the ability to do
work with respect to a unit weight of the explosive. This term does not indicate how the available
energy will be distributed or that the difference between calculated available energy and the
actual energy will vary depending on the departure of the explosive from theoretical behaviour.

Bulk strength (?) is the energy of the explosive compared with an equal volume of ANFO. It is
linked to the weight strength of the explosive through density as shown in the following formula:
Bulk strength reflects the amount of energy loaded into one metre of blasthole and is an
important parameter when evaluating the distribution of explosive energy in a blast because
both the density and weight strength of the explosive significantly influence this distribution.
This term has the same deficiencies as weight strength because it is based on calculated values
for the available energies of the different explosive products and may have little relationship to
the energy actually generated.

Weight and bulk strength parameters were improved by selecting a cut-off or venting pressure
for the energy calculation (to more accurately reflect actual blasting conditions) and this was
called the Effective Energy of the explosive.

Relative Effective Energy

The effective energy of an explosive is the energy released during detonation when the gases
expand and the pressure decreases from the detonation pressure to some cut-off or limit
pressure. The cut-off pressure normally used is 100 MPa (14500 psi) which represents the point
at which the detonation gases vent, that is, cease to be confined and therefore cannot do any
more effective work on the surrounding rock (Harries, (1985) and Sheahan and Minchinton,
(1988)). Available energy (Ao) from expansion of the gases produced by detonation is summed
incrementally with time until the last increment of pressure reaches one atmosphere. The
effective energy is the cumulative energy released to the cut-off pressure of 100 MPa (14500
psi) (see Figure 2).

Figure 2. Cumulative energy - pressure function (Harries (1985)).


Relative effective energy is the effective energy of the explosive compared with the effective
energy of ANFO and is usually expressed as a percentage. Relative effective energy does not
indicate how the energy will partition between the shock and heave components described
below (see Sections Detonation Pressure (Shock Energy) and Borehole Pressure (Heave Energy).
Because of this, an explosive product should not be selected for a specific task based on this
parameter alone. The effective energy of ANFO can vary depending on the type of ammonium
nitrate prill used, the size distribution of the prill, the density of the ANFO and on the method of
calculating the energy. It would be best if all of these were specified when relative effective
energies for different explosive products are quoted.
Figure 1. Illustration of explosive detonation (Atlas Powder Company
(1987)).

The pressure at which the detonation gases vent to the atmosphere in a blast is a function of
the characteristics of the rock mass, the type of blasthole initiation and the blast design. It is
almost impossible to measure the pressure within a blasthole to verify this or verify the
commonly used cut-off pressure of 100 MPa (14500 psi). A change in the cut-off pressure will
change the ranking of explosive products (Cunningham and Sarracino (1990)) as shown in Figure
2. This is a drawback to the use of this term for comparison of the performance of different
explosive products.

Gas Yield

The gas yield is the volume of gas produced by the detonation of one kilogram of explosive. The
temperature and pressure of the gas is assumed to be 25°C (77°F) and 1 atmosphere. In some
instances the gas yield is given as moles per kilogram of explosive which can be converted to
litres per kilogram by multiplying by 22.4 (i.e. volume of 1 mole of gas). The gases that occur in
the detonation products are determined through the evaluation of the detonation reaction and
analysis of samples of post-detonation fumes. These are usually performed by explosives
manufacturers.

The gas yield provides a rough indication of gas energy but so far it has not been directly related
to the performance of the explosive in a specific rock mass.

Velocity of Detonation

The velocity of detonation (?) (VOD) is the speed with which the detonation front travels through
the explosive charge. This speed is influenced by the density, charge diameter, particle size of
the explosive and degree of confinement of the charge. The VOD controls the rate at which
energy from the explosive is released and to a large extent, the ratio of the shock to bubble
energy components of the total energy. An explosive with a relatively low VOD releases its
energy at a slower rate and a larger proportion of the total energy usually is in the form of gas
pressure. In the converse situation a high VOD explosive is considered to have a high shock
energy or brisance. The higher the detonation velocity, the greater the ability of the explosive
to break rock. The strain to heave energy ratio increases with increasing velocity of detonation.

The measured VOD of an explosive column also is an indication of how completely the explosive
reacts chemically as it detonates. This is accomplished by comparing the measured value with
the theoretical VOD for the explosive charge. According to Bauer and Cook (1961), Cook (1974)
and Bauer et al (1984) the volumetric fraction of the explosive which has reacted in the
detonation is defined by the relationship:

where

N = volumetric fraction of the explosive which has reacted

D = measured VOD [m/s]

D* = theoretical thermohydrodynamic VOD [m/s].

The measured VOD of most commercial explosives will not reach the theoretical
thermohydrodynamic value because they are not normally ideal explosives. The VOD of these
products is influenced mostly by the size and intimacy of mixing of the components of the
explosive as well as by the charge diameter, degree of confinement of the charge, type of
initiation and charge temperature. The approximate range of VOD values for most commercial
explosives is from 3000 m/s (9800 ft/s) for low density ANFO products to greater than 6500 m/s
(21300 ft/s) for some emulsion explosives.

The measured VOD of an explosive in a blasthole provides a check on the quality of the explosive,
detonation pressure (see Section Detonation Pressure (Shock Energy)) and efficiency of the
detonation reaction when the result is compared with results from measurements of high
quality explosive in good blasthole conditions.

Detonation Pressure (Shock Energy)

The detonation pressure (?) (Pd) of an explosive is the pressure in the detonation wave as it
progresses along a charge. It is measured at the Chapman-Jouguet (CJ) plane as shown in Figure
3, which illustrates the detonation of an explosive and the pressure profile in the detonating
charge (after Given (1973)).
Figure 3. Simplified Illustration of detonation reaction (after Given ed. (1973)).
The detonation pressure is primarily responsible for the generation of a shock pulse in the
medium surrounding the explosive charge. The magnitude of this pulse reflects the shock energy
of the explosive and it is the effect of this pulse that causes breakage in the surrounding rock.

The magnitude of the detonation pressure has been found to be a function of the density of the
explosive and the VOD (Brown (1956); Johansson and Persson (1970); Cook (1974); Bjarnholt
(1980); and Atlas Powder Company (1987)). A commonly accepted formula for approximating
the detonation pressure is given by Bjarnholt (1980) as:

where

Pd = detonation pressure [Pa]

K = constant equal to 0.25

D = VOD [m/s]

ρ = explosive density [kg/m3].


Figure 4. Shock energy breakage mechanisms, first stages of explosive/rock
interaction showing expanding strain wave.
The typical range for detonation pressures generated by commercial explosives is 2 GPa (0.29 x
106 psi) to 12 GPa (1.74 x 106 psi). There is a loss of shock energy at the interface between the
detonation front and the wall of the charge as the shock pulse is propagated into the
surrounding material.

Figure 5. Pre-split blast utilizing shock energy.


This energy loss takes the form of heating, crushing and vaporisation of the material.
Quantification of this energy loss is almost impossible due to the very destructive environment
in the regions where it occurs.
The total effect of an explosive on rock includes an energy component due to the expansion of
gases generated by the detonation reaction. This is referred to as the borehole pressure.

Borehole Pressure (Heave Energy)

The borehole pressure (?) (Pb) is the pressure exerted on the medium surrounding the explosive
charge by the expanding gases of the detonation product (Atlas Powder Company (1987)). This
is illustrated in Figure 3. The pressure exerted by these gases and the time interval over which it
is exerted is a measure of the gas energy of the explosive (Lownds (1986) and Sheahan and
Minchinton (1988)). This energy is often referred to as heave or bubble energy.

The amount of pressure in the borehole is a function of the confinement on the charge as well
as the quantity and the temperature of the gas generated. Borehole pressure is often expressed
as a percentage of detonation pressure. This can vary from 30% to 70% for different explosives
and charge conditions but the average is approximately 50% (Atlas Powder Company (1987)).
The ratio between detonation pressure and borehole pressure represents the shock to gas
energy split of the explosive. This is known to vary depending on explosive formulation and
charge condition so there is no constant ratio which can be used for all situations.

Figure 6. Gas or heave energy breakage mechanisms - later stages of


explosive/rock interaction showing gas penetration and burden movement.
Direct measurement of borehole pressure is not possible due to the very high detonation
pressure preceding it which would destroy any instrumentation in the charge before the
borehole pressure develops.

Figure 7. Gas energy producing burden movement.


The borehole pressure in a blast is reflected by burden velocity or throw, swell and gas
penetration and is therefore often called heave or gas energy. Results of the borehole pressure
can be evaluated using crater tests, high-speed photography and muckpile surveys. The total
energy released by the detonation of an explosive is a sum of the shock energy and the heave
energy. The ratios of these energies can vary significantly for the range of types of explosives.

Shock/Heave Energy Split

The split between the shock and heave components of the energy released depends on the
composition, density and velocity of detonation of the explosive plus the physical characteristics
of the surrounding rock mass (Lownds (1986); Sheahan and Minchinton (1988) and Brinkmann
(1990)).

In general the higher the velocity of detonation of an explosive the more the energy split will
favour the shock component as illustrated in Figure 8. The effect that the physical properties of
the rock mass surrounding an explosive charge has on the performance of the explosive has
been proposed by Sheahan and Minchinton (1988) and is shown in Figure 9. Both of the
relationships between the shock and heave energies of an explosive mentioned above (i.e.
Figures 8 and 9) have not been verified by measurements made under mine site conditions.
Figure 8. Schematic illustration of the difference in energy partitioning for
high and low order detonation (Shock energy = area ABCD; Heave energy =
area CDEF) (Brinkmann (1990)).
Figure 9. Borehole expansion (Sheahan and Minchinton (1988)).

Detonator-Sensitive Packaged Explosives

Types of Detonator-Sensitive Explosives

Detonator-sensitive explosives (usually referred to as high explosives (?)) include emulsion


explosives, watergel explosives, dynamites and cast pentolite-type boosters. Each of these can
be reliably initiated by a No. 8 strength detonator or by a strand of 10 g/m (48 gr/ft) detonating
cord. Dynamites and some cast boosters can be reliably initiated by 5 g/m (25 gr/ft) cords.

Until quite recently, operators needing a reliable, water-resistant high explosive found it
necessary to use a NG-based composition (e.g., a dynamite). Today, operators can also select,
with a high degree of confidence, a (detonator-sensitive) emulsion or watergel explosive.
Figure 1. Packaged explosives including boosters, dynamite and emulsions.
Most packaged explosives are detonator-sensitive. The exceptions are some large diameter (i.e.
4+ in) packages of watergel and emulsion explosive.

Emulsion Explosives

A wide range of detonator-sensitive emulsion explosives are now being produced, principally for
the construction, underground metal mining and quarrying industries.

Emulsion explosives are of the water-in-oil type. They consist of microdroplets of super-
saturated oxidizer solution within an oil matrix. So as to maximize energy yield whilst minimizing
both production costs and selling prices, the oxidizer within the microdroplets consists mainly
of ammonium nitrate. Because the number of microdroplets per unit volume of emulsion is
exceedingly large, adjacent microdroplets are forced against each other, thereby deforming
their original spherical shapes. Consequently, the mean thickness of oil between microdroplets
is very small. Each of the ingredients of an emulsion explosive is a non-explosive. Cartridged
emulsions have the consistency of a firm putty, as shown in Figure 1. Because they are very
intimate mixtures, these products are visually homogeneous. The satisfactory viscosities of
emulsions are achieved without the use of gums or cross-linking agents.

Because emulsions are of the water-in-oil type, they exhibit excellent water resistance. Blasthole
water "sees" only the emulsion's matrix (i.e., the oil); it is prevented from contacting and, hence,
dissolving the nitrate-rich microdroplets by the enveloping film of (insoluble) oil. The water
resistance of emulsions is even greater than that of watergels and dynamites.

Whilst some degree of sensitivity results from the high degree of physical intimacy of the
oxidizer and fuel components, air or gas bubbles are the major sensitizer of emulsions. Where
charges could be compressed by static and/or dynamic pressures just prior to detonation (see
Section Pressure Desensitization), the bubbles should take the form of microballoons. (Each
bubble then has a high degree of dimensional permanence.)
Unlike most watergels, the sensitivities of emulsions do not decrease significantly with a fall in
temperature, as each microdroplet of oxidizer solution is isolated from adjacent microdroplets
by the oil film matrix. A decrease in temperature does not allow more salts within each
microdroplet to crystallize out; nor does it allow oxidizer crystals to coalesce or grow.

Most emulsion explosives exhibit relatively low sensitivities to initiation by a 10 g/m (48 gr/ft)
detonating cord downline. Where a charge consists of a continuous column of cartridges,
therefore, it is initiated most efficiently by a detonator located within the bottom cartridge. Only
when charge continuity is prevented by rough and/or caving blasthole walls should the explosive
column be traced with 10 g/m (48 gr/ft) detonating cord. This aspect of low sensitivity may
represent only a temporary limitation of emulsion explosives.

Although they are detonator-sensitive, emulsion explosives are relatively insensitive to


detonation by friction, impact and/or fire.

ANFO consists of AN and diesel oil. Because a standard (i.e., non-aluminized) emulsion consists
of oxidizers, fuels AND WATER, its theoretical weight strength (i.e., its theoretical energy yield
per unit weight) is not as high as that of ANFO. If ANFO's weight strength is taken as 100, the
calculated weight strength of a standard emulsion may be say 80. But this does not indicate that
the blasting effectiveness of this emulsion is only 80% of that of an equal weight of ANFO. The
actual energy yields of emulsion explosives are higher than one would expect. (This is due to the
emulsion's extremely high degree of oxidizer/fuel intimacy - exceeded only in molecular
explosives such as NG, TNT and PETN.) Therefore, the effective relative weight strength of a
standard emulsion is appreciably more than 80. (This beneficial effect of high detonation ideality
is being observed most in underground metal mines, hard-rock quarrying and construction
operations, where small-diameter cartridged emulsions are competing with dynamites and
watergel explosives.) This writer is not able to calculate the increment in relative weight strength
which results from an emulsion's more efficient and complete detonation. However, blasting
results indicate that this increment is significant and, therefore, that the effective relative weight
strength of an emulsion is significantly greater than its theoretical relative weight strength.

The densities of emulsions usually lie in the 1.1 - 1.2 g/cc range. Based upon a calculated relative
weight strength of say 80, an emulsion with a density of 1.15 g/cc has a calculated relative bulk
strength of:

(where 0.84 is the assumed mean density of bulk ANFO). But because the effective relative
weight strength of this emulsion is considered to be appreciably more than 80 (see preceding
paragraph), the effective relative bulk strength is considered to be significantly more than 110.
Even higher bulk strengths can be achieved when fuel-grade aluminium powder is added to an
emulsion.

Emulsions are packaged in a thin tough plastic film. This approach to packaging gives cartridges
which exhibit:

• a very satisfactory degree of rigidity


• resistance to rupturing during normal handling and use
• the ability to rupture and spread when tamped

If maximum possible energy concentration in downholes is required, cartridges should be slit


lengthwise as they are dropped; the emulsion then slumps, displaces more of the air or water,
and fills a greater percentage of the available charge volume. Considerable slumping can be an
advantage in front-row blastholes which have excessive toe burdens. Should running water cross
the blasthole, however, the cartridges should not be slit, but simply dropped, and the period
between charging and firing the blast should be kept to a minimum. For charging rough
blastholes, paper-cartridged emulsions have recently become available; the charging
characteristics of these products are similar to those of dynamites.

Because it has not been possible to detonate emulsion explosives in impact tests, charging with
these explosives considerably reduces hazards associated with careless or accidental drilling
near bootlegs (?) of previously fired blastholes. Under no circumstances, however, should
drilling take place in or near the bootlegs of blastholes previously fired with emulsion, due to
the possible presence of an unfired detonator.

Because of the presence of NG, the handling of dynamites often gives rise to headaches, whilst
vapours in closed magazines have a similar effect. Emulsions are free of ingredients which can
cause headaches. Therefore, the use of these products increases operator comfort and, hence,
efficiency.

Tests prove that emulsions burn with difficulty. This result is not really surprising when one
considers that emulsions usually contain 9 - 12% water. Fire hazards in storage, therefore, are
very much lower than those with dynamites.

Emulsion explosives have been shown to produce concentrations of toxic gases that are much
lower than those generated by dynamites and lower than those produced by watergels.
Emulsions also create less (visible) smoke and, hence, increase safety through improved
visibility. But fumes (i.e., toxic gases) should never be confused with smoke, and there could be
a hazard, unless it is fully recognized, in assuming that the very small amount of smoke after an
emulsion blast is synonymous with lack of fumes and a signal that it is safe to return to the face.
The fumes from emulsions are certainly much better than those from dynamites, but the
atmosphere is not necessarily safe when it is free from smoke.

For the following reasons, emulsion explosives have to a large extent replaced dynamites and
watergel explosives in most operations:

• Cartridges are firm; they are easy to handle, charge and tamp.
• Because of their very high detonation velocities, cartridges have high abilities for
priming ANFO-type mixtures.
• Their high strain wave energy: heave energy ratios make them suited to fragmenting
strong massive rocks.
• Their lack of NG eliminates the possibility of harmful or unpleasant physiological effects
(e.g., headaches) during handling.

As emulsions represent the most recently developed family of commercial explosives,


improvements in their properties and performance can be expected to continue.

Watergel (Slurry) Explosives


Watergel explosives are produced primarily for the construction, quarrying and underground
metal mining industries. These products contain relatively high concentrations of sensitizing
ingredients to ensure reliable detonation in small-diameter blastholes. Compared with
dynamites, they are more difficult to detonate by friction, impact and/or fire.

Watergels are based on saturated aqueous solutions of AN (often with sodium nitrate and/or
calcium nitrate) in which fuels, sensitizers (and sometimes more AN prills (see Blasting Agents))
are dispersed. The liquid phase is thickened with gums and gelled with cross-linking agents to
keep the solids in suspension, provide a satisfactory degree of cohesiveness and maximize water
resistance. When completely gelled, watergels have a rubbery, porridge-like consistency.

Most watergels are sensitized, at least partially, by bubbles of air and/or gas. Some watergels
are also sensitized by explosives of low sensitivity (e.g., monomethylamine nitrate or, for short,
MMAN). Fuels include aluminium powder. By varying the type and quantity of fuel (and
especially aluminium powder), watergels can be tailored to exhibit a wide range of weight
strengths and bulk strengths.

A certain amount of water is necessary to provide the required consistency and texture of a
watergel. If too little water is used, the liquid phase is insufficient and the resulting high viscosity
hinders cartridging of the watergel. To maximize weight strength, however, it is important to
minimize water content. If more water than that required for suitable consistency is used,
weight strength suffers. Apart from its contribution to consistency, the water content reduces
hazard sensitivity associated with fire, friction and impact (and especially projectile impact).

Temperature and density have pronounced effects on the sensitivity of watergels. All watergels
become less sensitive and less fluid at lower temperatures. Above certain densities, watergels
cannot be detonated even in large-diameter blastholes with powerful boosters. The density can
be lowered:

• by aeration
• by injecting a very small amount of 'gassing' solution into the watergel as it is cartridged
• by adding glass microballoons or equivalent to the watergel

In watergels which do not contain microballoons, the air and/or gas bubbles are compressed by
the weight of the overlying column of watergel and/or water and, therefore, charge density
increases towards the bottom of the blasthole. In watergels which contain microballoons, the
air bubbles are still compressed by the weight of the watergel/water column above, but the
microballoons retain their original size and, hence, give the watergel an irreducible minimum
level of sensitivity under limited pressures.

Because of their high bulk strength, high water resistance, good sensitivity and greater safety in
handling and use, watergel explosives have largely replaced dynamites. Where adverse ground
conditions do not encourage lateral precompression and dead-pressing of charges in closely-
spaced blastholes (e.g., in burn cuts), good watergels give satisfactory blasting results.

Watergels have been largely replaced by emulsion explosives. Watergel is the older explosives
technology, emulsion the new technology.

Dynamites
Dynamites (?) are NG-based explosives of high sensitivity. They can be formulated with high
plasticity, high densities and good water resistance, properties which are usually required in wet,
tough blasting conditions.

Figure 2. Loading dynamite.


When fired in a blasthole, gelatin-type dynamites develop moderately high detonation
velocities. These velocities combined with high density and bulk strength, make the gelatins
most effective in wet work where good breakage of strong massive rocks is required.

The future for dynamites depends essentially upon the assimilation of safety with cost efficiency;
whilst dynamites have higher hazard potentials, their performance reliability and energy yields
per dollar are sometimes higher than those for emulsion and watergel explosives.

Cast Pentolite-Type Boosters

Cast pentolite-type boosters (?) (e.g. PENTEX) have been developed to initiate ANFO type,
emulsion and watergel blasting agents. They are made principally of cast pentolite (a mixture of
TNT and PETN). In normal handling and use, they are much less sensitive than dynamites to
shock, friction and impact. As a result of their high density and very high detonation velocity,
these boosters, despite their small size, develop extremely high peak pressures during
detonation. Their exceedingly high detonation pressures explain why these boosters are, on an
equal weight basis, the best available boosters.

The longitudinal holes provided in these boosters permit the simple secure attachment of
detonating cord. Certain types of boosters are not reliably initiated by downlines having core
loads lower than 10 g/m (48 gr/ft); others can be reliably initiated by downlines with core loads
down to about 5 g/m (25 gr/ft). Additional boosters can be added up the explosive column by
simply threading them onto the downline and letting them fall to their intended positions.

The twin holes in some of these boosters allow the use of both detonating cord and a delay
detonator (preferably non-electric) in the same booster for bottom priming with the insurance
of an upline and additional upper later-firing boosters.

Cast boosters are completely waterproof and do not deteriorate during extended storage, even
in the most severe climates. Their performance is unaffected by immersion in properly
formulated blasting agents.

Blasting Agents

Introduction

Blasting agents consist largely of ammonium nitrate (AN). Blasting agents of the ANFO, emulsion
and watergel types now account for over 90% of the commercial explosives used in North
America.

Ammonium Nitrate (AN)

AN is not considered to be an explosive. AN is classified as an oxidizing agent and, therefore, is


subject to certain regulations regarding its transportation. Whilst it is a strong supporter of
combustion, AN is not flammable. It does not burn in the usual manner. AN can make fuels burn
more intensely than they would with air only.

When mixed or contaminated with a small percentage of a combustible material such as diesel
oil, AN acquires the properties of a blasting agent. Although many forms of AN could be used to
form a blasting agent, the porous prilled form is best for ANFO-type compositions.

Figure 1. Porous prilled AN.


The common properties of porous prilled AN used in bulk explosives are:

• bulk density: 0.79 - 0.84 g/cc


• micro-porosity: 10% - 15%
• coated and white in colour
• free-flowing
• prills 1 - 2 mm (0.04 - 0.08 in) diameter
• macro-pores = 35% of volume

ANFO

The annual consumption of ANFO greatly exceeds that of any other type of explosive. And with
good cause, since the cost per tonne for ANFO is lower and the energy yield per tonne is usually
higher than for other explosives. The energy yield per unit cost is always highest for ANFO.

Figure 2. ANFO.
Diesel oil continues to be the best fuel for ANFO. The microvoids within each porous prill absorb
and retain the optimum amount of oil. Achieving a uniform mix of AN and oil is important.
Pouring a measured quantity of oil into a bag of AN before tipping the (inadequately mixed)
contents into a blasthole is not good enough, as an appreciable proportion of explosion energy
is not realized. The relationship between the fuel content of ANFO and its available energy is
shown in Figure 4. This indicates that the optimum fuel content is 5.6% by weight and that a
significant loss in energy occurs if the fuel content varies much from the optimum.
Figure 3. Variation of energy per kilogram of ANFO with increasing fuel oil
content (Mercer, 1983).
Because of its unparalleled energy yield per unit cost, bulk ANFO is generally used in dry
blasthole situations, irrespective of the strength of the rock mass. But the mechanical efficiency
of ANFO charges cannot be optimum for the whole range of rock strengths encountered.
Optimum mechanical efficiencies will be achieved by using:

• regular ANFO in rock masses having a certain strength,


• ANFO containing increasing percentages of an emulsion blasting agent as rock mass
strength increases beyond this value (see Low Density ANFO below), and
• ANFO containing increasing percentages of say expanded polystyrene beads as rock
mass strength decreases below this value (see ANFO/Emulsion Blends below).

But for a given rock mass, the explosive composition of optimum mechanical efficiency should
be used only when this mixture is also the most cost-efficient. Explosives should be selected on
a basis of cost-efficiency rather than mechanical efficiency.

Aluminized ANFO (AlANFO)

The addition of aluminium powder to ANFO results in an appreciable increase in explosion


energy yield per unit weight as shown in Figure 5. But for the following three reasons, only the
first of which is adequately recognized, energy increments per dollar are low.
Figure 4. Variation of energy per kilogram of ALANFO with increasing
aluminium content.
• Al powder is expensive.
• The particle size range of Al needs to be matched to the blasthole diameter. With small-
diameter blastholes, the correspondingly small burden becomes detached from the rock
mass a relatively short time after detonation. If the reaction of each Al particle (by a
surface burning/erosion mechanism) is not complete by the time the explosion gases
start to escape via the broken burden rock and/or ejected stemming material, it will be
quenched; the detonation products will then contain the unreacted cores of Al particles.
Optimum blasting with AlANFO demands that, before the burden moves significantly,
there is sufficient time

- for the Al to react completely to form Al2O3, and

-for the extremely hot Al2O3 (which can exist only in the condensed form) to surrender
a large percentage of its heat to the surrounding lower-temperature gases so that these
can perform work.

• Even where ideal conditions exist, the Al2O3 and explosion gases fail to remain in
thermal equilibrium as the burden rock is broken and displaced. As a result, at least one-
third of the thermal energy bound up within the Al2O3 particles is not transferred to the
gases and, therefore, is lost. This explains, at least partially, why the observed
performances of AlANFOs are significantly lower than their theoretical ones.

These inherent deficiencies of AlANFO could well explain the low overall popularity of such
mixtures. In the last few years, the increasing use of emulsion blasting agents to raise the density
and, hence, bulk strength of ANFO has probably drawn the chequered career of AlANFO to an
end.

ANFO/Emulsion ('Heavy' ANFO) Blends


Since the early 1980s, ANFO has become closely linked with emulsion blasting agents (see
Section Emulsion-Based Blasting Agents). The density and bulk strength of ANFO can be
increased by partially filling ANFO's macropores with an emulsion. At the other end of the
ANFO/emulsion scale, ANFO can be blended into an emulsion matrix wherever the reaction rate
and strain wave energy to heave energy ratio of the charge need to be reduced to increase the
cost-efficiency of blasting (see Section Bulk Emulsion-Based Blasting Agents).

'Heavy' ANFO consists of a mixture of porous prilled AN, oil and an emulsion blasting agent.
Because it is a viscous yet non-sticky liquid, the emulsion is able to replace air in the interstices
between ANFO particles. When the selected ANFO to emulsion ratio is such that much of the
macrovoid within the ANFO component is filled with emulsion, the density of the resulting
'Heavy' ANFO usually lies in the 1.00 - 1.25 g/cc range. If all air within these interstices were to
be replaced, the resulting blasting agent would tend to exhibit an inadequate sensitivity. (The
air-filled microvoids within the ANFO particles provide a poor distribution and, except for highly
porous low-density prills, an insufficient number of initiation centres in a charge.)

'Heavy' ANFOs should be used only in blastholes having diameters of about 100 mm (4 in) and
greater. Even in such blastholes, reliable initiation often requires the use of large powerful
boosters.

Because standard emulsions have weight strengths which are significantly lower than that of
ANFO, the weight strengths of 'Heavy' ANFOs are lower than 100 and usually lie in the 90 - 95
range (see Section Weight Strength and Bulk Strength ANFO's weight strength of 100).

Despite their slightly lower weight strengths, 'Heavy' ANFOs exhibit higher densities and, as a
result, have bulk strengths which are about 20 - 45% greater than that of ANFO. Where 'Heavy'
ANFO is used in bulk form, therefore, it generates about 20 - 45% more energy than an equal
volume (and equal charge length) of ANFO.

When an emulsion rather than aluminium powder is added to ANFO, the bulk strength
increment per total dollar outlay increases. For this and other reasons, Heavy ANFO blasting
agents have all but superseded AlANFO wherever charges stronger than bulk ANFO are required
for dry or, at most, damp blastholes.

Where ANFO is replaced by 'Heavy' ANFO, opportunities exist to increase/improve:

• burden distances and/or blasthole spacings,


• fragmentation, displacement and muckpile looseness,
• the ability to break and displace heavy front-row toe burdens, and
• the breakage of a particularly strong bed within or overlying generally weaker rocks (as,
for example, where thinly bedded shales contain a thick bed (say greater than 2 m) of
strong massive sandstone or are overlain by an equally strong caprock).

'Heavy' ANFOs have been developed to replace bulk ANFO wherever the higher bulk strengths
of the former can be shown to lead to lower overall excavation costs.

Despite the high inherent water resistance of the emulsion component, Heavy ANFOs do not
exhibit high resistance to dilution and dissolution by water. The water resistance of 'Heavy'
ANFO is appreciably lower than that of watergel and especially all-emulsion blasting agents.
Therefore, bulk 'Heavy' ANFOs should not be drop-loaded into wet blastholes.
Low Density ANFO

In operations in which ANFO-type blasting agents are used, the advantages of overbreak control
can be achieved by bulk charging ANFO/diluent mixtures into perimeter blastholes.
ANFO/Polystyrene compositions(ANFOPS) hold high potential in this regard, as reliable
detonation in dry small-diameter blastholes can be achieved in mixtures which contain up to
about 75% (volume basis) PS. As one would expect, such blasting agents exhibit densities and
energy yields per metre of charge length that are down to about 25% of those of straight ANFO.
The correspondingly lower shock intensities and gas volumes help minimize overbreak in the
surrounding rock.

If such mixtures are transported considerable distances and/or handled a great deal between
mixing and charging, the resultant segregation may make it necessary to remix immediately
before charging. Special oils or other ingredients can be added as tackurfyers to reduce the
segregation problem.

The potential usefulness of ANFOPS mixtures is being realized in controlling overbreak in tunnels
and, to a lesser extent, alongside final or semi-permanent walls in surface mines and quarries.
But in weak, soft and/or friable rocks (in which significant fractions of ANFO's energy is likely to
be wasted in crushing and/or plastically deforming the adjacent rock), ANFOPS could well be a
cost-efficient replacement for ANFO in production blasts.

Other diluent materials besides polystyrene beads can be used to reduce the density of ANFO.
These include sawdust, peanut shells, rice husks and Pealite. The selection criteria for the diluent
material include:

• available in large quantities


• mixes well with ANFO
• low density
• low cost
• small size particles
• low water content
• contains little dust

High Density ANFO

High density ANFO is manufactured using the higher density mini-prilled ANFO. This results in
an ANFO with a density of approximately 1.05 g/cc. The higher density results in a 23% higher
Bulk Strength (i.e. 23% higher weight of explosive for the same charge length).

Emulsion-Based Blasting Agents

Cartridged Emulsion-Based Blasting Agents

Cartridged emulsion blasting agents are simply detonator-insensitive versions of the products
covered earlier; they can be used only in blastholes having diameters larger than about 100 mm
(4 in).

Bulk Emulsion-Based Blasting Agents


Bulk emulsion-based products are pumped into blastholes through a long flexible rubber hose.
They are either prepared in a static plant (often located on site) and then charged into blastholes
by a pump truck or prepared and charged by a mix-pump truck. Each system (i.e., the pump-
truck system and the mix-pump-truck system) has certain advantages over the other.

The ease and speed of charging emulsions from mix trucks and mix-pump trucks enable large
mining and quarrying projects to achieve substantial savings in labour costs (cf. where cartridged
emulsions are employed).

Doped Emulsions

Doped emulsions consist of prills of ANFO or AN within an emulsion matrix, the percentage of
emulsion usually lying in the 70 - 90% (weight basis) range. The particle size of the dry phase is
typically 400 - 500 times that of a microdroplet.

Provided that the dry phase content is maintained below about 35% (weight basis), there is
sufficient emulsion to completely cover the (soluble) dry phase. As the dry phase content
increases beyond about 35%, the ability of the emulsion matrix to protect the dry phase
decreases and, as a result, more and more of the dry phase can be dissolved by blasthole water.
Also, as the percentage of ANFO increases the viscosity of the explosive increases, making it
harder to pump. It is very hard to pump doped emulsions with concentrations of ANFO greater
than 35%.

As greater percentages of a dry phase are added to an emulsion, both the reaction rate and
detonation velocity decrease. The initial reaction proceeds through and takes place within the
emulsion matrix; the reaction in the (coarser) dry phase is initiated by the emulsion's reaction
and extends over an appreciably longer period. Therefore, by adding a dry phase, the overall
duration of the reaction is increased, thereby increasing heave energy (and, hence, muckpile
looseness) at the expense of strain wave (or shattering-type) energy. In relatively weak and/or
porous rocks, the presence of a dry phase has the beneficial effect of reducing energy losses
associated with excessive fragmentation and plastic deformation close to the charge.

Where the dry phase is totally enveloped by highly reactive emulsion, its reaction commences
with great vigor and proceeds to completion more easily and efficiently than in the case of ANFO
(where prills are surrounded essentially by air). Therefore, it seems reasonable to believe that
the actual:theoretical energy yield ratio for an emulsion/dry phase blend is greater than that for
ANFO. This is perhaps the most important factor to explain the observation that such products
perform better than their theoretical relative weight strengths suggest. In many published works
on blasting, the relative importance of fragmentation has been overemphasized at the expense
of muckpile looseness. Whilst fragmentation is usually the most influential feature of a muckpile,
looseness has a very considerable and hitherto neglected effect upon the speed and cost of
digging and hauling (and crushing?) operations. The displacement required to provide the
desired amount of muckpile looseness depends very largely upon the explosive's heave energy.
By adding a dry phase to an emulsion, the heave energy and, hence, muckpile looseness are
increased. The inclusion of up to about 35% dry phase has the beneficial effect of increasing
muckpile looseness without detracting from effective fragmentation.
Figure 5. Doped emulsion explosive.

Watergel (Slurry) Blasting Agents

Watergel blasting agents (i.e., detonator-insensitive watergels) can be used only in blastholes
having a diameter exceeding about 100 mm. They are of two types:

• cartridged watergel blasting agents


• pumped watergel blasting agents

Cartridged Watergel Blasting Agents

Because they cannot be reliably initiated by a detonator or by a single strand of detonating cord,
these watergels have important safety advantages over the dynamites. Like ANFO, they require
high-explosive boosters (e.g., PENTEX) to initiate them reliably.

Pumped Watergel Blasting Agents

Pumped watergels have been largely replaced by (newer) pumped emulsions (see Section
Emulsion-Based Blasting Agents, Bulk Emulsion-Based Blasting Agents and Doped Emulsions).

Water Resistant ANFO

The primary drawback with using ANFO in blasting operations is its lack of water resistance. In
order to overcome this a gelling agent is sometimes used to coat the ANFO prills. When the
gelling agent comes into contact with water the gel is activated, forming a water-resistant
barrier around the explosive. The explosive is usually packaged in 25 kg (50 lb) bags and is poured
into dry or damp holes. The explosive cannot be used in wet holes unless the holes are
dewatered.
Figure 6. Water resistant ANFO.

Pressure Desensitization

Causes of Desensitization

Most explosives become less sensitive at higher densities. This relationship is more
pronounced for those compositions which rely heavily or totally upon unreinforced air or gas
bubbles for sensitivity. For example, the variation of sensitivity with density is much greater for
air-sensitized, or gassed, watergels and emulsions than for dynamites.

Charges can be densified and, hence, desensitized in three major ways:

• by hydrostatic pressures
• by dynamic (i.e. blast-induced) pressures
• by the combination of hydrostatic and dynamic pressures

As one would expect, the degree of desensitization can be greatest (and blasting results
poorest) where appreciable hydrostatic and dynamic pressures act in unison.

Desensitization By Hydrostatic Pressures

As wet blastholes become deeper, there is a greater need to ensure that explosives have
adequate tolerance to hydrostatic pressures. Those emulsions and watergels which do not
contain glass microballoons or an equivalent ingredient are highly compressible (see Figure 1).
As the density of such products increases with hydrostatic pressure, the charge density
approaches the dead-press density (i.e., the density at which the charge cannot be initiated by
even the most powerful booster). Under adverse groundwater conditions, the dead-press
density could well be exceeded at the bottom of very deep blastholes.

Figure 1. Typical density curves for a bulk wet blasthole explosive sensitised
using entrained air or chemical gassing.

Desensitization by Dynamic Pressures

Explosives can be pressurized and, hence, desensitized by dynamic events:

• by the channel effect


• by the action of earlier-firing adjacent charges

Channel Effect

Where a small-diameter cartridged explosive is charged without tamping into a dry blasthole, a
shock wave in the resulting air annulus always outruns the detonation wave in the explosive
column. This air shock exerts a lateral pressure on the explosive just ahead of the detonation
wave front. If the stability of the detonation wave is little above or at its critical level, this
lateral pressure can densify and effectively desensitize the explosive just a few microseconds
before the detonation wave is due to arrive at the particular point in the charge. This
phenomenon, the so-called channel effect, can cause the detonation to fade and eventually
cease. The possibility of failure of charges through this effect can be reduced by:

• using higher-velocity explosives of relatively high sensitivity in cartridges of the largest


practicable diameter, and
• ensuring that the charge does not rest on a layer of drill cuttings or rock fragments
that drilling failed to clear from along a sub-horizontal blasthole.

The channel effect can be eliminated by successive tamping of charge elements. This may
mean the tamping of alternate cartridges or, for long cartridges, every cartridge.

Pressure Imposed by Earlier-Firing Charges

Where two blastholes are closely spaced (either by design or as a result of errors) and are
initiated on different delays, the earlier-firing charge can dynamically pressurize and, hence,
desensitize the other charge a fraction of a second before it is scheduled to detonate. This
pressure can be exerted:

• by the compressive strain wave passing through the later-firing charge,


• by lateral deformation of the blasthole and consequential squeezing of the charge due
to the movement of rock, and/or
• by explosion gases or groundwater that stream through pre-existing and/or blast-
generated cracks into the later-firing blasthole.

These three modes of application of pressure are presented in their expected order of
occurrence but by no means in their order of importance.

The first mechanism can be expected to occur only where the delay period between the firing
of adjacent charges is equal to

• the time of detonation of a long or slowly propagating trunkline, or


• the actual time period between the detonation of two down-the-hole delays (which
exhibit scatter).

The probability of the second mechanism occurring is greater in weaker, more porous and
water-saturated rocks.

The third mechanism tends to occur most readily in rocks which exhibit closely spaced physical
discontinuities. High-pressure explosion gases can stream though an interconnecting fissure
(or network of fissures) into a later-firing blasthole and compress and/or separate its charge
before its moment of initiation. With emulsion and watergel explosives (the products which
are most susceptible to dynamic pressure desensitization), such mechanisms could well cause
sufficient desensitization to result in an appreciable reduction in blasting performance. Such
desensitization can sometimes occur in tunnelling, shaft sinking and trenching-type operations
(where charges have small diameters and blastholes are close).

Should operators suspect that the rock and/or blast design at their operation is/are
encouraging such desensitization, efforts should be made to use charges which contain glass
microballoons or an equivalent ingredient. Such ingredients give the air/gas bubbles (upon
which sensitivity largely depends) dimensional permanence. With the mechanical strength that
is provided by such additives, bubbles are less susceptible to collapse during application of
dynamic pressures. However, dynamic pressures can often exceed the breaking strength of
microballons (which is typically about 400 psi). Also, there is some evidence that voids can
amplify the effect of shock waves and cause premature firing of detonators.

Desensitization By Hydrostatic Plus Dynamic Pressures

The resultant deleterious influence of hydrostatic and dynamic pressures on charges will be
greater than each effect in isolation. The effect of hydrostatic pressure on the performance of
a given charge can be quantified with reasonable accuracy. Unfortunately, the effects of
dynamic pressures cannot be predicted with sufficient accuracy, largely because of our lack of
knowledge of the ranges and durations of pressures which charges can experience. When
confronted with the possibility of charges experiencing significant (or, under adverse ground
conditions, even appreciable) dynamic pressures, operators would be wise to adopt a
conservative approach by subjecting each charge to a hydrostatic pressure which is
appreciably lower than that at which it becomes dead pressed.

Explosive System Selection

Explosive Performance Comparison

The following is a general comparison of some of the performance criteria for two common
bulk explosives used in wet blastholes.

Watergels

• high water resistance


• low shock energy
• high gas energy
• high cost

Emulsions

• high water resistance


• high VOD
• high shock energy
• low gas energy
• low cost

The above comparison of performance criteria may be influenced by the sensitisor and/or fuel
used in the manufacture of the explosives.

Explosive Selection Criteria

Following are some of the factors and guidelines to be considered when selecting an explosive
for a particular blasting situation.
Rock Properties

• density
• geological structure
• strength
• water conditions

Blast Requirements

• fragmentation
• muckpile shape and looseness
• environmental impact

Guidelines

• use high VOD and high density explosive where breakage is required
• when gas energy is required use low VOD explosive

Bulk Explosive Delivery Systems

There are many different bulk explosives delivery and loading systems available. They are
selected based on the requirements of the blasting operation. Requirements differ depending
on whether it is a surface or underground operation as well as with the explosives used and
the explosives charging capacity required.

For surface mines and quarries the common bulk explosives used are ANFO, ANFO/Emulsion
blends, and pumped emulsions.

Figure 1. ANFO auger truck.


The commonly used systems for loading blastholes with these explosives are ANFO auger
trucks, auger/blend trucks, auger/pump trucks, and emulsion pump trucks. ANFO auger trucks
(Figure 1) load only ANFO. They usually consist of a bin for AN prill, a diesel (fuel oil) tank, and
an auger for mixing and discharging the ANFO. These trucks can have varying capacities and
discharge rates.

Auger/blend trucks load ANFO or Heavy ANFO (ANFO/Emulsion blends). They are similar to
ANFO trucks except that they have an additional bin for emulsion. The AN prill, fuel and
emulsion are mixed in the auger and discharged. A hopper and pump are sometimes added to
these trucks (auger/pump truck) so that Emulsion and doped Emulsion explosives can be
manufactured and pumped to the bottom of blastholes for charging wet holes (Figures 2 and
3).

Figure 2. Auger/blend truck.


Figure 3. Auger/blend truck with pump.
Emulsion pump trucks are used only for loading emulsion and doped emulsion explosives.
Often the explosive is manufactured at a plant and loaded into a large tank on the truck. If
chemical gassing is used to sensitise the explosive and control its density then this is added to
the explosive as it is pumped down the blasthole.

Figure 4. Pressure vessel for loading underground blastholes.


For underground applications primarily ANFO and Emulsion explosives are used. ANFO is
usually loaded using a pressure vessel similar to that shown in Figure 4. These pressure vessels
can also be used to load emulsion explosive and can be mounted on an underground vehicle.
Emulsion explosives are also pumped for loading into both up and down holes.
Figure 5. Underground explosive loading vehicle.

Figure 6. Underground explosive loading vehicle


Figure 7. Pumped loading of emulsion explosive for underground blasting.

Explosive System Selection Considerations

Items that should be considered when selecting the explosives type, delivery system and
supplier include:

• type and range of products


• explosives quality
• consistency of supply
• delivery systems
• technical support
• cost

Review #2
The randomly selected multiple-choice questions below are designed to review your
understanding of the material covered in the preceding sessions. Your selections are lost when
you leave the review page. On return the review will start afresh with a new selection of
questions.
This review is currently set to practise mode. To optimize your learning experience you need to
register for certification before entering the course. Certification tests more rigorously, keeps
track of your answers to the multiple choice review questions, and enables you to report and
submit your review scores to complete the certification process. If you have already registered
and been approved for certification then you should Exit and re-enter before proceeding.

Each question below has one or more correct responses. Your selection of a response is
immediately marked correct or not.

Q1. Features of emulsion explosives include ...

excellent water resistance?

significant reduction in detonation sensitivity at low


temperature?

insensitive to detonation by friction, impact and fire?

relatively low production of smoke and toxic gases?

elimination of harmful or unpleasant physiological effects (e.g.,


headaches)?

cartridges are firm and easy to handle, charge and tamp?

a high shock energy : heave energy ratio ... suited to


fragmenting strong massive rocks?

a high heave energy : shock energy ratio ... more suited to


burden displacement / throw?

Q2. Relative effective energy (REE) is ...

a more accurate reflection of actual blasting conditions than


Available Energy?

the cumulative energy released by detonation to a cut-off


pressure of 200MPa?

the cumulative energy released by detonation to a cut-off


pressure of 100MPa?
a comparative measure based on the effective energy of ANFO?

Q3. Blasting agents are ...

detonator-sensitive explosives?

detonator-insensitive explosives?

consist largely of ammonium nitrate (AN)?

can be reliably detonated by a # 8 strength detonator or strand


of 10 g/m detonating cord?

require a strong stimulus, such as a booster, for detonation?

Q4. Explosive system selection criteria and considerations include ...

rock density?

geological structure?

rock strength?

groundwater conditions?

required rock fragmentation?

required muckpile shape and looseness?

environmental impact?

consistency of explosive supply?

delivery and charging systems?

explosive cost?

Q5. The primary ingredients of emulsion explosives are ...

a water-based solution of ammonium nitrate (AN)?


fuels (e.g. diesel)?

solid AN (AN-prills)?

trinitrotoluene (TNT)?

sensitizers (e.g. air, aluminium, micro-balloons)?

gum?

emulsifiers?
Blast Design and Assessment for Surface Mines and Quarries (Text Level)
Part 3: Initiating Devices and Systems

Introduction

Introduction

Initiating devices and systems are designed to activate explosive charges:

• from a safe distance


• at a pre-determined time
• in a pre-determined sequence
• with pre-determined time intervals between successive detonations

Today's initiating systems comprise explosive and inert components which transmit signals to
explosive charges by non-electric or electric means. They are wholly or partially consumed in
the blast. Small quantities of plastic tubing or wire will remain in the muckpile.

Non-electric initiating systems utilize chemical reactions, which can range from slow burning to
rapid violent detonation, to initiate explosive charges either directly or via non-electric
detonators. Electric initiating systems require a device which can generate or store electrical
energy; this energy is transmitted to electric detonators by a circuit of insulated conductors. In
some cases, a combination of electric and non-electric initiating systems is used to initiate
blasts, but there is an overall trend towards totally non-electric systems.

A wide range of initiating systems is now available. The diversity of products is necessitated by
large variations in applications and operating conditions. Selecting the most suitable system
for a given application requires one to consider factors which include:

• explosive type and charging method


• length, inclination and diameter of blasthole
• decking requirements, stemming material and stemming method
• rock mass properties and blasthole temperature
• desired blasting results (in terms of fragmentation, looseness, muckpile profile,
overbreak, etc.)
• environmental constraints, including limits on air vibrations and ground vibrations
• presence or risk of stray electrical currents, static electricity, radio frequency energy,
electrical storms, etc.
• the size of blasts and the required number of delays
• duration of charging or "sleeping" prior to blasting
• presence of groundwater and hydrostatic pressures

Before a system is selected, the manufacturer's product technical information should be


consulted for detailed recommendations.
Detonators

General

Detonators are compact devices which are designed to initiate explosive charges safely and
efficiently. Common elements include base charge, primary explosive and delay powder.
Typical primary explosives are mercury fulminate (HgC2N2O2), lead styphnate
(PbC6HO2(NO2)3) and lead azide (PbN6). A typical base charge would be PETN or RDX.

These sensitive components can be initiated prematurely if sufficient impact, heat, friction or
electrical energy is applied. Therefore, detonators should be stored, transported, handled and
used with care and respect, in accordance with government requirements and the
manufacturer's recommendations.

Detonators should be stored in cool, dry, well-ventilated magazines. As detonators may


become less reliable with increasing age, stocks should always be used within the time period
recommended by the manufacturer. This is particularly important if detonators are stored in
warm, humid magazines. The accuracy and precision of delay detonators deteriorate with
increasing age, especially under hot and moist storage conditions.

When used inside blastholes, detonators should always be secured inside suitable primers
which fully enclose the detonator shell to protect it from abrasion or impact during charging.
Bare (exposed) detonators should not be placed inside blastholes.

Shock-Tube Detonators

Non-electric detonators which incorporate "shock tube" are now used widely in blasting
operations. Figure 1 is a section through a non-electric shock tube, Figure 2 is a photograph of
a non-electric shock tube.
Figure 1. Section through a non-electric shock-tube.

Figure 2: Non-electric shock-tube.


These detonators are assembled from:

• a high-strength, non-electric detonator which features a PETN base charge and


pyrotechnic delay elements inside an aluminium shell
• a length of shock tube (one end of the tube is crimped into the detonator shell; the
other is closed off by a waterproof seal)
• an explosive charge of HMX and aluminum coating the inside of the plastic tubing
• an inert plastic connector (This is attached to provide a simple and secure means of
connecting the shock tube to a trunkline of detonating cord)

Non-electric detonators can be reliably initiated by a number of different mechanisms which


transmit a signal from an external source. This initiating energy is transmitted to the detonator
through an air shock wave which propagates down the shock tube at a velocity of
approximately 2000 m/s (6500 ft/s). An explosive load of roughly 0.2 g/m (1 gr/ft) provides the
means to support the air shock wave through the tube. The shock tube remains intact after the
shock wave has passed, causing no disruption to its surroundings.

The base unit of a non-electric detonator consists of a transition unit (T-Unit) that changes the
detonation wihtin the shock tube to a deflagration that starts the delay unit (a metal/metal
oxide or metal/salt formulation), this is followed by the priming charge that changes the
deflagration back to a detonation which then ignites the base charge of PETN or RDX. A section
through a detonater base unit is shown in Figure 3.
Figure 3: Section through a non-electric detonator.

A range of shock-tube detonators is available to suit most mine and quarry blasting
applications. Different types and lengths of shock tube are combined with various delays to
give the following products:

• LP detonators, having widely spaced "long-period" delay intervals and short lead
lengths
• MS detonators, having closely spaced millisecond-delay intervals and a wide range of
lead lengths
• Combination detonators, having both a surface delay unit and an in-hole delay
detonator on the same shock tube (a typical combination is 25ms/500ms)

Shock tube assemblies provide a high level of safety against accidental initiation by stray
electrical currents, static electricity and radio frequency energy. They cannot be initiated by
flame, friction or impact normally encountered in mining/quarrying operations. This non-
electric initiating system also eliminates the earth current leakage problems which often occur
with electric detonators in wet conditions. These safety features have led many operations to
replace electric by non-electric detonators.

Separate pieces of shock tube cannot initiate each other through direct contact or knots. In
order to avoid problems caused by assembly of components in difficult conditions, products
are factory-assembled with various lengths of shock tube attached to non-electric detonators.
Shock tube assemblies can be reliably initiated by another detonator or a suitable detonating
cord which has been connected to it in the recommended manner. Shock-tube detonators are
simple to use, and large initiating systems can be quickly built up by clipping or taping different
components together.

This type of detonator has proved to be robust and reliable, but must be handled with
reasonable care to prevent damage. If the tube is pierced, cut or split, moisture may enter it
and desensitize the reactive powder leading to a misfire.

When placed inside blastholes, every shock-tube detonator should be embedded within a
suitable primer and should point back towards the blasthole's collar. Where a cartridged
explosive primer is used, the shock tube should lie alongside the primer, not half-hitched
around it (Figure 4). Where a cast primer is used, a primer with a recessed base should be used
to protect the shock tube from damage. The shock tube should be threaded through the
primer's tunnel (Figure 5).
Figure 4. Non-electric detonators secured inside primers.

Figure 5. Non-electric detonators secured inside primers.

Shock-tube detonators do provide a high degree of protection against accidental initiation by


static electricity. Nevertheless, semi-conductive charging hose and properly grounded charging
equipment should be used when blow-loading blasting agents such as ANFO. Shock-tube
detonators should not be used inside non-conductive (e.g. plastic) blasthole liners into which
blasting agents are blow-loaded. A non-conductive liner would isolate the charging hose from
the blasthole wall, and this may allow electrostatic charge to accumulate.
Shock-tube detonators can be reliably initiated by clipping the plastic connector to a trunkline
of detonating cord which has a core charge of 3.6 to 5.0 g/m (18 to 25 gr/ft) of PETN. The
following method should be used to connect shock-tube detonators to a detonating cord
trunkline:

• Lay out the detonating cord in an orderly network which allows each shock tube to be
connected without stretching it.
• Use a closed loop of detonating cord wherever possible, to provide two paths of
initiation to each shock tube.
• Clip each connector to the detonating cord trunkline, keeping the cord and shock tube
at right angles.
• Pull the end of each shock tube through its connector until the tube is straight and taut
between the connection and blasthole's collar.
• Check to ensure that none of the shock tube between the connector and blasthole's
collar crosses over or lies within 200 mm (8 in) of the detonating cord.

Detonating cord functions at 3 to 4 times the speed of shock tube. Therefore, all connections
must be made with the detonating cord and shock tube at right angles, to avoid approach-type
"cut-off" failures in the shock tube.

Electric Detonators

Electric detonators are used as starter detonators to initiate many blasts, but are now rarely
used inside blastholes. Figure 6 is a photograph of an electric detonator assembly. Blastholes
are now generally charged using shock-tube detonators which are connected together by a
trunkline of detonating cord. When all charging and preparations for blasting are complete,
electric starter detonators are attached to the detonating cord. This enables the blast to be
fired safely from a remote location, using electrical cables attached to a portable exploder or
mains firing system.
Figure 6: Electric detonator assembly.
An instantaneous electric detonator consists of an aluminium shell which is closed at one end
and which contains a base charge of high-explosive PETN (a white powder), a sensitive priming
charge and an electric fusehead. A section through an electric detonator base unit is shown in
Figure 7. The fusehead consists of a fine metal bridgewire which is surrounded by a sensitive
flashing composition, and is soldered across the ends of two insulated lead wires. The lead
wires pass through a rubber plug which is securely crimped into the shell to provide a water-
resistant seal. When sufficient electrical energy is passed through the lead wires, the fine
bridgewire becomes hot enough to ignite the fusehead, which initiates the priming charge and,
hence, the powerful PETN base charge.

Figure 7. Section through electric detonator base unit.


The electrical resistance of each detonator depends on the length of its lead wires. The total
resistance of a circuit of electric detonators and the shot-firing line should always be
calculated, then measured with an approved blasting circuit tester to ensure that the exploder
or mains firing system will provide enough energy to reliably fire all detonators. The
recommended minimum firing currents are 1.5 amps DC or 2.0 amps AC for each series circuit.
If a capacitor-discharge blasting machine is used, each series requires an energy input of 20 MJ
per ohm of resistance.

Instantaneous electric detonators with a No. 8 or higher strength base charge are
recommended as "starter" detonators for initiating detonating cord trunklines. The
recommended procedure for using electric starters is as follows:

• Complete all charging operations and remove all equipment and tools from the blast
area. If an exploder is to be used to fire the blast, ensure that its key is in possession of
the shotfirer.
• Test the shot-firing cable for continuity and resistance using an approved blasting
circuit tester.
• Before touching electric detonators, make contact with an effectively grounded point,
to disperse any static electrical charges which may have accumulated.
• Use two electric detonators for insurance, and check both for continuity and resistance
using an approved blasting circuit tester before connecting them to the shot-firing
cable.
• Attach the detonator lead wires to the shot-firing cable in a simple series circuit.
Insulate the connections and ensure that they do not touch any electrical conductor
such as steel pipes, rails or pools of water.
• Securely attach the starter detonators to the detonating cord trunkline, with both
detonators pointing towards the blast.
• Cover the starter detonators with drill cuttings or shield them, to prevent any shrapnel
from damaging the shock tube system.
• Retreat to the safe shot-firing location, placing the required barricades or signs to
prevent any person from entering the blast area.
• At the shot-firing location, test the firing circuit for continuity and resistance using an
approved blasting circuit tester, then connect the ends of the shot-firing cable
together until firing time.
• At the approved firing time, connect the ends of the shot-firing cable to the exploder
or mains firing system, then fire the blast using standard procedures.
• Immediately after firing the blast, disconnect the shot-firing cable from the exploder or
mains firing system and connect the ends of the shot-firing cable together.
• Remove the key from the exploder, or close and lock the mains firing box.

Electric detonators can also be used inside blastholes, but this practice is now becoming
uncommon. The safety, simplicity and flexibility of shock-tube detonators has enabled them to
replace electric detonators in blastholes.

Electric delay detonators have a pyrotechnic delay element inserted between the electric
fusehead and the priming charge inside the detonator shell. Two types of electric delay
detonators are available for different applications, viz.:

• Long-period detonators - these have widely spaced delay intervals and short lead
lengths.
• Millisecond-delay detonators - these have closely spaced millisecond-delay intervals
and a wide range of lead lengths.

Electric detonators are supplied in tight coils, with the bared ends of the lead wires shorted
together to protect them from accidental initiation by stray electrical currents or radio
frequency transmissions. They should remain this way until just prior to connecting them
together or to the shot-firing cable.

Electronic Detonators

Non-electric and electric detonators utilize pyrotechnic delay elements that "burn" at a
controlled rate for the specified delay time. Inherent in all pyrotechnica delay elements is
"scatter" in the design firing time due to variances in the burn rate (see Delay Scatter).
Technology has been developed to replace the pyrotechnic delay element with an electronic
programmable chip. A number of manufacturers have developed delay systems that
incorporate these programmable devices. Figure 8 is a schematic of a such a system. In
summary, the system is as follows:

Figure 8. Schematic of electronic detonator system.


• Software that is used to program a field unit.
• The field unit is connected to all detonators in the blast. The unit communicates with
the delays, checks status and can be used to change delay time.
• Wiring used to connect all the delays in series.
• Delay units that are placed in each blasthole. The units have unique identifiers and can
be individually polled and assessed.

Electronic detonators offer the potential for very accurate delay times. In these detonators,
the pyrotechnic delay element is replaced by an electronic timing circuit. Precision timing, with
microsecond accuracy is achievable with this type of circuitry.

Potential benefits that may result from the use of these detonators have been reported as:

• significantly reduced ground vibrations


• increased fragmentation
• increased blast size
• improved contour blasting
• increased safety

The current high cost of manufacturing the electronic delay element has minimized the
application of this type of delay detonator. However, as production costs are reduced, more
units will be used in everyday blasting operations. Trial blasts have been conducted using these
systems as reported by Gregg (1994), Cunningham (1994), Dent (1994), Bernard & Laboz
(1995), Katsabanis et al. (1995) and Yamamoto et al. (1995).

Non-Electric Delay Connectors

Detonating cord trunklines can incorporate non-electric delay connectors which create a short
time delay between detonation of different sections of the trunkline network. Detonating
Relay Connectors (DRCs) and MilliSecond Connectors (MSCs) allow millisecond-delay timing to
be introduced into detonating cord trunklines. When used in conjunction with a range of in-
hole delays, DRCs and MSCs can be used to provide additional delays for large blasts (see
Shock Tube Detonators). Different delays are colour-coded for easy identification.

Both DRCs and MSCs function bi-directionally, which simplifies hookup and permits trunkline
layouts with two paths of initiation to each blasthole. All trunkline networks should be laid out
as a closed loop, to provide two paths of initiation to each blasthole, and security against poor
connections.

DRCs consist of two identical miniature delay detonators inside a small plastic body. DRCs are
compatible with 5 to 10 g/m (25 to 50 gr/ft) detonating cords but do not function reliably if
used with low-energy cords such as 3.6 g/m (18 gr/ft) cords. DRCs are not recommended for
use in wet conditions, as water may cause instantaneous firings or misfires.

DRCs should be connected into a detonating cord trunkline as shown in Figure 9.


Figure 9. Connecting DRC to detonating cord trunkline.

Figure 10. Connecting MSC into a detonating cord trunkline.

MSCs consist of two identical delay detonators inside plastic cleat blocks, connected together
by a short length of shock tube (Figure 11). The length of MSCs enables them to be directly
inserted into cut trunklines, whereas DRCs require adjustment of the trunkline to compensate
for their short length. MSCs are compatible with detonating cords.
Figure 11: MilliSecond Connector (MSC).
MSCs should be connected into a detonating cord trunkline network as shown in Figure 10.

DRCs and MSCs are designed for use in detonating cord trunklines. They should never be used
inside blastholes to introduce delays into downlines.

Detonating Cords

General

Detonating cords are strong, flexible linear explosives which consist of a continuous core of
high explosive, covered by a seamless plastic jacket which may be overwrapped with textile
yarns. When detonating cord is initiated, it detonates along its entire length with a steady high
velocity. Detonating cords are an effective but violent means of transmitting energy from one
place to another.

All cords suitable for use in mine/quarry blasting contain PETN, a high-explosive powder, and
detonate at between 6.0 and 7.5 km/s. Detonating cords designed for specific tasks have core
charges ranging from 1.5 to 85 g/m (6.5 to 400 gr/ft), enclosed in appropriate outer covers.

The core of PETN gives detonating cords a high velocity of detonation (VOD) and tremendous
brisance (shattering effect). PETN has a high melting point and is relatively insensitive to
initiation by impact or friction. The PETN core is covered by combinations of synthetic textile
yarns and plastics to give the finished product appropriate physical characteristics such as
water resistance, tensile strength, flexibility and abrasion resistance. Some detonating cords
contain PETN powder which has been treated with a special chemical to minimize penetration
by water (wicking) via a cut end.

Conventional detonating cords can be initiated by intense impact or friction, but are
insensitive to accidental initiation during handling and use. Energy sources such as static
electricity, stray electrical currents or radio frequency transmissions will not initiate detonating
cord. Some detonating cords may be initiated if subjected to extremely high temperatures, but
remain stable and safe to use below 80oC (176oF). For temperatures between 70oC and 80oC
(158oF and 176oF), exposure time should not exceed 24 hours.

Detonating cords should be stored with other high explosives. When stored correctly,
detonating cords have a minimum shelf life of 5 years.

Detonating cords should be cut using either an anvil-type cutting tool, or a sharp knife in
conjunction with a wooden block. Cutting devices with a shearing action must never be used to
cut detonating cords.

Detonating cords can be reliably initiated by means of a No.8 strength detonator which is
firmly attached to a dry section of cord at least 150 mm (6 in) from the cut end. Suitable
initiators include No.8 strength electric detonators, No.8 strength plain detonators with safety
fuse, and Trunkline Delays. The base of the detonator must point in the desired direction of
propagation (Figure 1).

Figure 1. "Starter" detonators connected to detonating cord.


Detonating cord which has been contaminated with oil or water may be less sensitive to
initiation and, therefore, care should be taken to avoid moisture penetration into cut ends. Cut
ends lying in water will absorb moisture by capillary action (wicking) at a rate which varies
from one type of cord to another. For initiation of wet cord tails, end-priming with a No.8
strength detonator is usually effective, but in extreme cases it may be necessary to use a cast
primer for reliable results. Once initiated, continuous (i.e., join-free) lengths of detonating cord
will continue to propagate through wet sections of undamaged cord. To ensure reliable
initiation of detonating cord, the use of a pair of starter detonators is always recommended.

Detonating Cord Trunklines

Detonating cords with core loads of 3.6 to 10.0 g/m (18 to 48 gr/ft) will reliably propagate
through joins made by simple knots. Therefore, they can be used as "trunklines" to transmit an
initiating signal from one point to another.
Figure 2: Detonating cords.
For joining or extending trunklines, separate lengths of cord should be tied together with a
square (reef) knot (Figure 3). For security, the knot should be 150 mm (6 in) from each cut end
and pulled tight, with the free ends taped back along the cord to ensure positive contact.

Figure 3. Connection for extending a trunkline.


Detonating cords with a core load of 3.6 to 5.0 g/m (18 to 25 gr/ft) will reliably initiate shock
tube, using standard connectors. Cords with loads greater than 5 g/m (25 gr/ft) are not
recommended for initiating shock tube, as an increased risk of approach-type cutoff failures
exists.
Figure 4. Connections between detonating cord downline and detonating
cord trunkline.
Detonating cord trunklines should be laid out in a neat and orderly manner, with no loops,
kinks, tight bends or excessive slack. A closed loop of detonating cord is recommended, to
provide insurance against poor connections. Where branchlines are required, they should be
attached to the main trunkline by either a double-wrap clove-hitch knot or a double half-hitch
knot (Figure 4). These connections should be pulled tight and made at right angles, to prevent
approach-type "cutoff" failures when the trunkline detonates. Figure 5 shows detonating cord
being used to initiate long period non-electric detonators in an underground development
round.

Figure 5. Example of detonating cord used as surface trunkline.


Detonating Cord Downlines

In most blasts, blastholes are initiated by in-hole delays to avoid downline cutoffs caused by
ground movement or flyrock during the blast. Thus, detonating cords are generally used only
when deck charging. The majority of blastholes are initiated by non-electric detonators whose
shock-tube downlines do not disrupt their surroundings in any way.
Detonating cord downlines can desensitize or continuously side-initiate the explosives which
surround them, depending on many factors which include:

• the explosive's sensitivity and method of sensitization


• in-hole density and temperature of the charge
• core charge and construction of the downline
• blasthole diameter
• rock type
• presence of water and its effect on the explosive's sensitivity
• number of downlines and their positions in the blasthole
• type of primer(s) and in-hole delay system

Many of these variables are difficult to measure or quantify, and thus it is not possible to
accurately predict the effect of detonating cord downlines on explosives in all circumstances.

Detonating cord downlines also affect collar stemming and interdeck stemming, creating a
chimney through which explosion gases can flow. This reduces the confinement and
effectiveness of explosives, and can cause sympathetic detonation of adjacent decked charges.

Every detonating cord downline should be lowered into its blasthole after securing the primer
to the free end of the cord. As soon as a downline is in position, it should be cut from the reel
and adequately secured at the blasthole's collar. Detonating cord downlines must not contain
knots or joins, as the lower section may not propagate reliably if moisture or fuel oil
penetrates the cut end of the cord. Such connections may also prevent additional primers
from sliding to their intended positions.

For attaching detonating cord downlines to a detonating cord trunkline, connections should be
made using either a double-wrap clove-hitch knot or a double half-hitch knot (Figure 4 above).
All connections should be pulled tight and made at right angles, to avoid approach-type cutoff
failures when the trunkline detonates.

Delay Scatter

Delay Scatter

Inherent in all pyrotechnic delay elements is some degree of "scatter" in the firing times of the
detonators. The burn rate of the delay element can be influenced by a number of factors,
including:

• composition of the element


• delay length and packing density
• temperature and humidity during storage
• age (oxidation of components)
• temperature in blasthole

An understanding of the magnitude of the scatter for a specific type of delay element can be
developed by undertaking measurements of approximately 30 units. The measurement of
delay scatter can be easily undertaken using high speed digital recorders. Figure 1 is a picture
of a typical measurement layout. In this case, non-electric detonators will be detonated on the
ground surface and the detonation recorded using a microphone. Output from the recording
device is shown in Figure 2. The detonation times for each cap can be easily identified and
compared to the design inititation time.

Figure 1. Delay scatter measurement.

Figure 2. Delay scatter measurement results.


Figure 3. Delay scatter distributions for different batches.

When the measurements are reviewed it becomes apparent that delay elements made during
the same production run tend to have delay times which are relatively closely "scattered"
about a common mean initiation time (but not necessarily the nominal initiation time). Typical
values for scatter in a single batch are one standard deviation being in the range of +/- 2.5% of
the average firing time. This is illustrated in Figure 3. However, when delay from different
batches are used in the same blast (which is usually the case) the combined scatter increases
significantly because the average firing time for each batch is different. Figure 4 is a summary
of measurements made on short period surface connectors. In this specific case the potential
for sub-optimal results is evident in the 42ms and 65ms connectors. Figure 5 shows the
cumulative results of several thousand measurements. The figure illustrates the general
tendency for delays to detonate somewhat later than the design delay time.

Figure 4. Example of measured scatter in surface MS connectors.


Figure 5. Histogram of delay scatter measurements.
Once recognized, the occurence of delay scatter should not pose significant problems if blast
designs incorporate the increased range of potential detonation times. Robust designs must be
developed to avoid the following problems:

• Overlapping, where a delay of an earlier period fires before a later period. This can
result in very large burdens, increasing confinement and the potential for detrimental
charge interactions.
• Crowding, where a delay fires early or late but very close in time to a dependent
blasthole. Similar effects, but to a lesser degree, can be expected.

Coarse fragmentation, flyrock, wall damage, high ground vibrations and air blast can result
from not properly incorporating delay scatter into blast design.

Delay Scatter and Blast Design

All blast designs should recognize the potential for delay scatter. This recognition can be in the
form of adequate delay times between dependent blastholes and compatible down hole to
surface delay ratios. Too often, blast designs are unnecessarily complex and can be reduced to
a near random detonation sequence by delay scatter. An example which is commonly
observed is a design which tries to have single hole detonation on an 8 or 9 millisecond
interval, but has down hole delays of 400 or 500 milliseconds. Delay scatter of 2% or 3% in the
down hole delays exceed the design interval between blastholes!

Blast designs can be evaluated with software programs that statistically assess the potential
for delay scatter to impact the design. The user provides delay scatter values based upon a
specific level of accepted risk and the model produces results which estimate the potential for
that level of risk to be exceeded. Output from a program of this type is shown in Figure 6.
Figure 6. Example of blast analysis program showing effect of delay scatter.
The use of electronic delay units (see Electronic Detonators) could greatly reduce the
magnitude of delay scatter. Figure 7 compares the relative amount of scattter between
pyrotechnic and electronic detonators. It can be seen that delay scatter in electronic delay is
reduced to microseconds rather than milliseconds or tens of milliseconds in the pyrotechnic
units. The lower amount of delay scatter will allow for designs that incorporate very close
initiation times for specific application such as cast blasting or vibration control without the
danger of overlapping or crowding.

Figure 7. Comparison of the relative amount of scattter between pyrotechnic


and electronic detonators.

Priming Systems
Priming System

The effectiveness of an explosive in priming or initiating another explosive is a function of the


amount of energy that is transmitted into the second explosive and of the rate at which this
energy is delivered. The detonation pressure of an explosive (see Detonation Pressure) is
defined as the pressure exerted at the detonation front as the explosive reacts. Detonation
pressure is the main factor which governs priming effectiveness.

Detonation pressure is a function of velocity and density and is approximated as follows:

where

P = detonation pressure (kilobars)


ρ = density (g/cc)
D = detonation velocity (m/s)

A rule of thumb is that the detonation pressure of the priming charge should exceed the
detonation pressure of the main charge in order to minimize overly long velocity run-up
distances.

Explosives that are readily initiated by a detonator and which have a high density and VOD are
the most effective primers.

For any type of primer, there is a minimum weight of explosive capable of imparting sufficient
energy to start a self-generating detonation front in another explosive charge. Where priming
is marginal, there is a risk of producing a low, transient VOD and reduced blasthole pressures
in the run-up zone. It has also been established that larger blasthole diameters require larger
primers because shock wave reflections do not contribute as much to the developing
detonation front. Even if the minimum primer weight is employed, the same steady-state VOD
will be achieved after a run-up distance of 2 to 8 blasthole diameters, and in many blasting
situations there may be no noticeable reduction in overall blast performance. In general,
recommendations on the size of primers take into account the non-ideal nature of detonations
in production blasting, and a factor of safety is included to reduce the risk of poor performance
and misfires.

Figure 1. Primer.
Figure 2. Primer.
Primer Effectiveness

Factors which may reduce priming effectiveness and which require a factor of safety are listed
below:

• contamination, dilution, non-ideal mixing of explosives


• excessive sleep times in the blasthole
• presence of ground water
• primer partially buried by drill cuttings or mud
• hydrostatic pressures in deep blastholes
• desensitisation by dynamic pressures transmitted from adjacent earlier-firing
blastholes
• water occlusion around the primer

Detonators represent a much more concentrated energy source than detonating cord and,
consequently, initiate primers with greater efficiency. Ultra high-speed photography has
established that detonation of the primer composition begins within a few millimetres of the
base charge of the detonator, after which the detonation propagates through the remainder of
the primer. Thus, virtually all of the potential energy of the primer is available for initiating the
surrounding bulk explosive.

Whilst common practice in mines is to use a detonator to initiate the primer, for those
operations that still employ detonating cord in blastholes the following points apply.

• Detonating cord downlines affect the explosive columns through which they detonate.
With ANFO, a downline can cause either detonation by side initiation, or compression
and desensitisation of the explosive around the cord.
• High-energy detonating cords tend to side-initiate ANFO in all blasthole diameters. The
initial VOD adjacent to the cord is low, with the detonation front accelerating as it
propagates across the charge. As a result of this, the average VOD of the overall charge
is lower than the steady-state VOD of an end-initiated ANFO charge.
• For a given blasthole diameter, there is a critical core load below which detonating
cord will not side initiate ANFO. Instead, detonation of the downline creates an
expanding shock wave and gas chimney in the charge. In large-diameter blastholes,
these lateral pressures cause local compression and desensitisation of the ANFO. The
cross-sectional area of the unaffected ANFO is still sufficient to allow the detonation to
propagate along the ANFO column, but there is some energy loss. In small to medium-
diameter blastholes, detonating cords side-initiate ANFO, severely desensitise it, or
dead-press it and cause complete failure of the detonation.
• In general, emulsion explosives are less sensitive to initiation than ANFO, and they are
generally not prone to side initiation by most detonating cords. The sensitivity of an
explosive increases with temperature, however, and some bulk explosives which have
been charged into blastholes at high temperature may remain sensitive to side
initiation by high-energy detonating cord until they have cooled.
• Emulsion explosives are more easily compressed and desensitised than ANFO, because
of the action of the shock wave on the gas bubbles within these explosives. Until the
bubbles re-form after compression, high-order detonation is prevented from
propagating beyond a certain distance from the primer, resulting in failure of the
explosive.

In general, the effects of a single detonating cord on bulk explosives are known to be
influenced by

• the weight of PETN per metre of detonating cord


• the type and quantity of covering around the detonating cord
• the sensitivity to initiation of the explosive
• the method of sensitisation of the explosive
• the initial density of the explosive when it is charged into blastholes
• the degree of confinement of the explosive (a function of the properties of the rock
and stemming material)
• the diameter of the blasthole
• the location of the detonating cord in relation to the axis or wall of the blasthole
• the temperature of the explosive

For these reasons, non-electric detonators are recommended for use in blastholes, as they
have no effect on the explosive prior to the detonator firing.

Primer Location

Experimental work carried out in small (<100 mm) diameter blastholes in a benching - type
operation has shown that the peak strain in the rock at the primer level increases by about
37% when the primer is moved from the bottom of the charge to bench floor level. This
increase in peak strain is due to the simultaneous detonation of those parts of the charge
which are equidistant from the primer (Figure 3). The improved blasting results with the
primer at bench floor level are especially noticeable where toe burdens are excessive.
Figure 3. Increasing peak strain by priming at bench floor level.
Similarly, more peak strain is generated in any relatively thick bed/band of strong strata in the
face by locating a primer centrally within the bed/band rather than well above or below the
bed/band.

Some operators place the primer at a known distance above or below bench floor level to
ensure that, should a misfire occur, the excavator operator does not dig directly into the
primer/detonator. This may be a valid reason for not placing the primer at bench floor level.

The advantages of bottom compared to top priming include:

• improved fragmentation, displacement and muckpile looseness;


• reduced toe problems, better floors and cleaner faces;
• reduced noise, airblast, flyrock and surface overbreak; and
• fewer cut-offs and misfires

There may be occasions when multiple primers are placed in a blasthole:

• to limit the potential for cutoff due to rock movement;


• to meet regulatory requirements;
• to ensure detonation of a sub-optimal explosive product;
• where multiple decks of explosives are placed in the blasthole.

In the first three cases above, the lower priming unit should be placed on a lower delay
number to ensure that optimal initiation is achieved. In the last case, delays should be
allocated for each deck of explosive.

Initiating Systems

General

The many available initiating devices can be combined together to create a large range of
initiating systems. The variety of products and possible combinations provide a choice for each
blasting application and there are several alternative methods available to fire a given type of
blast. In most blasts, however, a generally standard initiating system has been established,
based on the following components:

• a booster or primer containing the detonator


• non-electric in-hole detonators, initiated by shock-tube downlines
• a trunkline of detonating cord, to initiate the shock tubes
• electric starter detonators, to initiate the detonating cord trunkline
• an electrical shot-firing cable, with portable exploders, to initiate the blast from a safe
remote location

This general type of initiating system may not be appropriate for all applications, but it is used
in many surface mining/quarrying blasts today. It is widely accepted as being the safest, most
reliable and effective combination of products currently available. For each situation,
variations may be required to suit local conditions, but the same broad concepts will usually
apply.

Initiating Systems for Bench-Type Blasts

Bench-type blasts are now usually fired by an initiating system which includes the products
listed below (see Priming Systems):

• PETN/TNT cast boosters that contain the delay detonator and are used to detonate the
main charge.
• Millisecond (MS)-delay shock-tube detonators are used inside blastholes. Short delay
times are used between blastholes, to ensure that they interact effectively to produce
good fragmentation. Long-Period (LP) detonators are rarely used for benching,
because the relatively long delay intervals tend to produce coarser fragmentation.
However, LP detonators are sometimes used where it is more important to reduce
throw than to obtain good fragmentation.
• A trunkline of 3.6 to 5.0 g/m (18 to 25 gr/ft) detonating cord is used to initiate the in-
hole shock-tube detonators.
• For large blasts, shock-tube delay detonators can also be used outside blastholes, as
bridges to link several detonating cord trunklines together. This enables many rows of
blastholes to be fired in a single blast, using short delays to produce good
fragmentation. The in-hole detonators are connected to several separate detonating
cord trunklines, which are initiated in sequence by the bridge detonators.
• Either an instantaneous No.8 strength electric detonator or a NTD (noiseless trunkline
detonator) are used to initiate the detonating cord trunkline.

Initiating devices used in bench blasting should be charged and connected together according
to standard procedures which include the following.

• Each in-hole detonator is inserted into a primer which completely encloses the
detonator shell.
• The detonator/primer combination is lowered into the blasthole and the explosive is
loaded.
• A closed-loop detonating cord trunkline is used, to provide two paths of initiation to
each in-hole detonator. Several separate cord trunklines may be used, linked together
by shock-tube detonator bridges.
• Each in-hole detonator is joined by its connector to the nearest part of the detonating
cord trunkline. All connections between cord and shock tube are made at right angles,
with the tube pulled taut but not stretched. Connections are made at least 200 mm (8
in) away from knots or branchlines of detonating cord.
• The hookup is thoroughly checked, to ensure that all connections have been made and
that no shock tube crosses over or lies close to any detonating cord between the
connector and the blasthole's collar.
• If several detonating cord trunklines are to be fired in sequence, they are linked
together by shock-tube detonator bridges. The delay times of the bridge detonators
are selected to ensure that each trunkline is fired before it can be damaged by flyrock
or airblast. Separate trunklines are linked by pairs of identical bridges taped to short
lengths of detonating cord. This enables the detonators to be placed at least 2 metres
(6 ft) from any shock-tube connections.
• Two detonating cord lead-in lines are used to initiate the trunkline system. The lead-in
lines are long enough to locate starter detonators at least 2 metres (6ft) away from
any shock tube.
• Starter detonators are attached to the lead-in lines just before clearing the area for
firing. These two starters are securely taped to the lead-in lines, approximately 150
mm (6 in) from the cut ends. Both detonators point towards the blast and are covered,
to protect all shock tubes from detonator shrapnel.

Review #3

The randomly selected multiple-choice questions below are designed to


review your understanding of the material covered in the preceding sessions.
Your selections are lost when you leave the review page. On return the
review will start afresh with a new selection of questions.

This review is currently set to practise mode. To optimize your learning


experience you need to register for certification before entering the course.
Certification tests more rigorously, keeps track of your answers to the
multiple choice review questions, and enables you to report and submit your
review scores to complete the certification process. If you have already
registered and been approved for certification then you should Exit and re-
enter before proceeding.

Each question below has one or more correct responses. Your selection of a
response is immediately marked correct or not.

Q1. Detonating cord ...


is relatively insensitive to accidental initiation during handling
and use?

should be stored with other high explosives?

may be initiated by static electricity, stray electrical currents or


radio frequency transmissions?

should always be cut with a device which has a shearing action?

becomes less sensitive when contaminated by oil or water?

Q2. Delay scatter...

can be influenced by the composition of the delay element?

is independant of the temperature in the blasthole?

is generally lower within batch vs. between batches of delay


elements?

Q3. Which of the following is (are) true when discussing bench type
blasting?

In bench type blasting, MilliSecond (MS) delay shock tube


detonators are generally preferred over Long Period (LP)
detonators because they produce better fragmentation?

A closed-loop detonating cord trunkline is used, to provide two


paths of initiation to each in-hole detonator?

The detonator/primer combination is lowered into the blasthole


after the explosive is loaded?

Q4. Shock tube detonators...

are widely used in blasting operations?

are vulnerable to stray electrical currents?


may include a delay component?

may be initiated by flame or friction?

should be positioned pointing away from the blasthole collar?

Q5. Which of the following are common detonator components ...

mercury fulminate?

ammonium nitrate?

PETN?

aluminium powder?

lead styphnate?
Blast Design and Assessment for Surface Mines and Quarries (Text Level)
Part 4: Production Bench Blasting

Introduction

Production Bench Blasting

The following video clips provide examples of production blasting.

Video 1: Production blasting with delay sequencing.

Video 2: First blast in a quarry.

When starting to work a new mine or a new area of an existing mine, it is necessary to develop
one or more initial designs for production blasts. In this situation, some "rules of thumb",
derived over many years of relevant practical experience, should be used for developing these
designs. If a detailed assessment of rock mass properties has been carried out, modeling can
be used to assess the suitability of the designs developed, and possibly to indicate alternative
superior designs.

Initial blast designs must then be progressively improved through mine-site trials until they
have been optimised. Optimum blast designs are a prerequisite for minimising the total cost of
mining. Optimum designs help to produce the required fragmentation (Figure 1 below left),
muckpile looseness, muckpile profile, toe conditions and grade control. In some cases, blast
designs must also help to:

• minimise the probability of flyrock


• control ground vibrations and air vibrations
• minimise dilution
Figure 1: Optimum designs help to produce the required fragmentation,
muckpile looseness, muckpile profile, toe conditions and grade control.

Figure 2. Effect of fragmentation on the cost of drilling, blasting, loading and


hauling (Hoek & Bray (1977)).
The basic economics of rock excavation using explosives is shown in Figure 2 (above right). The
production of a well-fragmented and loose muck pile that has not been scattered around the
excavation area facilitates loading and hauling operations. This condition is at the minimum
total cost point on the graph. However, close to the final walls, drilling and blasting costs will
increase because more closely spaced and carefully charged blastholes will be required. In
order to achieve optimum results under both conditions, a thorough understanding of the
following parameters is required (see Figures 3 and 4 below):

• bench height
• nature of free face
• properties of the rock being blasted
• type of explosive

Figure 3: Design of effective burden and spacing for different blasting


patterns.

Figure 4: Production blasting.


• blasthole diameter
• blasthole inclination
• effective burden
• effective subdrilling
• effective spacing
• stemming
• initiation sequence for detonation of explosive charges
• delays between successive blastholes and/or rows of blastholes
Figure 5. Definition of blasting terms (Hoek & Bray (1977)).
Each of the factors listed above will be considered in relation to its influence upon the
effectiveness of the blast and its influence upon the amount of damage inflicted upon the
remaining rock. Figure 5 (right) illustrates blasting terminology for a typical bench blast.

Bench Parameters

Bench Height

Bench height normally lies in the 3 m (10 ft) to 18 m (60 ft) range. The selected bench height is
influenced by:

Figure 1: Bench excavation.


• statutory regulations (excessively high benches are unsafe and, therefore, not
permitted)
• rock mass properties
• the type and size of digging equipment
• grade-control requirements
• the need to maximise the overall cost efficiency of drilling and blasting

To date, drilling and blasting have not had a large effect on the selection of bench height. The
height of future benches should be selected only after considering the following points.

• An increase in bench height causes decreases in specific drilling (expressed in m/m3 or


m/t), the cost of drilling, the consumption and cost of primers and initiators, the
labour required for priming through firing, and the entire cyclicity of mining; and an
increase in the powder factor and cost of explosives.

• Optimum blasthole diameter increases with bench height; Table 1 shows the general
effect of bench height on optimum blasthole diameter. In general, an increase in
blasthole diameter causes a decrease in the cost of drilling.
• For vertical blastholes of a given diameter, the front row toe burden becomes
excessive for benches higher than a certain height. Where small diameter blastholes
are drilled in high benches, blastholes need to be angled, at least in the front row.
• Drilling accuracy becomes more critical in higher benches.

In surface mines, a high percentage of benches are 5 m (16 ft), 10 m (30 ft) or 15 m (50 ft) high.
These are "easy" bench heights to work with. But the probability of the optimum bench height
being an exact multiple of 5 m (16 ft) is low.

Consider a new large open pit in which the operator plans to blast in 10 m (30 ft) benches
down to a final depth of 240 m (800 ft). If the bench height were to be increased from 10 m
(30 ft) to 12 m (40 ft), the following advantages would be achieved.
• 24 benches would be replaced by 20 benches. The cyclicity of mining would decrease
by nearly 17%.
• For a given blasthole diameter, subdrilling would decrease by nearly 17% and the
drilled metres (feet) per year by about 2%.
• Drills would spend a higher percentage of their time drilling. Moving from one
blasthole collar to the next would take a smaller percentage of the drill shift.
• The consumption and cost of primers would decrease by nearly 17%.
• The cost of initiators would decrease by about 15%.
• The cost of labour for priming, charging, stemming, connecting up, checking the
trunkline network, guarding and firing would decrease by up to nearly 17%.
• Collar rock ( the rock alongside the stemming column and the origin of most
oversize/boulders) would decrease by nearly 17%. Therefore, overall fragmentation
would be finer and costs for digging, hauling and crushing would be lower.
• For a given area of blast block, the number of blasts per year would be nearly 17%
fewer. The time period required to withdraw from the blast area, wait for the blast
and then re-enter after the "all-clear" signal has been given is unproductive, and this
lost time would decrease by nearly 17%.

Disadvantages of these higher benches would include the following.

• Grade control may be more difficult. But any difficulty could possibly be minimised by
achieving greater control over the displacement of blasted material or by digging the
blasted ore in two equal benches, each with a height of 6 m (20 ft).
• It may become necessary to replace single-pass drilling by double-pass drilling. In this
event, blastholes in wet ground could be lost when adding the second drill steel. (With
sufficient planning, the selected drill would be able to drill the deeper blastholes in a
single pass.)
• There may be a greater need for angled blastholes in the front row of blasts. (With
sufficient planning, the blasthole diameter would be sufficiently large to prevent the
need for angled blastholes).
• The powder factor would increase by almost 8%.

There will be cases for which the resultant advantages outweigh the combined disadvantages
of higher benches. In other cases, the disadvantages will outweigh the advantages.

Nature of Face to which Blast Shoots

Where a blast shoots to a free face ( i.e., a rock/air interface), forward displacement of the
blasted rock can occur easily (see Figure 2a,b,c below). Where the face is choked by previously
blasted rock, forward movement is suppressed. To achieve the same fragmentation and
especially the same muckpile looseness, a change from free-face blasting to choked blasting
necessitates a slight contraction of the blasthole pattern and a corresponding increase in
powder factor. The required increase in powder factor is commonly 7% to 10%.
Figure 2a: Production blasting sequence ... blasting to a free face in a surface
coal mine.

Figure 2b: Production blasting sequence ... blasting to a free face in a surface
coal mine.
Figure 2c: Production blasting sequence ... blasting to a free face in a surface
coal mine.

Free-face blasting is recommended for bulk waste and for large orebodies in which selective
digging is not required. Where blast blocks contain thin or small orebodies within a waste
matrix, grade control can be promoted by suppressing lateral displacement by shooting to
choked faces.

Nature of the Rock Being Blasted

The nature and degree of heterogeneity of the rock mass is a very important factor in both the
design and results of blasting. Joints, bedding planes, faults, soft seams and other such
discontinuities can present considerable problems in blast design. In some cases, such
structures allow the explosive's energy to be wastefully dissipated rather than perform the
work intended.

The influence of the structural geology of a rock mass often overshadows that of the rock's
mechanical and physical properties. Pre-existing fractures such as joints and bedding plane
partings tend to dominate the nature of the fracture pattern produced by a detonating
blasthole.

Best fragmentation is usually obtained where the face is parallel to a natural fracture (e.g., a
joint). The newly formed face is then often a joint plane, and blasthole spacings appreciably
greater than the burden can be used satisfactorily.

In highly fissured rocks, there is always a high probability that a pre-existing crack will intersect
the charged section of the blasthole. In such cases, this crack can be extended (by the invading
gases) to a great length but, because it tends to prevent the development of high strains at
other points around the blasthole, few new cracks of any significant length are created. Even
when a new radial crack is developed, it often terminates prematurely where it intersects a
pre-existing crack, especially if the latter has been widened by gases from a blasthole on an
earlier delay.
It is heave energy that is important in highly fissured rock. The explosion gases jet into, wedge
open and, hence, extend the pre-existing cracks. The overall degree of fragmentation,
therefore, tends to be structurally controlled. Dense patterns of joints and bedding plane
partings increase fragmentation, and in strong rock a behaviour similar to that in weak, low-
strength strata occurs. As the burden becomes more highly fissured, longer stemming columns
and correspondingly lower powder factors or energy factors can be used. Because few new
fracture surfaces need be created in the blast, low-density, low-velocity explosives (e.g. ANFO)
are to be preferred. Higher velocity explosives tend to cause excessive fragmentation
immediately around the charge. Satisfactory results are achieved when the charges provide
sufficient heave energy to displace the rock into a loose, readily diggable muck pile, without
scattering the rock around the site.

Explosives and Priming System Selection

Explosives

The selection of one explosive or blasting agent out of the many available (see Explosives and
Charging Systems) must be based on an understanding of the basic properties of explosives,
the general types of explosives and blasting agents that have been developed, the meaning of
various government and industry classifications, the geological and groundwater conditions at
the job site, and the specific results desired from a blast.

Bulk ANFO is the most common explosive used under dry conditions in surface mines and
quarries. It has the advantages of the highest weight strength of the common commercial bulk
explosives combined with being economical, safe to handle and providing high charging rates.
It has no resistance to water therefore should not be used in damp or wet conditions.

ANFO has been replaced by Heavy ANFO's in some operations due to the reduction in drilling
requirements that can be realised through the use of higher bulk strength explosives. Heavy
ANFO's have an improved water resistance but should not be used in wet blastholes. An
analysis of the increased costs of the Heavy ANFO explosives compensated by the reduction in
drilling should be conducted before test blasts containing Heavy ANFO's are implemented.

In wet conditions water gel or emulsion explosives are used. Each explosive type is best suited
to different rock conditions. In general, water gel explosives are best suited to softer and
highly structured rock and the emulsions perform better in harder and more massive rock. The
performance of both explosive types depends on how they are made and their cost must also
be considered when choosing between them.

Priming System

The effectiveness of an explosive in priming or initiating another explosive is a function of the


amount of energy that is transmitted into the second explosive and of the rate at which this
energy is delivered. The detonation pressure of an explosive is defined as the pressure exerted
at the detonation front as the explosive reacts. Detonation pressure is the main factor which
governs priming effectiveness.

Explosives that are readily initiated by a detonator and which have a high density and VOD are
the most effective primers.

For any type of primer, there is a minimum weight of explosive capable of imparting sufficient
energy to start a self-generating detonation front in another explosive charge. Where priming
is marginal, there is a risk of producing a low transient VOD and reduced blasthole pressures in
the run-up zone. It has also been established that larger blasthole diameters require larger
primers because shock wave reflections do not contribute as much to the developing
detonation front. Even if the minimum primer weight is employed, the same steady-state VOD
will be achieved after a run-up distance of 2 to 8 blasthole diameters, and in many blasting
situations there may be no noticeable reduction in overall blast performance. In general,
recommendations on the size of primers take into account the non-ideal nature of detonations
in production blasting, and a factor of safety is included to reduce the risk of poor performance
and misfires.

Factors which may reduce priming effectiveness and which require a factor of safety are listed
below:

• contamination, dilution, non-ideal mixing of explosives;


• excessive sleep times in the blasthole;
• presence of ground water;
• primer partially buried by drill cuttings or mud;
• hydrostatic pressures in deep blastholes;
• desensitisation by dynamic pressures transmitted from adjacent earlier-firing
blastholes;
• water occlusion around the primer.

Detonators represent a much more concentrated energy source than detonating cord and,
consequently, initiate primers with greater efficiency. Ultra high-speed photography has
established that detonation of the primer composition begins within a few millimetres of the
base charge of the detonator, after which the detonation propagates through the remainder of
the primer. Thus, virtually all of the potential energy of the primer is available for initiating the
surrounding bulk explosive.

Whilst common practice in mines is to use a detonator to initiate the primer, for those
operations that still employ detonating cord in blastholes the following points apply.

• Detonating cord downlines affect the explosive columns through which they detonate.
With ANFO, a downline can cause either detonation by side initiation, or compression
and desensitisation of the explosive around the cord.
• High-energy detonating cords tend to side-initiate ANFO in all blasthole diameters. The
initial VOD adjacent to the cord is low, with the detonation front accelerating as it
propagates across the charge. As a result of this, the average VOD of the overall charge
is lower than the steady-state VOD of an end-initiated ANFO charge.
• For a given blasthole diameter, there is a critical core load below which detonating
cord will not side initiate ANFO. Instead, detonation of the downline creates an
expanding shock wave and gas chimney in the charge. In large-diameter blastholes,
these lateral pressures cause local compression and desensitisation of the ANFO. The
cross-sectional area of the unaffected ANFO is still sufficient to allow the detonation to
propagate along the ANFO column, but there is some energy loss. In small to medium-
diameter blastholes, detonating cords side-initiate ANFO, severely desensitise it, or
dead-press it and cause complete failure of the detonation.
• In general, emulsion explosives are less sensitive to initiation than ANFO, and they are
generally not prone to side initiation by most detonating cords. The sensitivity of an
explosive increases with temperature, however, and some bulk explosives which have
been charged into blastholes at high temperature may remain sensitive to side
initiation by high-energy detonating cord until they have cooled.
• Emulsion explosives are more easily compressed and desensitised than ANFO, because
of the action of the shock wave on the gas bubbles within these explosives. Until the
bubbles re-form after compression, high-order detonation is prevented from
propagating beyond a certain distance from the primer, resulting in failure of the
explosive.

In general, the effects of a single detonating cord on bulk explosives are known to be
influenced by:

• the weight of PETN per metre of detonating cord;


• the type and quantity of covering around the detonating cord;
• the sensitivity to initiation of the explosive;
• the method of sensitisation of the explosive;
• the initial density of the explosive when it is charged into blastholes;
• the degree of confinement of the explosive (a function of the properties of the rock
and stemming material);
• the diameter of the blasthole;
• the location of the detonating cord in relation to the axis or wall of the blasthole;
• the temperature of the explosive.

For these reasons, non-electric detonators are recommended for use in blastholes, as they
have no effect on the explosive prior to the detonator firing.

Experimental work carried out in small (<100 mm) diameter blastholes in a benching-type
operation has shown that the peak strain in the rock at the primer level increases by about
37% when the primer is moved from the bottom of the charge to bench floor level. This
increase in peak strain is due to the simultaneous detonation of those parts of the charge
which are equidistant from the primer (see Figure 1). The improved blasting results with the
primer at bench floor level are especially noticeable where toe burdens are excessive.
Figure 3. Increasing peak strain by priming at bench floor level.
Similarly, more peak strain is generated in any relatively thick bed/band of strong strata in the
face by locating a primer centrally within the bed/band rather than well above or below the
bed/band.

Some operators place the primer at a known distance above or below bench floor level to
ensure that, should a misfire occur, the excavator operator does not dig directly into the
primer/detonator. This may be a valid reason for not placing the primer at bench floor level.

The advantages of bottom compared to top priming include:

• improved fragmentation, displacement and muckpile looseness;


• reduced toe problems, better floors and cleaner faces;
• reduced noise, airblast, flyrock and surface overbreak;
• fewer cut-offs and misfires.

Blasthole Parameters

Blasthole Diameter

In surface mine and quarry operations, blasthole diameters usually range from about 75 mm (3
inches) for percussion drills to as large as 381 mm (15 inches) for large rotary drills. The
optimum diameter is larger for higher benches and for larger digging, hauling and crushing
equipment. Large diameter blastholes are less suitable

• in strong, sparsely fissured rock


• where it is very important to control blast-generated vibrations

At large surface mines, the total cost of mining is usually minimised by drilling large-diameter
blastholes. With larger-diameter blastholes, costs are lower for

• drilling (per cubic metre or tonne of rock)


• primers and initiators
• labour for priming through to firing

Provided that a drill is not operating at the top end of its diameter range in hard strong rock,
the overall cost of mining is often lower with larger-diameter blastholes.

But with larger-diameter blastholes, costs are higher for explosives. Higher powder factors are
required to achieve the same degree of fragmentation, especially in strong, sparsely fissured
rocks. Where the rock is difficult to break, smaller-diameter blastholes have the advantage of
better distribution of explosion energy throughout the mass of rock to be blasted. When the
blasthole diameter is increased and the explosion energy factor remains constant, the larger
blasthole pattern generally gives coarser fragmentation.

Where joints or bedding plane partings divide the burden into large blocks (see Figure 1),
acceptable fragmentation is often achieved only where each block is intercepted by a
blasthole. This usually necessitates the use of smaller-diameter blastholes and correspondingly
smaller blasthole patterns.

Figure 1. Effect of using (a) large-diameter and (b) small-diameter blastholes


in sparsely jointed rock.
In rocks which have closely spaced fissures, fragmentation tends to be structurally controlled.
For this reason, increases in blasthole diameter can be accommodated with only a small
deterioration in fragmentation.

Smaller-diameter blastholes also give better top breakage, since the charges can be brought up
higher in the blasthole (see Figure 2). This better charge distribution is an appreciable
advantage in blocky rocks. With 229 mm (9 in) to 381 mm (15 in) diameter blastholes, 4.5 m
(15 ft) to 7.5 m (25 ft) long stemming columns must usually be employed if excessive air
vibration or flyrock is to be avoided. With blasthole diameters of 89 mm (3.5 in) to 127 mm (5
in), stemming length can be reduced to as little as 1.8 m (6 ft) to 2.5 m (8 ft).
Figure 2. Effect of using (a) large-diameter and (b) small-diameter blastholes
in sparsely jointed rock.
Where vertical blastholes have a larger diameter, there is a greater probability of avoiding
excessive toe burdens in the front row. Excessive front-row toe burdens are common where
small-diameter blastholes are drilled in high benches with shallow-dipping faces. Where the
front-row toe burden is not broken and displaced sufficiently, the blast often fails in the front
row; it may not recover, irrespective of the number of rows of blastholes.

Blasthole Inclination

In production blasts in surface metal mines, blastholes are usually vertical. This is because

• angled blastholes are more difficult to drill


• some drills do not have an angled drilling capability
• drilling accuracy is greater with vertical blastholes

But angled blastholes provide better distribution of the explosive in the rock mass and are very
effective in overcoming toe problems and reducing surface overbreak (see Figure 3). Angled
blastholes also give greater displacement and muckpile looseness. With angled blastholes,
fragmentation is usually better because of the more efficient use of explosion energy and
because of the reduced volume of rock alongside the stemming column (where boulders
frequently originate - see Figure 3).
Figure 3. Less surface overbreak and collar rock volume with angled
blastholes.

Figure 4. Excessive toe burden for vertical blasthole adjacent to a high or


shallow-dipping face.
The use of vertical front-row blastholes often results in a considerable variation in burden
between the top and bottom of the charge (see Figure 4). This variation is greater where the
face is high or shallow-dipping. Front-row blastholes are often collared near the crest so as to
control the toe burden. But then, of course, explosion gases may blow out prematurely in the
upper face, causing noise, airblast and flyrock (see Figure 5a). The rate at which such venting
reduces blasthole pressure near bench floor level may be sufficiently great to prevent
adequate breakage and movement of the toe. This effect is more pronounced for top or
multiple-primed charges than for bottom-primed charges.
Figure 5a. Airblast and flyrock caused by inadequate burden alongside top of
charge.

Figure 5b. Airblast and flyrock caused by inadequate burden alongside base
of charge.
Where a vertical blasthole is drilled at the design burden distance from the crest, on the other
hand, hard immovable toe can be expected. One of the major advantages of angled blastholes,
therefore, is the greater uniformity of burden throughout the length of the blasthole. Ideally,
the blasthole should be parallel to the face.

Where drilling difficulties or an increase in caving of walls of blastholes do not preclude their
use, there may be good reasons to go to the extra trouble of drilling angled blastholes. Where
faces are high, angles of 20o to 30o (from the vertical) are recommended. Angles greater than
30o are seldom used because of excessive bit wear and difficulties in maintaining blasthole
alignment and in charging.

The potential advantages of angled blastholes can be realised only where drilling is carried out
with a high degree of accuracy. If the angle (to the vertical) of front-row blastholes is too large,
the toe burden will be too small, and excessive flyrock, airblast and noise could well emanate
from that part of the face immediately above bench floor level (see Figure 5b).

Where blast designs feature vertical blastholes, angled front-row blastholes may sometimes be
necessary to control the toe burden for front-row blastholes.

The drilling of vertical blastholes does not eliminate the need for accuracy in blasthole
inclination. In a 15 m (50 ft) high bench, 1o errors in blasthole inclination can result in 0.5 m
(1.5 ft) errors in toe burden or toe spacing. An absolute error of this magnitude may represent
a significant percentage error, and may cause sub-optimum blasting results just above bench
floor level.

Direction of Dip of Blasthole

Another potential problem of angled blastholes, one that is less obvious, is the need for
correct alignment. Although generally less influential than errors in blasthole inclination, errors
in alignment are common. If one takes the (unlikely but possible) example shown in Figure 6,
where front-row blastholes are drilled at the correct angle (to the vertical) but with their dips
in the wrong direction, the following errors arise.

Figure 6. Errors in toe spacings and toe burdens caused by errors in the
direction of dip of angled blastholes.
• The toe burdens on blastholes A, B and C are too large.
• The toe spacing between blastholes A and B is too small and that between B and C too
large.

Wherever blastholes are inclined, therefore, there must be an accurate and reliable system of
aligning the drill to achieve blasthole dips in the design direction.

Effective Subdrilling

Efficient digging operations require the fragmentation and displacement of the rock at bench
floor level to exceed certain critical values. Toe conditions are influenced appreciably by the
amount of effective subdrilling.

Effective subdrilling is the length of that part of the explosive charge which lies beneath bench
floor level. Subdrilling is a less important blast parameter and is the drilled length of blasthole
beneath bench floor level. Because unavoidable fallback (of drill cuttings and small rock
fragments) is always finite, effective subdrilling is always less than subdrilling. It is effective
subdrilling, not subdrilling, that influences toe conditions. It is recognised good practice to drill
a certain extra distance (which is longer for higher benches and weaker rocks) to allow for
unavoidable fallback.
The optimum effective subdrilling varies with

• rock mass properties


• the type of base charge (and, more particularly, the energy generated per metre of
blasthole)
• the diameter and inclination of the blasthole
• the effective burden distance
• the location of primers in the charge

If effective subdrilling is insufficient, toe (or high bottom) will result. Blasting under such
conditions usually sets a vicious cycle in motion which results in more overbreak and requires
even heavier concentrations of explosion energy in the bottom of the blasthole.

In strong massive rocks or steeply dipping formations, effective subdrilling of about 8d (where
d = blasthole diameter) is usually found to be satisfactory. Where vertical blastholes are drilled
alongside high or shallow-dipping faces, effective subdrilling of 10d or even 12d may be
necessary in front-row blastholes because of the heavy toe burden (see Figure 7). Effective
subdrilling shorter than 8d can often be used satisfactorily where:

Figure 7. Greater effective subdrilling to remove excessive toe burden.


• a very high energy per metre of blasthole can be generated
• there is a distinct shear or separation plane at or near bench floor level
• blastholes are inclined significantly to the vertical

Even where there is no change in blasthole diameter, effective subdrilling may need to be
increased with any increase in burden or spacing.

Where pronounced bedding plane partings parallel the bench floor, as is the case in most
sedimentary formations, very little, if any effective subdrilling is necessary. If an open fissure
(eg, bedding plane parting) exists at floor level (see Figure 8), effective subdrilling may, in fact,
be a disadvantage, since too much gas expansion energy could be released into this plane, thus
producing excessive spread of the muckpile.
Figure 8. Influence of horizontal bedding on effective subdrilling.
Too much effective subdrilling is to be avoided because it leads to

• a waste of drilling and explosives expenditure


• an increase in ground vibration
• undesirable shattering of the bench floor (which may create drilling problems,
abandoned blastholes and, hence, off-design patterns on the bench below)

The last of these tends to cause increases in both cut-off and overbreak problems.

Blasthole Pattern

Type of Blasthole Pattern

Blasthole patterns vary greatly and depend on blasthole diameter, rock properties, explosive
properties, the degree of fragmentation and displacement required, as well as the bench
height.

Confirmed blast modelling results and considerable operating experience have shown that
fragmentation and productivity are generally greater with staggered patterns than with either
square or rectangular patterns. The difference between these two types of pattern increases
with the mean fissure spacing and strength of the rock.

In strong massive rocks, the best staggered patterns are those that are based upon equilateral
triangular grids. Equilateral triangular patterns provide the optimum distribution of explosion
energy in the rock to be broken. In general, staggered patterns give the best performance,
followed by rectangular and then square patterns. However, initiation sequence can alter the
performance of these patterns considerably.

Where equilateral triangular patterns are used, the drilled blasthole spacing (S) is equal to the
drilled burden distance (B) multiplied by 1.16 (ie, S = 1.16B) for most situations.
Burden Distance and Blasthole Spacing

Experience has shown that the explosive charge is most efficient when the burden is equal to
approximately 25 to 35 times the blasthole diameter, depending on the explosive used and the
properties of the rock. Where optimum fragmentation is required in ground which tends to
break coarsely, conservatively small values of burden distance (B) and blasthole spacing (S)
should be selected. Where good fragmentation is less important, or when blasting friable
ground which tends to be broken easily, satisfactory results may be obtained by drilling larger-
diameter blastholes on a correspondingly larger pattern.

Blasthole spacings appreciably smaller than the burden distance tend to cause premature
splitting between blastholes and early loosening of the stemming. Both of these effects
encourage the premature release of explosion gases to the atmosphere, and overbreak is
usually considerable. This loss of heave energy detracts from overall breakage, and large slabs
are often found in the muckpile.

On the other hand, where the drilled S:B ratio is too large, the face region midway between
back-row blastholes may remain intact, especially near bench floor level, where the
unsuitability of the spacing would result in tight digging and possibly unbroken toe.

As the burden distance increases beyond its optimum value

• fragmentation becomes coarser, the muckpile tighter, and digging slower and more
costly and
• overbreak, ground vibrations and the instability potential of pitwalls increase
Special attention should be paid to the positions of front-row blastholes. If the burden on
front-row charges is excessive, it will not be broken and detached from the rock mass by the
time second-row charges detonate, and this restriction of motion at the beginning of the blast
will prevent optimum blasting results being achieved throughout. Where B is too small,
explosion gases burst rapidly through the face, causing noise and airblast, and in some cases
flyrock.

Changes in B generally affect fragmentation, muckpile looseness and the presence of toe much
more rapidly than changes in S. If blasting results are more than adequate and it is decided to
try a larger pattern, S should usually be increased rather than B. If the current pattern is
already quite elongated, however, it may be necessary to keep S constant and increase B by a
reasonable amount (10% maximum). When modifying B and S, it should be remembered that a
S:B ratio of 1.16 gives the optimum distribution of explosion energy in the rock mass. With any
trial blast, it is preferable that the muckpile be completely removed and the cost-efficiency of
digging, hauling and crushing assessed before the next blast is drilled out.

Normally, B and S are related to blasthole depth and, more particularly, to charge length. Table
1 above can be used as a guide for multi-row blasts of 89 to 349 mm diameter blastholes in a
rock mass of average strength.

Once the pattern has been designed, it is most important that the blastholes are drilled in the
correct place, at the correct angle and to the correct depth. There is a very good reason to
have the face surveyed and every blasthole measured out carefully. Optimum blasting results
can be achieved only where the correct burdens and spacing are selected and then
implemented.

Compared to free-face blasts, choked blasts and drop cut (ie, sinking cut) blasts require a
higher effective energy factor to achieve comparable fragmentation and particularly muckpile
looseness. Such an increase in energy factor is usually achieved by reducing B and/or S.

Effective Burden and Effective Spacing

The effective burden Be and the effective spacing Se depend not only upon the blasthole
pattern but also upon the sequence of firing. As illustrated in Figure 1, a square blasthole
pattern which is fired row by row from the face gives an effective burden equal to the spacing
between successive rows parallel to the face. On the other hand, an identical pattern of
blastholes can be fired en echelon resulting in completely different burdens and spacings.
Figure 1. Effective burdens and spacings for blasthole patterns (Hoek & Bray
(1977))

Stemming

The use of stemming to fill the blasthole above the explosive column is a generally accepted
procedure for directing explosive effort into the rock mass. The same arguments as were used
in the discussion on burden apply in the case of stemming. Too little stemming allows the
explosion gases to vent and generate flyrock and air blast problems as well as reducing the
effectiveness of the blast. Too much stemming gives poor fragmentation of the rock above the
top of the explosive charge.

The following video illustrates the results of proper stemming design.

Video 1: An example of proper stemming design.


Type of Stemming

Drill cuttings are the most convenient and usual means of stemming blastholes. However, it
has been found that angular crushed rock containing a well graded distribution of particle sizes
ranging from fine to coarse is a much more effective stemming material. In situations where
control of airblast is critical, consideration could be given to the use of angular crushed rock
stemming.

For blasthole diameters in the 76 to 381 mm (3 to 15 in) range, (angular) crushed rock in the
approximate size ranges of 5 to 10 mm (1/4 to 3/8 in) and 10 to 20 mm (3/8 to 3/4 in) make
very effective stemming material. When impulsively compressed at its lower end by the
tremendous pressure of the explosion gases, crushed stone arches, wedges into the blasthole
walls and provides better interlocking and confinement of the gases than (relatively fine) drill
cuttings. This confinement helps to maintain peak pressures in the blasthole for a longer
period of time. The longer these gases can be bottled up, the more they will be able to
fragment, heave and loosen the ground. Where blasthole water exists above the charge,
coarse angular stemming also sinks more readily to its intended position. Unlike drill cuttings,
it does not form a soupy suspension.

Length of Stemming

For every set of blasting conditions, there is an optimum stemming length. As stemming length
decreases below its optimum value, collar rock breakage by shock energy increases but

• overall breakage and displacement by heave energy decrease ( because explosion


gases vent to atmosphere more easily and rapidly), and
• there is a higher probability of flyrock, surface overbreak, noise and airblast.

As stemming length increases beyond its optimum value, more effective work is done on the
rock alongside each explosive charge, but fragmentation of the collar rock becomes coarser.
There is a certain stemming length beyond which further increments in heave energy
effectiveness are not achieved.
The optimum stemming length is shorter for

• smaller-diameter blastholes,
• stronger, more massive rocks,
• stemming material with a higher ejection resistance,
• explosives with lower bulk strengths, and
• benches with heights less than a certain value.

As a general rule, the stemming length should be not shorter than the burden distance (B). The
optimum stemming length, however, depends very largely on rock mass properties and can
vary from about 0.7B to 2B. Columns shorter than 0.7B generally cause noise, airblast, flyrock
and overbreak.

For strong massive rock, the stemming column should be the shortest which prevents flyrock,
overbreak or excessive noise and airblast. As a general rule, the stemming length should lie in
the ranges shown in Table 2 below. If the shortest stemming columns fail to give adequate
breakage of the collar rock, the use of small pocket charges should be considered (Figure 2
below).

Figure 2. Use of pocket charges to break collar rock.


Figure 3. Airdecks or longer stemming columns can be used in back-row
blastholes to reduce overbreak.

Where the upper burden contains closely spaced fissures and planes of weakness, relatively
long stemming columns (and correspondingly low powder factors and energy factors) can be
used. Alternatively, an airdeck can be used above the charge with a stemming column of
normal length. Airdecks or longer stemming columns can also be used in back-row blastholes
to reduce overbreak (Figure 3 above).

Relatively long stemming columns should also be used for any front-row blasthole which has
an inadequate burden alongside the top of the charge, so as to prevent flyrock and airblast.
(Such conditions are common where vertical blastholes are drilled alongside faces which are
high or shallow-dipping). Pocket charges can also be applied in such front-row blastholes.

Review #4

The randomly selected multiple-choice questions below are designed to


review your understanding of the material covered in the preceding sessions.
Your selections are lost when you leave the review page. On return the
review will start afresh with a new selection of questions.

This review is currently set to practise mode. To optimize your learning


experience you need to register for certification before entering the course.
Certification tests more rigorously, keeps track of your answers to the
multiple choice review questions, and enables you to report and submit your
review scores to complete the certification process. If you have already
registered and been approved for certification then you should Exit and re-
enter before proceeding.

Each question below has one or more correct responses. Your selection of a
response is immediately marked correct or not.

Q1. The choice of bench height influences the costs associated with drilling
and blasting; in general, an increased bench height will ...

decrease the amount of subdrilling?

increase the efficiency of drilling and blasting personnel and


equipment?

decrease the consumption of primers and initiators?

decrease the powder factor?

decrease the fragmentation of rock?

increase the optimum blasthole diameter?

Q2. In blasthole pattern design ...

burden (B) should be 15 - 30 x the blasthole diameter?

burden (B) should be 25 - 35 x the blasthole diameter?

the ratio of blasthole spacing (S) to burden (B) should be


approximately 1.16?

the ratio of blasthole spacing (S) to burden (B) should be


approximately 1.28?

Q3. Effective measures for dealing with excessive toe burden include ...

effective subdrilling of up to 12 x blasthole diameter?

effective subdrilling to a maximum of 18 x blasthole diameter?


positioning of the primer at bench floor level?

positioning of the primer at the bottom of the blasthole?

Q4. When selecting explosives, some are better suited to certain


conditions; of the following selections, which are true for the indicated
explosive?

Bulk ANFO has high charging rates but no resistance to water.

The higher bulk strength of Heavy ANFO can reduce the overall
amount of drilling required.

Heavy ANFO has improved water resistance compared to bulk


ANFO and can be used in wet conditions.

Water gel explosives are good in wet conditions and are best
suited to softer and highly structured rock.

Emulsion explosives are good in wet conditions and are best


suited to softer and highly structured rock.

Q5. To achieve the same fragmentation when blasting into previously


blasted rock as blasting to a free face, one needs to...

decrease the blasthole spacing?

increase the blasthole spacing?

decrease the powder factor?

increase the powder factor?


Blast Design and Assessment for Surface Mines and Quarries (Text Level)
Part 5: Production Bench Blasting (cont.)

Powder Factor, Size and Initiation

Powder Factor

The powder factor (PF) is the weight of explosive in kilograms (or pounds) used to break each
cubic metre (cubic yard) of solid rock. In surface metal mines, powder factors for production
blasts generally vary from around 0.2 to 0.8 kg/m3 (0.3 to 1.4 lb/yd3) (if the powder factor is
expressed in kg/tonne, it is important to qualify the figure with a statement of material
density).

Too much importance has been placed on powder factor as a blast design criterion. Because
powder factors are defined by the weight rather than the strength of the explosive, a
comparison between different products may not be possible.

From experience gained over a period of time, a shotfirer often knows how much explosive of
a given type is needed to blast a particular rock type satisfactorily. In these cases, powder
factor can be used as a good design tool.

Many operators have realised overall gains in operational efficiency with increased powder
factors related to lower digging and maintenance costs and/or increased crusher throughout.

Size and Shape of Blast

Larger blasts produce generally finer fragmentation and result in less time being lost in
withdrawing workers and equipment from the blast area, in waiting for the blast to be fired,
and in re-entering after the "all-clear" signal has been sounded.

Coarsest fragmentation usually occurs along the perimeter of a blast (ie, along the face to
which the blast shoots and along the sides and back of the blast). In a typical bench blast, a
significant amount of oversize material is often produced at the ends of the blast where the
forward movement of blasted material tears unblasted boulders from the side faces formed.
Oversize also originates from the front-row burden in which large blocks, effectively detached
from the rock mass by the previous blast, are heaved out without being fractured or
influenced by the explosion-generated strain waves or propagating cracks.

For blasts of a given volume, therefore, there is an incentive to minimise the combined area of
its sides. This area is minimum for square (plan view) blast blocks. But where large blasts in
strong massive rock have a square shape, the large number of rows tends to result in tight
muckpiles and, hence, increasingly difficult, slow and costly digging towards the back of the
blast. This problem can be reduced, but not eliminated, by using

• a high powder factor


• relatively long delay times between dependent blastholes

Increases in mean blast size can be achieved best (with least risk of digging problems) by
increasing the mean length of blasts. If efforts were to be made to increase the mean width of
blasts, the number of rows of blastholes should be increased gradually. Several blasts with n
rows should be dug out cost efficiently before a blast with n + 1 rows is attempted. Several
blasts with n + 1 rows should be dug out cost efficiently before a blast with n + 2 rows is
trialled.

Figure 1: Long, narrow (3 row) blast showing loose, well-fragmented


muckpile.
In developing designs for blasts with larger numbers of rows, operators should be aware that
muckpiles towards the back of a wide blast are looser for

• smaller burden distances


• angled (cf. vertical) blastholes
• stemming columns which do not allow explosion gases to escape to atmosphere
prematurely
• large values of effective subdrilling
• relatively long delay times between dependent blastholes

As mean blast size increases, so does the probability that the blast block will contain two or
more rock types with different blasting characteristics. Blasting strong and weak rocks in the
same blast poses the following problems.

• Where a blasthole passes through weak and strong rock, the dimensions of the
blasthole pattern should be dictated by the properties of the strong rock.
• Where a blast block contains an area of strong rock and an area of weak rock, there is
an incentive to use two different blasthole patterns: a small pattern in the strong rock
and a large pattern in the weak rock. But this approach is complex and, therefore,
invites errors and, consequently, poor blasting results. A better approach is to use a
standard blasthole pattern (the pattern for the strong rock) and to use a less dense
explosive or shorter, lighter charges with longer stemming columns resulting in a lower
powder factor in the weak rock.
• Where a charge detonates in weak rock and very close to a weak/strong rock contact,
it breaks the weak rock well but has little effect on the strong rock. This part of the
strong rock may be inadequately broken. Similarly, where a charge detonates in strong
rock and very close to a strong/weak contact, explosion gases jet prematurely into the
weak rock, and may leave this part of the strong rock inadequately broken. For these
reasons, poor blasting results are often obtained at and near the contact between
weak and strong rocks.

With larger blasts, environmental disturbances are fewer and, when initiating charges with
non-electric detonators, the magnitude of the disturbance need be little or no greater than
that for a small blast.

Figure 2: Large blast pattern.


Figure 3: Example of complex blast pattern tie-in.

Blast size should be such that explosive charges do not "sleep" in blastholes for excessive
periods of time. Nor should the size of a blast block be increased by joining sub-blocks
together to make excessively complex shapes. Blasts with complex shapes invite errors,
especially when allocating delays.

Initiation Sequence and Delay Allocation 1

General

The outcome of any multi-hole production blast is dependent on interactions between


blastholes. The sequence in which blastholes are initiated and the time interval between
successive detonations has a major influence on overall blast performance. A poor blast design
(up to the point of initiation design) cannot be rectified by good initiation design. The
performance of production blasts can only be optimised when charges detonate in a
controlled sequence at suitable discrete but closely spaced time intervals. The result of a well-
designed multi-hole blast cannot be duplicated by firing the same number of blastholes
individually or at random.

The optimum delay allocation for a blast depends on many factors which include:
• rock mass properties (strength, Young's modulus, density, porosity, structure,etc)
• blast geometry (burden, spacing, bench height, availability of free faces, etc.)
• diameter, inclination and length of blasthole
• type and length of stemming
• explosive characteristics, degree of coupling, decking, etc.
• initiating system (surface or in-hole delays, type of downline, etc)
• type and location of primer
• environmental constraints (air and ground vibration levels and frequency)
• the desired results (fragmentation, muckpile displacement and profile etc.)

It is not possible to determine optimum delay allocations from first principles, but blast
modelling, monitoring, analysis and interpretation have led to a greater understanding of the
mechanisms and significance of blasthole interaction.

Inter-Hole Delay

The delay time between adjacent blastholes in a drilled row is commonly referred to as the
"inter-hole" delay. Firing a single row of blastholes with the optimum inter-hole delay
produces the following results.

• Fragmentation cannot be improved upon without altering geometry, explosive type or


some other variable to increase the explosion energy per cubic metre of rock.
• Forward displacement is always somewhat less than that for an instantaneous single-
row blast, but displacement and muckpile profile can be altered by changing the inter-
hole delay.
• Overbreak is similar to that produced by a single-hole blast, and a smooth wall profile
can be produced by manipulating the inter-hole delay.
• Ground and air vibration levels can be maintained at or near the level of a single-hole
blast. If necessary, seismic energy can be channelled into a more appropriate
frequency band.

In practice, there is likely to be some trade-off between fragmentation, displacement and


vibration levels. The appropriate balance for each situation can be achieved only by using the
optimum inter-hole delay, and some experimentation is usually required.

For a brittle, elastic, homogeneous rock, a short inter-hole delay is usually appropriate, whilst a
porous, plastic, closely fissured rock requires more time between the detonations of adjacent
blastholes, tending to maximise forward displacement at the expense of fragmentation and
vibration levels. Long delays tend to make each blasthole work more independently, reducing
positive interaction.

Results from a wide range of conditions indicate that the optimum inter-hole delay for
conventional blasting usually lies between 3 and 7 ms/m (1 - 3 ms/ft) of blasthole spacing (as
measured along a row or echelon). The optimum delay interval for each situation is influenced
by rock properties, blast geometry and the desired result, but a value of 5 ms/m (2 ms/ft) of
spacing is usually a good starting point.

Inter-Row Delay
The delay time between the initiation of dependent blastholes or successive effective rows of
blastholes is commonly referred to as the inter-row delay. The inter-row delay time has a
major influence on the outcome of any multi-row production blast. In many situations the
inter-row delay is just as important as the inter-hole delay in controlling overall blast
performance.

In a properly designed multi-row blast, charges adjacent to free faces have a manageable
burden of rock to break and displace. However, all blastholes in subsequent rows depend on
earlier-firing charges to create new free faces during the blast. Charges with an excessive
lateral burden tend to crater upwards towards the collar, their only alternative free face.
Blastholes which are subjected to this type of lateral confinement do not perform efficiently,
and the overall blast result is sub-optimum.

In a simple multi-row blast with a single free face, all blastholes could be initiated
simultaneously. In this event, front-row blastholes would have a finite burden and would
produce a certain amount of fragmentation, forward displacement, ground vibration, etc.
However, every blasthole behind the front row would perform less effectively because of
excessive lateral confinement at the time of firing. These charges would tend to crater
upwards, with the effect becoming more pronounced towards the back row; consequently,
fragmentation and forward displacement would suffer, whilst overbreak, ground vibration
levels and airblast would increase. The resulting muckpile would become progressively more
difficult to dig towards the sides and back, where the rock may be virtually unmoved towards
the toe.

The following video clip provides an example of production blasting with delay sequencing.

Video1: Production blasting with delay sequencing.

A multi-row blast can be fired to a free face in a row-by-row sequence by introducing a time
delay between the detonation of successive rows of blastholes (see Figures 1a and 1b). If this
time delay is adequate, overall blast performance may increase significantly because
progressive relief of burden is provided. This ensures that each row of blastholes has an
effective sub-vertical free face to which to shoot, because the preceding row has already
broken and displaced its burden. However, if the delay time between successive rows is not
sufficient to provide proper relief, later-firing rows become progressively more crowded.
Consequently, blast performance will not be optimum, and results will tend to deteriorate
towards the back of the blast.

Figure 1a. Initiation sequences - square blasthole pattern.

Figure 1b. Initiation sequences - staggered blasthole pattern.


The use of a square blasthole pattern and row-by-row initiation leads to a highly unfavourable
blasting geometry for side and corner blastholes (see Figure 1a), which causes tight digging
and excessive overbreak in these areas. The use of a staggered pattern with row-by-row
initiation produces finer fragmentation (and possibly looser muckpiles), as explosives are more
evenly distributed and perimeter blastholes have a more favourable breakout angle (see
Figure 1b). For most operations, however, the result will not be optimum, and may be
unsatisfactory unless blasting is used to simply loosen the rock mass before digging.
Figure 1c. Initiation sequences - standard square 'V' pattern.

Figure 1d. Initiation sequences - standard staggered 'V' pattern


Figure 1c (above) shows a standard square 'V' pattern. Figure 1e (below) shows a square 'V1'
pattern. Figure 1d (above) shows the standard staggered 'V' pattern; this is, in fact, a tighter
firing configuration than that shown in Figure 1c.
Figure 1e. Initiation sequences - square 'V1' pattern.

Figure 1f. Initiation sequences - staggered 'V1' pattern.


Blast performance can usually be improved significantly by initiating staggered blasts in a 'VI'
sequence (see Figure 1f (above)). Drilled rows of blastholes are split into effective rows which
have an apparent reduction in burden. Progressive relief of burden is enhanced by a longer
time delay per metre of burden between effective rows, reducing crowding towards the back
of the blast. This initiation sequence leads to a reduction in geometric confinement for side
and corner blastholes, promoting forward displacement and reducing overbreak.
Fragmentation is improved by the progressive development of suitable effective free faces
during the blast, and the positive interaction of adjacent charges. Vibration levels are reduced
because of progressive burden relief, whilst side and corner holes are less constrained and the
blast is spread over a longer period of time.

The optimum inter-row delay affects blast performance in the areas listed below.
• Diggability is enhanced, particularly in the toe region of blastholes towards the back of
the blast. The optimum inter-row delay ensures that each blasthole has an effective
free face to which to break. This occurs because the earlier-firing blasthole has broken
its burden and detached it before the dependent blasthole fires. This progressive relief
of burden during the blast affects the amount of oversize rock produced, although
fragmentation is often influenced more by the inter-hole delay than by the inter-row
delay.
• Lateral movement and muckpile profile can be optimised by manipulating the inter-
row and inter-hole delays to control the extent and direction of rock displacement.
• Overbreak at the back and side(s) of the blast is minimised because forward
movement is promoted by providing an adequate time delay between the detonation
of dependent blastholes.
• Ground and air vibrations are controlled and can often be maintained at levels similar
to those produced by a single-row blast. This is a direct result of progressive relief of
burden, which promotes lateral movement and minimises uplift, cratering and
stemming ejection. Subsequent blasts are also likely to be quieter because reduced
overbreak means that front-row burden rock contains minimum cracking from the
previous blast. Correctly matching inter-row and inter-hole delays may also permit
seismic energy to be channelled into more appropriate frequency bands.

The optimum inter-row delay for a specific application cannot be calculated from first
principles. In any situation, some experimentation is recommended. However, results from a
wide range of conditions indicate that the optimum inter-row delay for conventional blasting is
usually in the range of 10 to 20 ms/m ( 3 to 6 ms/ft) of effective burden (measured between
dependent blastholes or successive rows). The optimum inter-row delay for each situation is
heavily influenced by rock properties, blast geometry and the desired result, but a figure of 15
ms/m (5 ms/ft) is usually a good starting point.

For a brittle, elastic, closely fissured rock , a relatively short inter-row delay is usually
appropriate. A porous, plastic, dense, homogeneous rock requires more time for burden
movement. Long delays encourage forward displacement and muckpile looseness. Short
delays tend to restrict lateral movement (with the possibility of blasthole cut-offs), reduce
diggability and cause higher ground vibrations.

Initiation Sequence and Delay Allocation 2

In-Hole Delays

Where a surface delay system is used to fire instantaneous downlines (say detonating cord)
with no in-hole delays, a blasthole detonates immediately after its downline has been initiated.
This will disrupt the surrounding rock mass and cause ground movement which may damage
or cut unfired trunklines and downlines in the vicinity. The probability of this occurring is
higher where relatively long surface delays are used between adjacent blastholes, and in many
situations the maximum practicable delay time is limited by the need to prevent excessive
misfires from this cause. In massive rock, in which high energy factors and short stemming
columns are used, the allowable delay time may be significantly shorter than that required for
optimum blast performance. As a general guide, the maximum inter-hole delay should not
exceed 6 ms/m (2 ms/ft) of effective blasthole spacing if downline cutoffs are to be avoided,
unless in-hole delays are used.

In-hole delays provide a time interval between the initiation of each downline (either
detonating cord or nonel (?)) and detonation of the corresponding explosive charge. Provision
of a suitable time delay ensures that the initiating signal has reached the detonator(s) within
each charge before it or adjacent charges begin to disrupt the surrounding rock mass. This
minimises the probability of downlines being physically damaged or cut off by ground
movement during the blast, and permits the use of longer inter-row delays, which are often
essential for optimising blast performance.

The 4:1 Ratio

Where a single in-hole delay is used, the in-hole delay should be 3 to 5 times the longest
surface delay. A ratio of in-hole to surface delay in this range provides a practical balance
between two conflicting factors.

• Providing sufficient in-hole delay to avoid surface downline cut-offs due to ground
movement during the blast. If surface delays are used with no in-hole delays (ie, the
ratio is zero), there is a significant risk of cutoffs.
• Providing reliable blasthole sequencing. If an extremely long in-hole delay is used with
short surface delays, firing time variations within the in-hole delays may be large
compared to the surface time. The probability of blastholes firing out of sequence is
minimised with a ratio of 3:1 to 5:1 for in-hole to surface delays, because delay time
variations within the in-hole delays are small compared to the most critical surface
delay interval. A higher ratio increases the probability of blastholes firing out of
sequence.

Consistent results are obtained by using in-hole delays which are approximately 4 times the
longest surface delay. A smaller ratio may be acceptable in favourable conditions (e.g. where
blasts shoot to a free end with a low energy factor and long stemming columns). Under severe
conditions, however, a larger ratio may be chosen to increase the burning front between the
surface delay system and the subsequent detonation of charges (eg, in strong massive rock
with a high energy factor in a drop cut with no free face). In-hole delay selection for 114 mm
(4.5 in) diameter blastholes in a strong rock is illustrated by the following example:

Surface delays (based on experience):

Inter-row delay = 42 ms

Inter-hole delay = 17 ms

Therefore, design in-hole delay = 4 x 42 ms = 168 ms.

A 175 ms non-electric detonator would be appropriate. For double priming, use a 175 ms delay
in the bottom primer and a 200 ms delay in the top insurance primer.
As inhole delay elements become more accurate the inhole delay time can be increased. A
commonly used inhole delay time is 350 ms. If much longer delay times are used the scatter in
their initiation times may cancel out the effect of the surface delays.

Hole-by-Hole Initiation

In VI blasts, blastholes in each effective row fire virtually simultaneously. This rarely produces
optimum blast performance, especially in terms of fragmentation or ground vibrations. In
many cases, blasting results can be improved by introducing hole-by-hole firing, where every
blasthole is initiated in sequence at a unique time. Where appropriate delays are selected,
hole-by-hole initiation exploits the positive benefits of blasthole interaction whilst avoiding
most of the negative effects. This can lead to finer fragmentation, looser muckpiles, less
overbreak, lower ground vibration, and better control over the position and profile of the final
muckpile.

Hole-by-hole firing can be achieved by using in-hole detonators of various delay times.
However, the available range of times of commercial in-hole delays is limited and, hence, it is
rare to see large production blasts fired with sequencing controlled by in-hole delays. In
practice, hole-by-hole initiation is usually achieved by using a surface delay system to control
blasthole sequencing.

If relatively long inter-row or inter-hole delays are required to produce the desired results, a
combination of surface and in-hole delays will be required to avoid downline cutoffs caused by
ground movement during the blast. Where this combination system is used, each blasthole
usually contains the same in-hole delay (see Figure 1). Inter-hole delay times are controlled by
the surface initiating system, whilst the in-hole delay provides a safety factor against potential
trunkline or downline cutoffs caused by ground movement during the blast.

Figure 1. Hole-by-hole firing using both surface and in-hole non-electric


delays.
Delay Time Interaction

The performance of a conventional blast can be influenced significantly by altering delay


timing to change the degree of interaction between adjacent blastholes. Whilst absolute
values of inter-row and inter-hole delays are important, the ratio of these times also has a
major impact on the final result. This can be explained by the following concepts (which are
over-simplifications of a complex subject).

• The inter-hole delay controls interaction between adjacent blastholes and determines
whether blastholes act independently or together.
• The inter-row delay controls interaction between dependent blastholes, as it affects
the progressive creation of new effective free faces for later-firing blastholes.
• The ratio of inter-row delay to inter-hole delay controls the geometry and orientation
of new free faces created as the blast progresses. For a later-firing blasthole, the
location, shape and extent of any effective free face depends on this ratio of delay
times. This influences the direction and extent of displacement of each blasthole's
burden and thus the final muckpile shape and position. This is sometimes referred to
as the apparent direction of movement of a blasthole or the overall blast.

The manipulation of delay timing to control blast performance is best illustrated by considering
several alternatives for initiating the same blast. Where blastholes are initiated row-by-row,
forward displacement is enhanced and the general direction of movement is perpendicular to
the rows. If the same blast is initiated in a 'V' pattern, the direction of movement will be
oblique to the effective rows or towards the free corner (see Figures 2 - 5 below).

Figure 2: Blast Tie-in with 17ms Inter-hole and 65ms Inter-row Delays.
Figure 3: Detonation Simulation for Blast with 17ms Inter-hole and 65ms
Inter-row Delays.
Figure 4: Blast Tie-in with 65ms Inter-hole and 17ms Inter-row Delays.

Figure 5: Detonation Simulation for Blast with 65ms Inter-hole and 17ms
Inter-row Delays.

The above sequences can be used for blasting in a wide range of applications. Best results
occur where a free face and/or free end is available. Delay times can be altered to suit
different conditions, in conjunction with in-hole delays if necessary.

The above sequences are best achieved by using a non-electric initiating system.

Evaluation and Modification

Evaluation of a Blast

Once the dust has settled and the fumes have dispersed after a blast, an inspection of the area
should be carried out. The main features of a satisfactory blast are illustrated in Figure 1.

The front row should have moved out evenly but not too far. Excessive throw is unnecessary
and very expensive to clean up. The heights of most benches are designed for efficient
excavator operation; low muck piles, due to excessive front-row movement, represent low
excavator productivity.
The main charge should have lifted evenly and cratering should, at worst, be an occasional
occurrence. Flat or wrinkled areas are indicative of misfires or poor delaying.

Figure 1. Features of a satisfactory production blast (Hoek & Bray (1977)).

The back of the blast should be characterized by a "power trough", indicating good forward
movement of the blasted rock. Tension cracks should be visible in front of the final diglines.
Excessive cracking behind the final digline represents damage to the slopes and wastage of
explosive's energy.

The quality of a blast has a significant effect on the cost of secondary drilling and blasting,
digging rate, the condition of the haul roads, loader/shovel and truck maintenance, etc.
Therefore, careful evaluation of the blast to determine how improvements could be made to
the design are very worthwhile.

Oversized fragments, hard toes, tight areas and low muckpiles (caused by excessive throw)
have the most significant detrimental effect on the digging rate and digging conditions. A study
of loader or shovel performance and of complaints from operators helps to maintain an
awareness of these problems among blasting personnel. An attempt should be made to
measure digging rate by noting the time required to fill loader/shovel buckets or by comparing
average daily production rates. Similarly, loader/shovel wear and tear should be noted since
this may reflect difficult digging conditions.

Poor fragmentation of the toe due to an excessive toe burden can lead to poor digging and
uneven haul road condition. Uneven haul roads lead to suspension and tire wear on the trucks
and also spillage which can cut the truck tires. An attempt to correct this problem by additional
subdrilling rather than by correcting the front-row charge can lead to excessive sub-break
which can give rise to blasthole instability and also to poor bench crest conditions if a lower
bench is to be removed.

Modification of Blasting Methods

When it is evident that unsatisfactory results are being obtained from a particular blasting
method and that the method should be modified, the engineer may have to embark on a
series of trials in order to arrive at an optimum design. As with any trials, careful
documentation of each blast is essential and, whenever possible, only one variable at a time
should be changed. The following sequence of test work is an illustration of the type of
experiment which would be carried out to evaluate the cost effectiveness of using a higher-
energy explosive. Similar test sequences could be carried for each of the other factors which
are relevant in a particular situation.

a) Rationalization - Document present powder factors on an equivalent energy basis using the
weight of the explosive in current use. Weight strength data should be obtained from the
explosives manufacturer if these are not already available.

b) Evaluation - For a blast with the explosive currently in use, document the behavior of the
blast during initiation and the condition of the resulting muck pile.

c) Document rate and conditions of digging.

d) Document fragmentation based upon the ratio of oversized material requiring secondary
blasting to the total blast tonnage.

e) Document drilling and blasting costs.

f) Experimentation - Select a similar area of ground and carry out a blast with a higher powder
factor which is obtained by using a higher energy explosive or a smaller blasthole pattern.

g) Evaluation - Document the results as for steps (b) to (e).

h) Carry out a cost-benefit study.

Repeat the experiment to determine its validity.

Review #5

The randomly selected multiple-choice questions below are designed to


review your understanding of the material covered in the preceding sessions.
Your selections are lost when you leave the review page. On return the
review will start afresh with a new selection of questions.

This review is currently set to practise mode. To optimize your learning


experience you need to register for certification before entering the course.
Certification tests more rigorously, keeps track of your answers to the
multiple choice review questions, and enables you to report and submit your
review scores to complete the certification process. If you have already
registered and been approved for certification then you should Exit and re-
enter before proceeding.

Each question below has one or more correct responses. Your selection of a
response is immediately marked correct or not.

Q1. Which of the following are characteristics of a good blast?

A lot of cratering in the middle of the blast.

Front row moved evenly but not too far.

Tension cracks visible in front of final diglines.

Q2. The optimum delay allocation for a blast depends on many factors
which include...

rock mass properties?

type and length of stemming?

environmental constraints?

Q3. Which of the following is (are) true of inter-row delay?

It refers to the delay time between adjacent blastholes in a row.

It refers to the delay time between the initiation of dependent


blastholes or successive effective rows of blastholes.

The row furthest from the face is initiated first to ensure the
burden for the next row is managable.

Q4. Which of the following is (are) true?


The inter-hole delay controls interaction between adjacent
blastholes and determines whether blastholes act independently
or together.

The inter-row delay controls interaction between adjacent


blastholes and determines whether blastholes act independently
or together.

Q5. Coarsest fragmentation usually occurs along the perimeter of blasts,


and reducing the combined area of the sides of a blast will reduce the
amount of coarse fragmentation.

True?

False?
Blast Design and Assessment for Surface Mines and Quarries (Text Level)
Part 6: Overbreak Control and Secondary Blasting

Overbreak

Introduction

When the extremely concentrated energy of a detonating explosive is generated in a normal


production blast, it is not surprising that it can destroy, or at least reduce, the structural
strength of the rock around the blast block. New fractures and planes of weakness are created,
as well as joints and bedding plane partings, which may have been tight before the blast, are
opened up. As a result, there is an overall reduction in the ability of the rock mass to hold
together. This shows up as overbreak, and the fractured face (or wall) is left with a higher
likelihood of rockfalls.

Because the amount of overbreak has an appreciable influence on the steepness of the bench
and, subsequently, on the overall steepness of the excavation, it may be necessary to minimize
overbreak, especially where blasting operations are approaching the final limits of the
excavation. It will then be possible to steepen rock walls significantly. In surface mines and
quarries, steeper walls cause increases in:

• recoverable rock reserves


• the final volume of the excavation.

The successful application of overbreak control blasting techniques also lessens the hazard of
rockfalls and minimizes the need for scaling, rock bolting and/or cable bolting.

Figure 1: Coal mine highwall in horizontally-bedded strata.


Figure 2: Quarry excavation.
Figure 3: Open pit mine wall.

Overbreak Mechanisms

There are at least four mechanisms by which blast-induced overbreak can occur, viz.

• radial fracturing
• internal spalling
• the opening and possible extension of fissures and strain wave-induced cracks by
explosion gases
• release-of-load fracturing

The above mechanisms are listed approximately in their chronological order of occurrence but
by no means in their order of importance.

The overbreak resulting from mechanisms of radial fracturing and internal spalling is
influenced almost entirely by the characteristics of both the rock mass and back-row charges.
The presence of an effective free face in front of back-row charges does not prevent fracturing
by these mechanisms, the amount of breakage beyond the blast boundary being similar to that
in the burden rock.
Figure 4: Overbreak with many joints and bedding planes.
The opening of fissures and cracks, and release-of-load fracturing are usually the dominant
modes of overbreak. The overbreak resulting from these mechanisms is affected not only by
the characteristics of the rock mass and back-row charges; it is influenced to a greater degree
by certain design features of the entire blast (e.g. energy factor, effective burden distance and
delay allocation).

Overbreak Mechanisms

Radial Fracturing

When the front of an explosion-generated strain wave passes, a cylindrical shell of rock
immediately around each charge is subjected to intense radial compression, and tangential
tensile strains then develop. If these tangential strains exceed the dynamic tensile breaking
strain of the rock, a zone of dense radial fractures is produced immediately around the charge.
This intense fracturing terminates quite abruptly at that radial distance at which the wave's
tensile strain attenuates to a value which is incapable of generating new cracks.

Beyond this zone of dense radial fractures, some evenly-distributed radial cracks propagate
into the rock as long as wave-induced tension is applied normal to the crack tips. Crack lengths
in both the inner and outer radial fracture zones increase with

• an increase in the peak blasthole pressure (Pb)


• decreases in the rock's dynamic tensile breaking strain and rate of absorption of strain
wave energy
An approximately linear relationship exists between crack extension and length of pre-existing
crack; longer cracks are preferentially extended. Where a fissure intersects the blasthole over
much of the length of the charge, it opens under the action of the strain wave and limits the
development of radial fractures in other directions.

A fissure which is parallel to but some distance behind a blasthole arrests the propagation of
radial cracks. Where fissures are widely spaced, the probability of peripheral blastholes being
remote from a fissure is relatively high, and the full development of radial fractures is
promoted. But where fissures are closely spaced, they are more capable of localising radial
fracturing as a result of

• preferential extension of the fissure(s)


• termination of radial fractures

Radial cracks behind a back-row blasthole tend to produce interlocking sectors of rock beyond
the design boundary of the blast. Where radial cracks from adjacent perimeter blastholes
overlap and intersect other cracks or fissures, the newly formed face may suffer local wedge or
block failures. To minimise overbreak, therefore, the spacing of back-row blastholes should be
as large as is practicable.

Internal Spalling

Where the radial compressive strain wave strikes an effective free face (usually a pronounced
rock/air interface), a reflected radial tensile wave is created. Since the strength of rock in
tension is much lower than that in compression, tensile waves can be effective breakers of
rock. If a reflected tensile wave is sufficiently strong, spalling occurs progressively from the
effective free face back towards the blasthole.

Insofar as they affect reflection of the compressive wave, air-filled fissures beyond the
peripheral blastholes can be regarded as effective free faces. Internal spalling produces more
intense overbreak between the blasthole and such fissures but, because of reflection of the
outward-propagating strain wave and termination of radial cracks, overbreak beyond the
fissure is reduced. Breakage by this mechanism is expected to become appreciable only within
about 20d (where d = blasthole diameter) of a perimeter charge.

In rocks with closely spaced fissures, internal spalling together with radial fracturing
encourages intense overbreak between the blasthole and the fissures nearest to the blasthole.
Because fracturing beyond these fissures is not appreciable, the amount of overbreak by these
two mechanisms is small. But, as a result of the following two mechanisms, overbreak is
usually greater in more closely fissured rocks.

Opening and Extension of Fractures/Fissures by Invading Explosion Gases

Immediately after the formation of radial cracks, the explosion gases start to penetrate both
fissures and the radial cracks. Where the charged section of a blasthole does not intersect a
fissure, the high-pressure gases first jet into, wedge open and, hence, extend the radial cracks.
In the absence of fissures running along the side of the charge, this mechanism can increase
the length of strain wave-generated radial cracks by a factor of about five.
Because stress concentration at a crack tip increases with crack length, the longest crack
(whether this be a fissure or a strain wave-generated crack) is the least stable and requires the
lowest gas pressure for its further propagation. Therefore, longer fractures always extend first
and propagate at a higher velocity than shorter adjacent fractures. The further they get ahead,
the greater is the velocity difference until the shorter ones stop altogether. Where an open
fissure intersects a blasthole alongside its charge, high gas flows cause this fissure to be
preferentially expanded by a wedging action.

In thinly bedded or closely jointed rocks, the extensions of blast-generated cracks are, in most
cases, masked by those of the fissures. Indeed, fissures such as joints and bedding plane
partings frequently dominate both the nature and extent of overbreak. Because major crack
development is along such fissures, the overbreak zone has a uniform thickness only where
such fissures are closely spaced and quite evenly distributed throughout the rock mass. Where
fissures are widely spaced and distributed unevenly, fragmentation in the overbreak zone is
coarse, the overbreak zone has a variable thickness, and the profile of the newly formed face is
irregular.

High-pressure explosion gases use the radial cracks and any fissures which intersect the
charged section of a blasthole as access routes to any outer cracks produced by the strain
wave and, more particularly, to the network of fissures beyond the immediate vicinity of the
blasthole.

Figure 1: Blast damage in less massive rock.


Figure 2: Less damage in more massive rock.

The invasion and wedging action of high-pressure explosion gases can extend fissures well
beyond the blast perimeter. In the common case in which vertical blastholes are drilled in
horizontally bedded strata, the upward push exerted (by invading gases) on the sides of the
expanding horizontal fissures tends to cause rotational uplift and, hence, tensile fractures
normal to bedding. The amount of breakage by this mechanism increases

• towards the top of the bench


• with the brittleness of the rock
• with a decrease in the spacing between bedding plane partings

Release-of-Load Fracturing

Before the initial strain wave reaches the effective free face, the total energy transferred to
the rock by the initial compression of the rock is probably as much as about 60% of the blast
energy. After the compressive strain wave has passed, a state of quasi-static equilibrium exists,
the pressure in the blasthole being balanced by the strain at the blasthole wall. When the
pressure in the blasthole subsequently falls (as gases push the burden rock forwards and
escape via the stemming column), this strain energy is very rapidly relieved, rather like a
compressed coil spring being suddenly released.

In rocks with fissures which parallel back-row blastholes, behaviour of the rock behind the
blast may be likened of that of a multi-layered mass of rubber which is impacted at normal
incidence by a heavy steel plate. After contacting the surface of the rubber, the plate
progressively compresses the layers until its momentum is exhausted. The highly compressed
layers then expand, thereby accelerating the plate in the opposite direction; in propelling the
plate forwards, the layers separate from each other. Such separation explains the so-called
tension fractures so often observed behind blasts.
Where large multi-row blasts with an in-line initiation sequence are fired, all charges in a given
row act in unison to produce tensile overbreak fractures parallel to and over the length of the
back row. As one would expect, there is some reinforcement of these effects as a result of the
detonation of earlier-firing rows of charges. Large blasts of this type can cause release-of-load
cracks parallel to and up to 60 m behind the back-row charges (McIntyre and Hagan (1976)). As
is indicated by the analogy to a multi-layered mass of rubber, the rock mass fails preferentially
at sub-vertical fissures (e.g. joints). The amount of overbreak produced by release-of-load
fracturing is smaller where burden rock is heaved forwards with reasonable ease.

Figure 3: Cracks parallel to the wall after a production blast.

Redesigned Production Blasts

Redesigned Production Blasts vs. Smoothwall Techniques

Special blasting techniques such as presplitting (also known as preshearing) can produce rock
walls of great soundness and smoothness. This has been adequately proven by the often
spectacular results which are visible in some spillways and in highway and railway cuts. Such
techniques are important and, in the appropriate situations, can be very useful. Before any
attempt is made to apply presplitting or any other smoothwall technique, however, efforts
should be made to find out (by trial) to what extent production blasts of modified design can
reduce overbreak. If a high degree of success is achieved with redesigned production blasts,
then the need for smoothwall blasting methods can be eliminated. Contrary to what one might
expect, the redesign of a production blast to reduce overbreak does not necessarily reduce,
and could very well improve, breakage and looseness of the muckpile.
Reducing Overbreak in Production Blasts

Overbreak from production blasts can be reduced by observing the following guidelines (These
recommendations apply more particularly to those blasts which are close to the design
excavation limits, but can be applied wherever it is necessary to limit overbreak.)

• If a continuous column of explosive is too powerful for back-row blastholes, deck


charge the explosive and initiate the charge decks with cast boosters on a detonating
cord downline. The material between the charge decks should be dry angular crushed
rock. If bulk explosives are used then cardboard or plastic tubes can be lowered into
the blastholes and the bulk explosive loaded in them.
• In wet blastholes, an emulsion, watergel or dynamite should be used to give charges
which build up as a continuous column but which are a loose fit in the blasthole. A 55
mm (2 in) diameter cartridge in a 76 mm (3") blasthole would be a suitably decoupled
charge, as would a 65 mm (2.5 in) cartridge in a 90 mm (3.5 in) blasthole.
• The burden on the back row holes should not be greater than about 25 times the
blasthole diameter.
• The best spacing between back-row blastholes is the largest spacing that gives a
straight face. The spacing usually lies between 28 and 35 times the blasthole diameter.
In multi-row shots, blastholes should be staggered.
• Drill inclined rather than vertical blastholes at least for the last 3 - 4 rows in front of
the final wall of the excavation. Angled blastholes are particularly necessary for the
back row. Angles of 15 - 30o to the vertical are recommended.
• The length of the stemming column should be about 25 blasthole diameters.
• Effective subdrilling should never be greater than 8 blasthole diameters. Efforts should
be made to minimize effective subdrilling over a berm, especially where this is a
permanent berm and/or a haul road.
• The initiation sequence should be selected such that:
- there are minimum numbers of blastholes firing on the same delay;
- the blastholes along the back row and side(s) of the shot are detonating in a
delayed sequence.
• The staggered 'V1' pattern (see Figure 1) is recommended (for firing to both an open
face and a free end). Where blasts are fired to an open face, the pattern shown in
Figure 2 is even better than that shown in Figure 1.
Figure 1. Staggered V1 pattern for multi-row blasts to a free face using
millisecond-delay detonators (Hoek & Bray (1977)).

Figure 2. Three-row blasts in which charges are initiated by millisecond-delay


detonators located down the blastholes; three blastholes per delay (Hoek &
Bray (1977)).
• In single-row shots adjacent charges should be initiated on separate millisecond-
delays. In multi-row shots, each blasthole in the second (or a later) row should be
delayed so that the 2 or 3 closest blastholes in front of it have had time to break and
detach their burdens. The delay patterns shown in Figures 1 and 2 are recommended.

The specially designed blast shown in Figure 3 indicates a combination of features which give
reduced overbreak. This design also gives good breakage, movement and looseness of the
muckpile. The limited amount of overbreak results from:

• the small number of blastholes firing on the same delay number


• the relatively long duration of the blast
• the delayed rather than simultaneous detonation of back-row and side blastholes
• the relatively wide spacing between back-row blastholes
• the small burdens on blastholes
• the very suitable angle of blastholes

Figure 3. Production shot designed to give reduced overbreak and good


breakage, movement and muckpile looseness.
The chance of successfully using a modified production blast design without involving a
smoothwall blasting technique (e.g., presplitting) is much lower where it is very important to
maximize the soundness and stability of the rock wall.

Examples of components of redesigned production blasts are shown in the following figures.
Figure 4: Redesigned production blast.

Figure 5: Initiation design.


Figure 6: Detonation simulation.

Figure 7: Example of a tie-in for a single free face.


Figure 8: Wall control for structure parallel to pit wall.

Figure 9: Example of a revised blast design where blasting to a moderately


dipping structure.

Smoothwall Blasting
Smoothwall Blasting Techniques

Smoothwall blasting techniques are important methods of achieving rock walls that are
sounder and smoother than those that are left by redesigned production blasts. These
techniques are most successful in massive rocks and in tight horizontally-bedded formations
which are relatively undisturbed by faults, etc. In unconsolidated, weathered or highly-fissured
ground, problems in obtaining a consistently smooth wall are usually encountered. The very
technique that produces the desired effect in a strong massive rock may be entirely unsuitable
in weak, highly-fissured ground. Blasthole patterns and charge loads, therefore, will change
appreciably with the rock properties encountered.

Figure 1: Pre-split blast in massive sandstone.


Because of the greater amount of drilling required, smoothwall blasting techniques (described
below) are more costly than regular or redesigned production blasting. It is essential that
drilling be carefully supervised so that blastholes are at the design spacing, in the intended
direction and parallel to one another; drilling accuracy becomes particularly important as the
bench height (and blasthole length) increases. Even where good supervision and drilling care
are applied, deviations usually limit the inclined length of 50 - 90 mm (2 - 3 1/2") diameter
smoothwall blastholes to about 12 m (40 ft.) or at most 15 m (50 ft.). Larger diameter holes
can be drilled deeper with less hole deviation. This does not overcome drill errors due to poor
setup.
Figure 2: Variability of spacing in pre-split row ... resulting from poor drilling
control.

Presplitting (Preshearing)

The following video clip shows a pre-split blast.

Video 1: A pre-split blast.


Presplitting involves drilling a row of closely-spaced blastholes along the design excavation
limit. These blastholes are very lightly charged and then detonated simultaneously or in groups
separated by short (i.e. 17ms) surface delays. The presplit blastholes can be fired either as a
separate shot or with the production blast. In the latter case, the presplit should detonate
about 50 ms before the earliest-firing production blasthole(s). Firing of the presplit charges
splits the rock just along the design excavation limit, creating an internal surface to which the
production blast can then break. Firing of the presplit charges will itself create overbreak, of
course, if the presplit blastholes are too close together or charged too heavily.

Figure 3: Angled drilling for pre-split holes.


Figure 4: Initiation of a pre-split blast.
Unlike trim blasting (see Trim Blasting), presplitting never makes it necessary to move digging
equipment into the blast site twice. Presplitting can give more spectacular results (especially in
strong massive rocks), but is generally more costly than trim blasting. Presplitting rarely gives
impressive results in highly-fissured ground where, in cases of overcharging, it can be quite
detrimental. Trim blasting is very often preferred to presplitting in such poor ground.

The spacing between presplit blastholes normally increases with the blasthole diameter as
shown in Table 1. But, because rock properties have an over-riding effect on blasthole spacing
and charge load, the distances shown in Table 1 should be considered only as recommended
starting values. The best spacing and charge load for a particular rock should be determined by
field trials. In easier-breaking formations, for example, trials may show that the spacings given
in Table 1 can be increased by 25% or more.

Explosive charges used in pre-splitting usually consist of one of the following:

• a continuous column of emulsion or watergel explosive with a length of detonating


cord through the centre
• cartridges of explosive suspended at specific intervals down the blasthole initiated by a
length of detonating cord
• a small base charge of bulk explosive
Combinations of these charges are also used, depending on blasthole diameter and rock
properties. A specifically designed high strength detonating cord is sometimes used as the sole
charge in smaller diameter blastholes.

The most efficient pre-split charge consists of a continuous de-coupled length of explosive in
the blasthole. The utilization of a small base charge of bulk explosive as the only charge in the
pre-split blasthole is usually only effective in large diameter blastholes in a rock that is
relatively easy to split.
Figure 5: Typical pre-split blast design.
Presplit holes are generally charged to within about 8 blasthole diameters of the collar when
de-coupled charges are used. If noise and airblast can be tolerated, presplit blastholes should
not be stemmed. In residential areas, presplit blastholes should be stemmed by pushing down
a wad of damp paper or hessian to the top of the charge and then backfilling above with drill
cuttings.

If the best possible presplitting action is to be obtained, presplit charges must be fired very
quickly or simultaneously. This is achieved most successfully by joining up all detonating cord
downlines with a trunkline of 3.6 - 5.0 g/m (18 - 25 gr/ft) cord. Where ground vibrations are
likely to disturb residents, however, 17ms or 25 ms surface delays should be inserted into the
trunkline at intervals, so as to obtain the consecutive firing of groups of holes. The number of
holes in each group is made sufficient to obtain a satisfactory splitting action while not
exceeding the maximum charge weight that can be fired per delay.
Figure 6. A line of presplit blastholes fired with a double-row production
shot.
Where noise is to be minimized, the detonating cord trunkline should be covered with not less
than 200 mm (8 in) of sand or fine screenings, or nonel type detonators should be employed.
If, for some reason, trunklines cannot be used, each downline can be initiated at the collar of
the blasthole by a millisecond-delay detonator (preferably the zero delay). If ground vibration
limitations prevent the use of all zero delay detonators, blastholes should be initiated in
groups using zeros and delay numbers 1, 2, 3, etc..

To enable the presplit fracture to develop to its fullest extent, the presplit blastholes should be
initiated at least 50 ms before the first front-row holes in the production blast, or as a separate
blast. Some examples of advance pr-split blasting are shown below.
Figure 7: Example 1 of pre-split holes being initiated in advance of production
holes.

Figure 8: Example 2 of pre-split holes being initiated in advance of


production holes.
Figure 9: Example 3 of pre-split holes being initiated in advance of production
holes.
The presplit face will be damaged or even destroyed if production blastholes are drilled too
close to it. If the distance between the presplit and the back row of production blastholes is
too great, on the other hand, rock is left in front of the presplit face; this wedge or fillet of rock
could create a hazardous mucking condition and may make it necessary to reblast this area.
The best distance between the presplit and back-row production blastholes should be
determined by trial; it is usually 1/3 - 1/2 times the normal burden distance for production
blastholes.

Disruption of the presplit face can also be minimized by careful design of the initiation
sequence and delay timing of the production blast. Production holes in the back row must be
able to push their burden forward easily. This is achieved by using the initiation sequence
shown in Figure 6.

Trim Blasting

Sometimes referred to as post-splitting, slashing or slabbing, trim blasting consists of drilling a


row of closely-spaced blastholes with a suitable burden:spacing ratio along the final excavation
limit. All blastholes are charged with light, well-distributed charges and fired simultaneously
(or nearly so), preferably with a detonating cord trunkline, to remove the narrow berm left in
place after the final production blast in that area has been dug out. To avoid loss of blastholes
through caving, trim blastholes should be drilled after firing the adjacent production shot.

Trim blastholes are charged and fired in the same way as those for presplitting, so that the
detonation tends to split the rock web between holes giving a smooth wall with minimum
overbreak. For 51 - 89 mm (2 - 3.5 in) diameter blastholes, the charge usually consists of a
string of small-diameter cartridges suspended at 0.5 - 1.0 m (1.5 - 3 ft) intervals on a
detonating cord downline. To achieve the best charge distribution, however, continuous
columns of 22 mm (7/8 in) (for 76 mm (3 in) blastholes) or 25 mm (1 in) diameter explosive (for
89 mm (3 1/2 in) blastholes) should be used.

The burden and spacing increase with the blasthole diameter (see Table 1 above). The burden
must always be greater than the spacing. Stemming lengths are typically 0.6 - 1.0 m (2 - 3 ft)
for 51 - 89 mm (2 - 3 1/2 in) blastholes.

Trim blasting does not usually give the spectacular type of result so often produced in strong
massive rocks by presplitting, but it does give a considerable reduction in overbreak. Trim
blasting tends to give better results than presplitting in unconsolidated ground.

Despite the need to bring back the digging equipment to clear up the muckpile produced by
the trim blast, this method is less costly and, for many operations, more cost-efficient than
presplitting. The lower cost of trim blasting results from the increased blasthole spacings (see
Table 1 above).

Secondary Blasting

Introduction

It is often necessary to blast boulders to facilitate digging operations and, in some cases, to
break the oversize material down sufficiently to feed into a crusher. This operation, which is
known as secondary blasting, can be done by either popping or plaster shooting.

Popping

With this method, a 32 - 38 mm (1.25 - 1.5 in) blasthole is drilled into a boulder, and this is
charged with just enough explosive to give the desired breakage. The charge performs most
effectively and produces least flyrock where the blasthole is drilled to just past the centre of
the boulder. Popping is the more economical method for explosives consumption, but it
involves considerable drilling expense. An emulsion or dynamite explosive is usually preferred
because of ease of handling and lack of spillage when part-cartridges are charged. Where
ANFO is used for this purpose, the tendency is to overcharge the blasthole. This not only
wastes explosives but, more importantly, increases noise and airblast and the possibility of
hazards from flyrock. Where part cartridges of 25 mm (1 in) diameter explosives are used, the
blaster has a very high degree of control over the weight of charge in each boulder. The
powder factor should usually lie in the 0.075 - 0.110 kg/m3 (0.125 - 0.185 lb/yd3) range. To
increase breakage and reduce noise and airblast, all pop holes should be carefully stemmed to
the collar with coarse sand or fine screenings.

Each charge can be initiated:

• by a detonating cord line


• by an electric detonator (instantaneous or delay)
• by a shock-tube detonator (instantaneous or delay)

Initiation of pop charges with detonating cord is considered to be the safest firing method.
Although more costly in materials, it is a very simple and rapid procedure and, for this reason,
is low on labour costs. Unfortunately, the firing of a detonating cord trunkline and the tails of
the downlines in the atmosphere creates a sharp crack which could possibly cause complaints
from any residents nearby. The benefits of detonating cord initiation in secondary blasting,
therefore, can usually be gained only at those sites which are remote from residential areas.

Where pop charges are initiated directly by electric detonators, the wires are usually
connected up in a single series circuit and fired by a blasting machine. Any grounding of bare
connections must be prevented, and the recommended maximum number of detonators
which the blasting machine can initiate should be strictly observed. It is also most important to
recognize and remain alert to the hazards of lightning and stray currents. This is particularly
the case when the site is located in an area of high electrical storm activity. The operation of
radio transmitters in the vicinity should be strictly controlled.

Where pop charges are initiated by shock-tube detonators rather than electric detonators,
hazards are much reduced and there is less need to calculate and measure electrical
resistances. It is also possible to fire a greater number of pops at one time. But on sites which
are close to residential areas, the firing of detonating cord trunklines creates a sharp crack
which could possibly cause complaints from neighbours.

Plaster Shooting

If relatively high noise and airblast levels can be tolerated, it may be more practical to break
large boulders by plaster shooting. The required number of cartridges, in close contact with
each other, should be laid on top of, and in intimate contact with, the boulder and covered
with 100 - 150 mm (4 - 6 in) of stiff mud or clay (free from stones which might constitute a
missile hazard) to provide some measure of confinement. This clay cover increases the amount
of effective work done by the charge and reduces noise and airblast slightly. Although
dynamites are frequently used for this work, emulsions, because of their much higher
detonation pressure, have proved to be the most effective explosives.

Because rock varies greatly in hardness, powder factors for plaster charges range from about
0.6 to 1.8 kg/m3 (1.0 - 3.0 lb/yd3). Powder factors for strong massive rocks like granite and
basalt usually range from 0.75 to 1.0 kg/m3 (1.25 - 1.70 lb/yd3). If the rock is either very large
or its width greatly exceeds its height, the charge should be divided into two or more parts, so
placed that they have maximum effect.

Because 10 - 15 times as much explosive is required to do the same amount of work, plaster
shooting is generally considered to be more expensive than popping. But if drilling equipment
is not available, or if the rock is very hard and difficult to drill, there may be no cost-efficient
alternative. Plaster shooting may actually be cheaper in the long run, especially if digging and
hauling operations are being delayed by the presence of large boulders in the muckpile.

Obviously, several charges in a group must be fired instantaneously either by detonators or


detonating cord.
Review #6

The randomly selected multiple-choice questions below are designed to


review your understanding of the material covered in the preceding sessions.
Your selections are lost when you leave the review page. On return the
review will start afresh with a new selection of questions.

This review is currently set to practise mode. To optimize your learning


experience you need to register for certification before entering the course.
Certification tests more rigorously, keeps track of your answers to the
multiple choice review questions, and enables you to report and submit your
review scores to complete the certification process. If you have already
registered and been approved for certification then you should Exit and re-
enter before proceeding.

Each question below has one or more correct responses. Your selection of a
response is immediately marked correct or not.

Q1. Overbreak from production blasts can be reduced by ...

using decked or decoupled charges in the back row blastholes?

using vertical blastholes in the back row?

minimizing the number of charges firing on the same delay?

increasing the burden on the backrow blastholes to about 30


times the blasthole diameter?

increasing the stemming in the back row blastholes?

Q2. For minimizing overbreak in unconsolidated material, the best results


are achieved by ...

presplitting?

trim blasting?
Q3. Minimizing overbreak should ...

increase the recoverable rock reserves?

decrease the overall size of the excavation?

allowing steepening of benches?

reduce the occurrence of rockfalls?

minimize the need for rock reinforcement?

Q4. The most effective explosive for plaster shooting is ...

dynamite?

emulsion?

watergel?

Q5. Opening and extension of fissures by invading explosion gases


propagates faster in ...

shorter fissures?

longer fissures?
Blast Design and Assessment for Surface Mines and Quarries (Text Level)
Part 7: Damage Control, Safety and Accident Prevention

Introduction

Blast Damage Prevention and Noise Control

Blasting in urban areas must be controlled to ensure that damage to structures and
disturbance to people living and working in the vicinity is prevented. Three types of damage
are caused by blasting (Figure 1):

Figure 1. Causes of blast damage.


• structural damage due to ground vibration induced in the rock by the detonating
explosive
• damage due to flyrock or boulders ejected from the blast area
• damage due to air blast and noise

When designing a blast in an urban area, one must take into consideration not only the
potential for damage in the vicinity of the blast, but also possible disturbance to people a
considerable distance from the blast site. This disturbance to people living outside the
potential damage zone can give rise to complaints and possibly spurious damage claims. This
and the following two sessions provide information on allowable vibration and air blast levels
and discusses methods of preventing structural damage, and disturbance to people.

On most urban quarry projects, blasting is only one of the many possible sources of vibration.
Figure 2 shows how the peak particle velocity produced by various types of construction
equipment compares with the vibrations produced by detonation of 0.5 kg (1.1 lb) of
dynamite. This chart represents approximate values, and actual vibration levels will vary from
site to site. However, the chart does show the relative magnitude of vibrations produced by
different sources.

Figure 2. Peak particle velocities produced by construction machinery


compared to explosive charge.
Detonation of an explosive charge produces a short duration, transient vibration compared to
other sources, such as heavy machinery, which produce steady-state (or pseudo-steady)
vibrations. Generally, the steady-state vibrations are more disturbing to people than transient
vibrations, even if the latter have a higher peak particle velocity. Vibrations produced by the
steady-state sources can damage structures very close to the construction site. Therefore, in
some circumstances, the effects of non-blasting vibrations should be considered in quarry
projects.

Structural Damage

The fragmentation of rock by detonation of an explosive charge depends on the effects of both
strain induced in the rock and upon the gas pressure generated by the detonation of the
explosive. Structural damage resulting from vibration is dependent upon the strains induced in
the rock.
When an explosive charge is detonated near a free surface, two body waves and one surface
wave are generated as a result of the elastic response of the rock. The faster of the two waves
propagated within the rock is called the primary or P wave, while the slower type is known as
the secondary or S wave. The surface wave, which is slower than either the P or S wave, is
named after Rayleigh who proved its existence, and is known as the R wave. It is suggested
that, in terms of vibration damage, the R wave is the most important since it propagates along
the surface of the earth, and because its amplitude decays more slowly with distance traveled
than the P or S waves. While this may be true for damage to surface structures, rock in the
final slopes is also of concern and, hence, the effects of all three waves should be considered.

The wide variations in geometrical and geological conditions on typical blasting sites preclude
the solution of ground vibration problems by means of elastodynamic equations. Therefore,
the most reliable predictions are given by empirical relationships developed as a result of
observations of actual blasts.

Of the three most easily measured properties of the stress waves, that is, acceleration, velocity
and displacement, it is generally agreed that velocity can be most readily correlated with
damage to structures. The stress wave has three components - vertical, radial and transverse
and it is necessary to measure all three components and use the greatest, termed the Peak
Particle Velocity (PPV), to assess damage potential. The particle velocity is a measure of the
velocity of particles of ground during passage of the vibration wave, and not the velocity of the
wave itself.

Of the many empirical relationships between blast geometries and vibration velocities that
have been postulated, the most reliable appears to be that relating particle velocity to scaled
distance. The scaled distances is defined by the function R/W , (where R is the radial distance
from the point of detonation and W is the weight of explosive detonated per delay. The U.S.
Bureau of Mines has established that the maximum particle velocity, PPV, is related to the
scaled distance by the following relationship:

where:

PPV = Peak Particle Velocity

R = Distance from blast to point of concern

W = Charge weight per delay

k = site factor

b = site factor

k and b are constants which have to be determined by measurements on each particular


blasting site.
The above equation plots as a straight line on log-log paper, and the value of k is given by the
PPV intercept at a scaled distance of unity, while the constant b is given by the slope of the
line. A hypothetical example of such a plot is given in Figure 3.

Figure 3. Hypothetical plot of measured particle velocity versus scaled


distance from blast.

See the interactive tool Peak Particle Velocity (PPV) for Assessment of Structural Damage
Potential for a demonstration of blast vibration principles.

In order to obtain data for the construction of a plot, such as that given in Figure 3, some form
of vibration measuring instrument must be available. Table 1 shows typical specifications for a
seismograph suitable for measuring blast vibrations, as well as vibrations produced by a wide
range of construction equipment such as pile drivers. Of importance for monitoring non-
blasting applications is the useful frequency range of the equipment (i.e. 2 to 250 Hz). Some of
the features of instruments currently available are measurement of acceleration, displacement
and frequency in addition to velocity, a trigger that starts recording when a pre-set vibration
level is detected, and digital recording of the results which can be transferred to a computer
for further analysis.

Table 1 Technical Specifications for Blast Vibration Monitor


Vibration Damage and Control

Vibration Damage Thresholds

Much work has been carried out to determine threshold vibration levels for damage to many
different types of structures. These peak particle velocity levels, which are listed in Table 1,
have been established from many observations in the field and are now used with confidence
in the design of blasts. However, variability in site conditions mean that caution should always
be exercised if vibration levels are close to the threshold values quoted in Table 1 right.
The damage criteria most frequently used in urban areas is a peak particle velocity of 50 mm/s
(2 in/s), which is the level below which damage to most residential structures is very slight.
However, for very poorly constructed buildings, and old buildings of historic interest, allowable
vibration levels may be as low as 10 mm/s (0.4 in/s). The required limit should be determined
from the structural condition of the building.

Figure 1 shows the relationship between the distance of a structure from the blast, the
explosive weight detonated per delay, and the expected peak particle velocity at the structure.
These graphs have been drawn up using the relationship

for values of k = 200 and b = 1.5. It is intended for use as a guideline in the assessment of blast
damage. If this graph indicates that the vibration level is approaching the damage threshold,
then it is usually wise to carry out trial blasts and measure the vibration levels produced. These
data can be used to produce a graph similar to that shown in Figure 2, from which the values
of the constants k and b for the site, and thus the allowable charge weight per delay, can be
determined. It has been found that the values of k and b will vary considerably from site to
site, so vibration measurements are useful in critical situations unless very conservative
restrictions are placed on allowable explosive weights per delay.
Figure 1. Typical blast vibration control diagram.
Figure 2. Hypothetical plot of measured particle velocity versus scaled
distance from blast.
See the interactive tool Blast Parameters for a Specified Vibration Threshold for a
demonstration of the blast vibration principles involved.

Effect of Vibration Frequency

The frequency of the vibration is also of importance in assessing damage potential. If the
principal frequency, that is the frequency of greatest amplitude pulse, is approximately equal
to the natural frequency of the structure, then there is a greater risk of damage than if the
principal and natural frequencies are significantly different. The natural frequency of two-
storey residential buildings is in the range of 5 to 20 Hz, and the natural frequency decreases
with increasing height of the structure. The principal frequency of a blast will vary with such
factors as the type of blast, the distance between the blast and the structure, and the material
through which the ground vibrations travel. Typical construction blasts produce vibrations with
principal frequencies in the range of about 50 to 200 Hz. Figure 3 It is found that large quarry
and surface mine blasts produce vibrations with lower principal frequencies than do
construction blasts, that principal frequencies decrease with increasing distance due to
frequency attenuation, and that vibrations measured in rock have higher frequencies than
those measured in soil.

Figure 3. Allowable vibration levels for residential structures due to blasting.


Figure 3 shows relationship between peak particle velocity, vibration frequency and
approximate damage thresholds for residential structures. This diagram demonstrates that
with increasing frequency the allowable PPV also increases.

Effect of Geology

It has been found that blast vibrations are modified by the presence of overburden at the
measurement location. In general, vibrations measured on overburden have a lower frequency
and higher amplitude than those measured on rock at the same distance from the blast. A
consequence of this is that if the particle velocity is approximately the same at two locations,
the lower frequency of the vibrations in overburden will make the blast vibrations more readily
felt by humans.

Electronic Equipment and Machinery

Some types of electronic/electrical equipment are sensitive to vibrations. Studies have been
done on computer disk drives, telecommunication equipment such as relay stations and fiber
optic cables, as well as electrical equipment such as older model power transformers which
have mercury cut-out switches. Vibrations from blasting, pile driving, or other construction
activities can interfere with the operation of this equipment. Table 2 shows some typical
vibration threshold values for these types of equipment. The manufacturer should always be
consulted to determine the specific vibration limit for each type of equipment subjected to
vibrations.

Human Response to Blast Vibrations


Humans are very sensitive to vibrations and can feel the effects of a blast well outside the
potential damage zone. Figure 4 shows the relationship between peak particle velocity,
frequency and the possible human response to the vibrations. This indicates that low
frequency vibrations are more readily felt than high frequency vibrations. The frequency of
blast vibrations is usually in the range of 10 to 50 Hz.

To illustrate the distances over which vibrations are perceptible, consider the effect of a blast
in which 25 kg (55 lb) of explosive is detonated per delay. Using either the equation for PPV
(above) with k=1600 and b= -1.5 , or Figure 1, it may be shown that there is a potential for
damage to houses at distances up to 50 m (160 ft), but that the vibrations can be perceptible
for distances up to 400 m (1300 ft). This can give rise to complaints and possible damage
claims at considerable distances from the blast.

Control of Vibrations
The magnitude of blast vibrations at a particular location is dependent upon the distance from
the blast and the charge weight per delay. There are also the blast design factors discussed
earlier in the course, the most important of which is the use of delays and the correct
detonation sequence to ensure that each hole, or row of holes, breaks toward at least one free
face.

To control vibration levels at a particular distance from the blast, it is necessary to limit the
explosive charge detonated per delay according to the PPV relationship given in the previous
session. Calculation of the allowable charge weight per delay, from either trial blasts or design
charts, will determine how many holes can be detonated on a single delay. If the charge
weight in a single hole is more than that allowable, then either shorter holes must be drilled,
or a decked charge could be used. In a decked charge, the explosive load is separated into two
halves with drill cuttings, and each charge is detonated on a separate delay. The minimum
delay interval between charges or holes in order to ensure no constructive interference of
vibrations produced by each charge is about 15 to 20 ms.

Pre-Blast Surveys

At any location where there is a potential for damage to structures due to blast vibrations, it is
usually wise to carry out pre-blast surveys on all structures within the potential damage zone.
These surveys should record, with photographs and/or video tape where appropriate, all pre-
existing cracks, other structural damage, and drainage problems. The U.S. Office of Surface
Mining has drawn up a standardized method of recording structural damage which ensures
that the survey is systematic and thorough. In addition to the building survey, a public
relations program informing people of the blasting program and actual vibration
measurements will usually minimize complaints and eliminate spurious damage claims.

Flyrock, Air Blast and Noise Control

Control of Flyrock

When the front row burden is inadequate or when the stemming column is too short, a crater
is formed and rock is ejected from the crater and it may be thrown a considerable distance
(Figure 1a,b,c). This figure also shows that flyrock can be caused by poor drill alignment and
geologic conditions which allow venting of the explosive gases along discontinuities in the rock
mass.
Figure 1a. Common causes of flyrock.

Figure 1b. Common causes of flyrock.


Figure 1c. Common causes of flyrock.

In a study by the Swedish Detonic Research Foundation, the maximum distance which
boulders were thrown was studied for a range of powder factors, and the results are plotted in
Figure 2. This plot shows that, for the particular rock mass and blasting geometry tested, the
flyrock problem could be eliminated by reducing the powder factor to 0.2 kg/m3 (0.012 lb/ft3).
A low powder factor, such as that required to eliminate flyrock, may not give adequate
fragmentation and, hence, other forms of flyrock control would have to be used.

Figure 2. Maximum throw of flyrock as a function of powder factor in tests by


the Swedish Detonic Research Foundation.
An alternative to changing the powder factor would be to increase the front row burden
and/or the length of the stemming column but, as pointed out in Burden Distance and
Blasthole Spacing, this could give rise to choking the blast and to poor fragmentation of the
rock above the top load. A stemming column length of 40 times the blasthole diameter is
recommended by the Swedish Detonic Research Foundation for the control of flyrock, and this
is in line with the generally recommended optimum stemming column length of 0.7 to 2 times
the burden.

In practice, the complete control of flyrock is difficult, even if the blast is designed with the
recommended stemming and burden dimensions. Therefore, in areas where there is a
possibility of damage to structures, blasting mats should be used to control flyrock. Blasting
mats consist of rubber tires or strips of conveyor belting chained together, which, in extreme
circumstances, should be weighted with soil, or anchored to the bedrock.
Control of Air Blast and Noise

These two problems are taken together because they both arise from the same cause. Air
blast, which occurs close to the blast itself, can cause structural damage such as the breaking
of windows. Noise, into which the air blast degenerates with distance from the blast, can cause
discomfort and give rise to complaints from those living close to the mining/quarry site.

Factors contributing to the development of an air blast and noise include overcharged
blastholes, poor stemming, uncovered detonating cord, venting of developing cracks in the
rock, and the use of inadequate burdens giving rise to cratering. Figure 3

The propagation of the pressure wave depends upon atmospheric conditions including
temperature, wind, and the barometric pressure-altitude relationship. Cloud cover can also
cause reflection of the pressure wave back to ground level at some distance from the blast.
Figure 3 gives an indication of how the propagation of the shockwave is effected by the
variation of temperature with altitude. This shows that air blast problems can be most severe
during temperature inversions.

Figure 3. Effect of atmospheric conditions on airblast.


Figure 4 gives a useful guide to the response of structures and humans to sound pressure level.
The maximum safe airblast levels recommended by the U.S. Bureau of Mines are:
Figure 5 gives the results of pressure measurements carried out by the U.S. Bureau of Mines in
a number of quarries. The burden, B, was varied and the length of stemming was 0.3 m/mm
diameter of borehole. For example, if a 1000 lb (454 kg) charge is detonated with a burden of
10 ft (3 m) then the over-pressure at a distance of 500 ft (152 m) is found as follows:

Figure 4. Human and structural response to sound pressure level.

From Figure 5, over-pressure equals about 0.022 psi (0.152 kPa) or 137 db.
Figure 5. Over-pressure as a function of scaled distance for bench blasting.
These calculations are related to the air blast produced by the explosive charge itself.
However, a significant component of the air blast is produced by the detonating cord and
considerable reduction in noise can be achieved by either covering the detonating cord with
sand, or using a detonating system such as the Nonel.

Safety and Accident Prevention 1

General

Blasting operations need to be both safe and cost-efficient. Whilst there are rapidly mounting
pressures to minimize the cost and period of excavation and to maximize overall cost
efficiency, all relevant personnel must recognize and remember that total safety is the first
objective. Only after a safe blasting operation has been established can one start to consider
cost efficiency.

Everyone who handles and uses explosives should remember, at all times, that he is dealing
with materials that can be hazardous. When properly treated, explosives are safe. But in
careless or inexperienced hands they can cause death and destruction.
These notes are intended as a supplement to local, state and national regulations applicable to
the storage, handling and use of explosives as well as drilling and blasting for the construction
and mining industries; and they do not supersede these regulations. The applicable regulations
and procedures must be followed.

Personnel

All blasting operations must be conducted under the direct supervision and responsibility of an
experienced shotfirer or blaster with a valid license for the type of blasting being conducted.
As the number of people handling explosives decreases, so does the probability that an
accident will take place. Where a crew of people works together, the several tasks involved in
charging and firing the blast should be divided up so that each man has a clearly defined job to
do. The firing of the blast should be the responsibility of the licensed blaster only, and it is their
responsibility to also ensure that the shot has been loaded and tied in correctly. With such a
system, everything can be done in a precise and orderly manner. People, other than those
directly concerned with the blast, should be kept away from the blast area.

The people who handle explosives should understand explosives and know which practice is
safe and which is dangerous. A totally inexperienced person should always be given definite
instructions before he is allowed to handle explosives at all, and then should work under the
direct supervision of a careful experienced man until he demonstrates that he can be relied
upon not to endanger himself or his fellow workers.

People who are in the habit of using explosives but who, through ignorance, carelessness or
bravado, follow unsafe practices, constitute the greatest problem. Some virile young people
show a tendency to live dangerously and take risks; others also tend to ridicule laws and
regulations which appear (to them) to have an unmanly and/or pointless nature. If a person is
careless to the point of recklessness and is unwilling to change their ways, the sooner they are
removed from all contact with explosives the better for all concerned. People must not only be
taught safe practices; they must also be made to apply them. Therefore, close supervision and
strict discipline must be maintained.

Very few blasting accidents are caused by lack of knowledge or experience. They are usually
the result of a mental attitude that does not put safety before all other considerations. Rules,
regulations and even the most safety-conscious supervisor cannot prevent accidents. Only the
proper attitude towards safety by every member of the crew can achieve this. Every person is
responsible for the safety of both himself and his fellow workers.

Preparations in the Blast Area

The blast area should be cleared of all unnecessary people and equipment before any
explosives or detonators are delivered to the area. Explosives, detonators or primers should
not be placed:

• in locations where they may be impacted by falling rocks or other objects


• in or near the path of vehicles
It is essential that electric detonators be kept well away from electric pumps, cables, lights,
switch gear, starter motors, generators, batteries, etc. Detonator boxes should be kept locked
until charging is to commence.

In benching, activity of a continuous nature, such as the operation of front-end loaders, should
be minimized in front of a face being charged with explosives; it is particularly important to
observe this rule if electric detonators and dynamites are being charged into blastholes. Where
non-electric detonators and an emulsion explosive and/or ANFO are employed, the hazards of
operating in front of the blast are reduced.

Drilling into Explosives

In tunneling and trenching, drilling into explosives is probably the most common cause of
serious blasting accidents. It is also one of the most difficult to eliminate in tunneling, since
there is always a possibility that undetonated explosives remain in the face after the blast.
Under certain conditions, this hazard will still be present regardless of the precautions taken.
Great care is always necessary wherever blastholes are drilled into a face which has been
previously blasted.

Before drilling commences, the face should be thoroughly examined for any evidence of
undetonated explosive. Under some conditions, such explosive may be very difficult to detect,
since it may have been driven into cracks in the rock. Therefore, considerable efforts should be
made to locate and remove undetonated explosives:

• by completely flushing out all bootlegs and cut-off blastholes


• by carefully washing down the face with a strong jet of water

Even though it may appear to be free of undetonated explosive, a bootleg must never be
drilled into. If a blasthole is still intact, one should assume that it was charged and was fired
along with the other blastholes in the round. Bootlegs should be circled with florescent paint
so their location is easily identified by the drilling crew.

Each hole should be collared at a distance not less than 150 mm from any bootleg, and drilled
in a direction such that it will not converge with the bootleg. This rule should be strictly
observed; there is never a valid reason to ignore it.

To further reduce accidents caused by drilling into explosives, blastholes should not be
collared adjacent to the toes of burn cut holes of the previous round, even though there may
be an absence of bootlegs. The position of the cut should be alternated from one side to the
other side of the centre-line for successive rounds.

The probability of leaving explosives in the face can be minimized by drilling the round so that:

• all blastholes bottom at about the same distance from the face
• one charge cannot cut off an adjacent later-firing charge

Every primer should be placed at the bottom of the blasthole.

Unsafe Charging Practices


Most of the accidents in this category are the result of pounding (i.e., excessively vigorous
tamping) during charging operations. This type of accident can also be caused through the use
of metal tamping or charging devices. Only a wooden tamping rod (or one made of an
approved non-metallic material) should be used; it should contain no metal parts. Even the so-
called non-sparking metals (e.g., brass, copper, aluminium) cannot be safely used, because
some explosives can be detonated by impact and/or friction between such metals and a rock
surface.

Together with the above recommendations, the following precautions will help to eliminate
this type of accident.

• All blastholes should be cleaned before any attempt is made to charge them.
• Blastholes should be measured with a charging pole (for horizontal blastholes) or
weighted tape (for deep downholes) to check their length and condition. If any
obstructions are present, no attempt should be made to force explosives past them. A
blasthole should not be charged if its diameter is insufficient to take the cartridges
without having to force them into place.
• Paper or cardboard shells must never be removed from cartridges of dynamite, as this
greatly increases the friction hazard.
• The primer cartridge should never be slit. Only one detonator should be placed in any
one cartridge of explosive.
• The primer cartridge must never be tamped, but merely pushed or lowered into
position. If the primer should become lodged, so that it cannot be moved forwards or
back by gentle pressure, no additional pressure should be applied to dislodge it.
• Care should be taken to prevent the detonator from being pulled out of the primer
during the charging operation.

Safety and Accident Prevention 2

Premature Firing of Electrical Blasts

When firing rounds electrically, it is essential to guard against:

• accidental application of the firing current before everyone has reached a safe place
• entry of extraneous electricity into the detonator circuit (which might result in a
premature blast).

Extraneous electricity is the unwanted electrical energy that may enter detonator circuits from
sources which are generated by either nature or man. Sources generated by nature include
lightning. Man-made sources include static electricity and stray ground currents from
improperly installed or faulty electrical equipment.

The protective shunt should not be removed from the leg wires of electric blasting caps until
ready to hook up the shot. The firing circuit should be short-circuited while the leads from the
blasting caps are being connected to each other and the short-circuit should not be removed
until immediately prior to blasting and until all workers have left the blast area. The energy
source for the blasting operation should be disconnected from the firing circuit immediately
after firing.

Static Electricity

Certain conditions can cause dangerous charges of static electricity to accumulate. Earthing of
these charges through an electric detonator may result in a spark of sufficient intensity to fire
the detonator.

This hazard must be avoided where ANFO is blown over or near detonator lead wires. Because
static electricity can be generated as ANFO is being blown into a blasthole, this cannot be
allowed to accumulate; it must be continuously drained away through the use of a good semi-
conductive charging system.

The pneumatic charger should be constructed of conductive materials and should be effectively
earthed. The resistance between the charger and earth should not exceed 106 ohms. This can
usually be accomplished by direct physical contact between the charger and earth, especially if
the floor (or invert in a tunnel) is wet.

If the charger is mounted on the jumbo or some other vehicle, it should be positively grounded
by a heavy flexible earth wire (bolted or welded to the charger) which is laid in a pool of water
or attached to a rockbolt. The earth wire should NOT be attached to metal air or water lines or
other fittings through which stray electrical currents could be conducted to the face. A chain
should NOT be used in place of an earth wire.

A semi-conductive charging hose provides a path for the static electricity to drain off to earth
via the hose and charger provided:

• that the hose is electrically connected to the charger


• that the charger is effectively earthed

The semi-conductive hose should have a minimum resistivity of about 15000 ohms per metre
and a maximum total resistance of about 2 x 106 ohms.

The operator handling the charging hose should NOT wear gloves. He should be in direct physical
contact with the hose. This precaution is necessary to prevent the accumulation of static charges
on the operator. The ambient relative humidity should be not less than 50%.

Where electric delay detonators are placed within downholes, plastic blasthole liners should
NOT be used in an effort to keep ANFO charges dry. The non-conductive nature of these liners
prevents the earthing of static charges which build up on the ANFO particles.

Stray Currents

Electric current flowing from a battery, generator or transformer through a power line to motors
or other electrical equipment will always return to that source by whatever paths are available.
These paths may be through bare conductors such as rails or through the earth itself. Electric
haulage is probably the greatest source of stray currents. Faulty insulation of cables which are
feeding electrically-driven equipment is a further source. Wherever such equipment is used,
stray currents may be anticipated. To guard against this hazard, electric detonator circuits
should be kept away from the immediate vicinity of continuous conductors.

The hazards which stray currents present can be greatly reduced by:

• connecting all electrical equipment to low resistance earths (not more than 1 ohm)
• electrically bonding all rails, pipe lines, armored cables, etc. together at frequent
intervals
• keeping power line insulation and insulators in good repair
• keeping the shotfiring cable and the detonator circuit completely insulated from earth,
other conductors and possible sources of stray currents
• leaving all detonator wires shorted and sheathed until connections are to be made to
fire the blast

Power Transmission Lines

When the detonator circuit is located near a transmission line, there is a possibility that an
electrical blast may be detonated prematurely. The following hazards may be involved.

• stray current - caused by accidental earthing of the power line


• lightning discharge - because of their excellent earthing systems, power lines can carry
a lightning discharge over long distances (up to about 15 km (8 miles))

Lightning

A number of fatal accidents have been caused by lightning strikes while electrical blasting
operations were in progress. Electrical blasting well within a tunnel is safe from this hazard only
if pipes, power lines, rail track or other continuous conductors that might carry the lightning
discharge are not directly adjacent to the face. Instances are known where strokes of lightning
several kilometres away have induced sufficient current into the blasting circuit to fire the
detonators.

In cases where an electric blasting is carried out, it is recommended

• that a lookout be kept during seasons of high electrical storm activity


• that a standard warning signal system be devised to alert the blasting crew
• that blasting operations be discontinued whenever an electrical storm seems imminent
(instruments for detecting the approach of an electrical storm are now in use at many
operations overseas)

The use of a lightning break of at least 5 m (16 ft) is recommended adjacent to the blasting switch
or source of power.

Accidental Initiation by Flame, Sparks, etc.

Explosives and detonators should not be subjected to flame, sparks or excessive heat. Accidents
have occurred from this cause, and it is recommended that the following precautions be
observed.

• Open flames must not be allowed close to charging operations.


• Smoking must be forbidden near charging operations or near explosives and detonators.
• Flammable materials or any machinery that is liable to throw sparks should not be
allowed in the blast area.
• Stemming should be used in all downholes. This not only protects the charge from flame
and sparks, but also increases the confinement and blasting efficiency of the explosive.

Explosives Fires

One of the most important lessons to be learned in explosives/blasting safety is to NEVER fight
an explosives fire. Dynamites, ANFO, detonating cords and even some emulsion explosives will
burn. Firemen are trained to put out all fires. Some of them may not realize that burning
explosives can, and often do, detonate. While detonation does not always occur, it should be
assumed that any burning explosives will detonate immediately. Do NOT fight explosives fires.
Do everything possible to evacuate the area and keep firemen away. Let the fire burn, and if it
explodes, hopefully no one will be near.

Safety and Accident Prevention 3

Delaying Too Long in the Blast Area

Numerous accidents have been caused by delaying too long in or near the blast area after a fuse
assembly has been ignited. Such accidents are frequently referred to as premature blasts, but
they really occur through lack of knowledge of the burning speed of safety fuse or failure to
gauge the time required to reach a place of safety.

The following precautions should be rigidly enforced.

• Where moisture is present, care must be taken to protect the end of the fuse from
contact with wet ground.
• Blasters must know the burning speed of the fuse being used.
• Fuse shorter than that specified in the Regulations must NEVER be used.
• No work should be carried out after fuse has been ignited. All necessary jobs should be
completed beforehand.

Inadequate Guarding

Before applying the firing current, initiating the shock tube or lighting the fuse, the shotfirer
must be absolutely sure that all routes to the blast area are effectively guarded. When
tunnelling, this protection should be extended to any adjacent tunnel if there is any possibility
that the blast could break through into this other working area.

Guards should halt any approaching people or vehicles at a safe distance from the blast. The
guards and the shotfirer must have some system of communication so that all know the danger
area is definitely clear before the blast is fired. No worker or individual should remain in an area
where their only means of egress passes the blast area.

Before firing any blast, the shotfirer must ensure that a clear positive warning or signal is given.
Posting of signs warning that blasting is being carried out does not provide sufficient protection.

Taking Insufficient Cover

Serious and often fatal accidents have resulted from persons being struck by flying rock.
Personnel should:

• be out of range of flying rocks (that may rebound from an adjacent slope or from the
walls of a tunnel or ricochet into areas that were erroneously considered to be safe)
• be sufficiently far from the blast to avoid possible injury from airblast

Even though past experience with normal shots may indicate that sufficient distance has been
provided, an abnormal blast may result at any time. Undetectable zones of weakness may exist
in the rock that may result in considerable quantities of rock being thrown considerable
distances. Standing behind a vehicle should not be considered as effective cover, since it offers
poor protection from rebounding rocks. It is often better to take a position of safety behind the
blast (when surface blasting) than in front of the face.

The shot-firing location should be not less than 700 m (2300 ft) from the blast when blasting
within a tunnel, but a shorter distance may be acceptable where say a cross tunnel enables
personnel to be outside the direct line of the blast. When deciding upon a suitable safety
distance under such circumstances, however, due consideration should be given to the
possibility of rock fragments ricocheting from the walls of the tunnel.

Insufficient Ventilation After Blasting

After firing a tunnel round, sufficient artificial ventilation must be provided to ensure that
working conditions at the face become safe within a reasonably short period of time. However,
ideal conditions do not always exist, and toxic fumes from blasts have resulted in a number of
serious and, in certain cases, fatal accidents. These were caused either by an inadequate
ventilation system and/or by people returning to the face too soon.

Only oxygen-balanced explosive compositions should be used in tunnel blasts. Carbon monoxide
(CO), the most persistent toxic gas, is present to some extent after every blast. CO is insoluble
in water and cannot be dissipated by means of a water spray or by wetting down the muckpile.
The only adequate method by which CO levels can be controlled is through dispersal or dilution
by positive ventilation. Oxides of nitrogen may be present in harmful quantities immediately
after a blast. Dissipation of the oxides of nitrogen can be accelerated by absorption in a damp
atmosphere, particularly when a water spray is used.

Facts and precautions which need to be recognized by all tunnel workers include the following:

• Some toxic fumes are liberated by the detonation of any commercial explosive.
• The quantities of toxic gases produced by improperly-detonated or burning charges are
greater than those generated by properly-initiated charges.
• A clear atmosphere is not necessarily a safe one. The principal toxic gas, CO, is
colourless, tasteless, odourless and non-irritating.
• The CO concentration increases appreciably when wooden spacers are used.
• Adequate positive ventilation is the only certain method of dissipating toxic gases or
diluting them to harmless concentrations.

Returning Too Soon After Blasting

No one should return to the blast area until all smoke, dust and fumes have cleared from the
tunnel. Visibility will be reduced by the smoke, obscuring possibly dangerous roof conditions or
falling rock. In addition, the gaseous products of detonation from all types of explosives contain
toxic ingredients that may be present in harmful concentrations immediately after a blast.

Some accidents have also been caused by people returning to the face too hastily as a result of
their impatience or uncertainty. By rushing into smoke in an attempt to determine the
performance of the blast, a man may well stumble over pieces of rock, fall and sustain painful
injuries.

For surface blasting operations, no one but the licensed blaster should return to the blast area
until the they have checked to ensure all charges have fired and the area is safe.

Handling of Misfires

Unexploded blastholes which have detonator wires or lengths of damaged detonating cord
exposed must be treated as misfires. A misfired charge or any undetonated explosive, whether
remaining in cut-offs or bootlegs or thrown out into the muck, introduces a hazard into later
operations. Every misfire is a potential accident, and the safest way to handle a misfire is to
prevent it. Therefore, all efforts should be made to prevent misfires by drilling, priming,
charging, stemming, connecting up and firing properly.

Handling of a misfired charge calls for the shotfirer's very best judgment, since it is impossible
to formulate rules and regulations to cover all cases that can occur. The shotfirer must never
forget that working on or near a missed blasthole is probably the most hazardous operation
associated with blasting. There are certain precautions which should be taken, and if these are
followed, much of the hazard associated with misfired blastholes will be eliminated. Under
practically all conditions, the safest way to handle a misfired blasthole is to fire it at the earliest
opportunity, before allowing unnecessary personnel to return to the blast area or any other
operations to be carried out.

If electric detonators are used, the firing cable should be disconnected from the exploder or
mains firing box before returning to the blast area. If the lead wires are accessible, test the
detonator with an approved blasting circuit tester and, if it shows a circuit, connect it to the
cable and attempt to fire it in the usual way. This will work successfully if the trouble was caused
originally by a faulty connection or by insufficient power. If the shot fails again or if the lead
wires cannot be reached, try to remove the stemming (if used) in the manner described below
and insert a fresh primer.

No attempt must ever be made to recover any part of a misfired charge either by withdrawing
it or by using any tool (especially a metal one) to scrape or prod the explosive from the blasthole.
If, in downhole benching operations, it is necessary to remove the stemming from a misfired
blasthole, this is done with greatest safety by blowing it out with a stiff rubber or plastic hose
and a strong jet of water or water plus compressed air. Do NOT use compressed air alone. A
metal auger for digging out the stemming must NOT be used, because it is difficult to determine
the location of the top of the charge, and the shotfirer may dig into and detonate the explosive.
Sufficient stemming should be removed to allow for insertion of a new primer close enough to
the misfired charge to ensure initiation. Misfired blastholes containing ANFO must be washed
out completely before taking measures to reblast them. In all instances, a new primer must be
inserted before any attempt is made to reshoot a misfired blasthole.

The investigation and correction of a misfire should be left to an experienced blaster who should
be allowed to carry out their work in a methodical manner and without interference. The blaster
must be familiar with any relevant statutory requirements.

Some operators handle misfires by drilling another blasthole far enough away for safe drilling,
but close enough so that the explosion of the charge in the new blasthole either detonates or
exposes that in the missed blasthole. This can be a hazardous procedure and should be used
only where removal of stemming from, and repriming of, the missed blasthole is not possible.
Close attention must be paid to the location and direction of the second blasthole to eliminate
the possibility of drilling into the missed charge. Where possible, it is preferable to drill the
second blasthole alongside of, rather than behind, the missed blasthole in order to prevent
throwing undetonated explosive into the muckpile.

The operation of front-end loaders in muckpiles containing unexploded dynamite is extremely


dangerous. The frequency of this danger can be significantly reduced by priming blastholes at or
very near to the bottom with delay detonators and by allocating delay numbers so that
blastholes are initiated in the desired sequence. This hazard is greatly minimized, if not
eliminated altogether, by the use of emulsion explosives and/or ANFO.

Keen observation in the investigation of a misfire will often disclose the cause of the trouble
(such as improperly-made primers, the use of non-water-resistant explosives in wet work,
improper charging practices, damage to lead wires, failure to connect a detonator into the
circuit, or an improper electrical hookup for the power available).

Any specific regulations covering the treatment of misfires must be observed.

Review #7

The randomly selected multiple-choice questions below are designed to


review your understanding of the material covered in the preceding sessions.
Your selections are lost when you leave the review page. On return the
review will start afresh with a new selection of questions.

This review is currently set to practise mode. To optimize your learning


experience you need to register for certification before entering the course.
Certification tests more rigorously, keeps track of your answers to the
multiple choice review questions, and enables you to report and submit your
review scores to complete the certification process. If you have already
registered and been approved for certification then you should Exit and re-
enter before proceeding.

Each question below has one or more correct responses. Your selection of a
response is immediately marked correct or not.

Q1. Common causes of flyrock include ...

missing or inadequate stemming?

uncovered detonating cord?

insufficient delay sequencing?

inadequate burden?

failure to evaluate unfavourable geological conditions?

poor drilling alignment?

Q2. The vibration damage potential of a blast is best determined from an


empirical relationship involving ...

the velocity of the vibration wave?

the velocity of ground particles during passage of the wave?

the displacement of ground particles during passage of the


wave?

the charge weight per blast delay?

distance from the blast?

Q3. The peak particle velocity (PPV) threshold that is generally accepted for
urban areas is ...
50 mm/s?

10 mm/s?

80 mm/s?

Q4. Techniques for controlling the vibration level at a particular distance


from a blast include ...

limiting the amount of explosive charge detonated per delay?

using decked charges on separate delays?

providing an air gap around blasthole charges?

using shorter blastholes with less charge?

Q5. Structural damage resulting from vibration caused by blasting is


primarily dependent on ...

strains induced in the rock by detonation?

gas pressure generated by detonation?

primary (P) waves propagated within the rock?

Rayleigh (R) waves propagated along the surface?

Blast Design and Assessment for Surface Mines and


Quarries - Author References

"C.I.L. Blaster's handbook", 1968. Canadian Industries


Limited, Montreal, Quebec.

"Caterpillar Performance Handbook", 1991. Caterpillar Tractor


Co.
Atlas Powder Company, 1987. "Handbook of Explosives and
Rock Blasting", USA.

Bauer A. and Cook M.A., 1961. "Observed Detonation


Pressures of Blasting Agents." CIM Bulletin, January 196l.

Bauer A., 1978. "Trends In Drilling And Blasting", CIM


Bulletin, September 1978.

Bauer A., Glynn G., Heater R. and Katsabanis P., 1984. "A
Laboratory Comparative Study Of Slurries, Emulsions, And
Heavy AN/FO Explosives." Proceedings of the 10th Conference
on Explosives and Blasting Technique, Society of Explosives
Engineers.

Bernard T. and Laboz, J.M., 1980. "Radio controlled


detonators and sequential real time blast applications." Proc.
11th Annual Sym. on Explosives and Blasting Research, ISEE,
Nashville 1995.

Bjarnholt G., 1980. "Suggestions On Standards For


Measurement And Data Evaluation In The Underwater
Explosion Test", Propellants and Explosives, 5, 67-74.

Bollinger, G.A., 1971. "Blast vibration analysis." Feffer and


Simons, Inc. Lodong and Amsterdam.

Brinkmann J.R., 1990. "An Experimental Study Of The Effects


Of Shock And Gas Penetration In Mining", Third Int. Symp. on
Rock Fragmentation by Blasting, Brisbane, August 1990.

Brown F.W., 1956. "Determination Of Basic Performance


Properties Of Blasting Explosives." Proc. 1st. Sym. on Rock
Mechanics, Colorado School of Mines .

Bureau Of Public Roads, 1966. "Presplitting - A controlled


blasting technique for rock cuts." U.S. Dept. of Commerce,
Washington, D.C., 36 pages.

Clay R.B., Cook M.A., Cook V.O., Keyes R.T. and Udy L.L.,
1965. "Behavior Of Rock During Blasting". Proceedings of the
7th Symp. on Rock Mechanics, vol. 2, 438-461.
Cook M.A., 1974. "The Science Of Industrial Explosives."
Ireco Chemicals, Salt Lake City, Utah.

Cunningham C., 1994. "Experiences with electronic delay


detonators in major production blasts." Fifth High-Tech
Seminar, Blasting Analysis International, Louisiana.

Cunningham C.V.B. and Sarracino R.S., 1990. "The


Standardisation Of Explosives Rating By Ideal Detonation
Codes." Third Int. Symp. on Rock Fragmentation by Blasting,
Brisbane, August, 345-351.

Dent N., 1994. "Electrodet - a new precise, reliable, easy to


use and cost effective electronic delay detonator system."
Fifth High-Tech Seminar, Blasting Analysis International,
Louisiana .

Devine, J.F., Beck, R.H., Meyer, A.V.C., and Duvall, W.I.,


1966. "Effect of charge weight on the vibration levels in
quarry blasting." U.S. Bureau of Mines Report of
Investigations. 5774, 37 pages.

Devine, J.F., Beck, R.H., Meyer, A.V.C., and Duvall, W.I.,


1966. "Vibration levels transmitted across presplit failure
plane." U.S. Bureau of Mines Report of Investigations. 6695,
29 pages.

Dowding, C.H., 1985. "Blast vibration monitoring and


Control." Prentice-Hall, Englewood Cliffs, NJ.

Dowding, C.H.,1984. "Effects of Repeated Blasting on a Wood


Frame House." US Bureau of Mines, Report of Investigations
8896.

Duvall, W.I. and Fogelson, D.E. 1962. "Review of criteria for


estimating damage to residences from blasting vibrations."
U.S. Bureau of Mines Report of Investigations 5968, 19
pages.

Duvall, W.I., 1965. "Design requirements for instrumentation


to record vibrations produced by blasting." U.S. Bureau of
Mines Report of Investigations. 6487, 5 pages.
Duvall, W.I., Devine, J.F., Johnson, C.F., and Meyer, A.V.C.,
1963. "Vibration from blasting at Iowa Limestone quarries."
U.S. Bureau of Mines Report of Investigations. 6270, 28
pages.

Duvall, W.I., Johnson, C.F., Meyer, A.V.C., and Devine, J.F.,


1963. "Vibrations from instantaneous and millisecond-delayed
quarry blasts." U.S. Bureau of Mines Report of Investigations.
6151, 34 pages.

Edl J.N., 1983. "Estimating Explosive Gas Pressure


Distribution." Proceedings of the 9th Conference on
Explosives and Blasting Technique, Society of Explosives
Engineers.

Finger M. et al., 1976. "The Effect Of Elemental Composition


On The Detonation Behavior Of Explosives." Sixth Int. Symp.
on Detonation, 710-722.

Given I.A. et. al., 1973. "SME Mining Engineering Handbook",


The American Institute of Mining, Metallurgical and Petroleum
Engineers Inc., New York.

Gregg, W.B., 1994 "Accublast Detonator - A new era of


precision in all-electronic detonators." Fifth High-Tech
Seminar, Blasting Analysis International, Louisiana.

Gregory, C.E., 1973. "Explosives for North American


Engineers." Trans Tech Publications.

Gustafsson, R., 1973. "Swedish blasting technique." Published


by SPI, Gotherburg, Sweden, 378 pages.

Hagan T.N., 1967. "Performance Characteristics Of


Ammonium Nitrate/Fuel Explosives." PhD thesis, University of
Queensland.

Hagan T.N., 1973. "Rock Breakage By Explosives." National


Symp. on Rock Fragmentation, Australian Geomechanics
Society.

Hagan, T.N. "Blasting Physics - What the operator can use in


1975." Proc. Australian Inst. Min. Metall. Annual Conf.
Adelaide. Part B., pages 369-386.
Harries G., 1985. "Relative Effective Energy - A Meaningful
Way To Rate Explosives." ICI Technical Newsletter.

Harries, G. and Mercer, J.K., 1975. "The science of blasting


and its use to minimize costs." Proc. Australian Inst. Min.
Metall. Annual Conf. Adelaide, Part B, pages 387-399.

Hemphill, G.B., 1981. "Blasting Operations". McGraw Hill.

Johannson, C.H. and Persson, P.A., 1970. "Detonics of high


explosives." Academic Press, London.

Katsabanis, P.D. et. al.,1995. "Blast control using accurate


detonators." Proc. 11th Annual Sym. on Explosives and
Blasting Research, ISEE, Nashville .

Ladegaard-Pedersen, A. and Dally, J.W.,1975. "A review of


factors effecting damage in blasting." Report to the National
Science Foundation. Mechanical Engineering Department,
University of Maryland. January 1975, 170 pages.

Langefors, U. and Kihlstrom, B., 1973. "The modern


technique of rock blasting." John Wiley and Sons, New York,
Second edition, 405 pages.

Lownds C.M., 1986. "The Strength Of Explosives, The


Planning and Operation of Open Pit and Strip Mines", SAIMM,
Johannesburg, 151-159.

Lundborg, N., Persson,A., Ladegaard-Pedersen, A and


Holmberg, R., 1975. "Keeping the lid on flyrock in open pit
blasting." Engineering and Mining Journal. May , pages 95-
100.

McCauley, M.L., Hoover, T.P., and Forshyth, R.A., 1974.


"Presplitting." California Division of Highways, Research
Report DA-DOT-TL-2955-4-74-32

McIntyre, J.S. and Hagan, T.N.,1976. "The design of


overburden blasts to promote highwall stability at a large
strip mine." Proc. 11th Canadian Rock Mechanics Symposium.
Vancouver, October 1976.
Mercer J.K., 1980. "Some Aspects Of Blasting Physics",
Quarry Mine and Pit, Vol. 19, No. 2.

Meyer R., 1987. "Explosives." 3rd edition, VCH


Verlagsgesellschaft, Germany.

Oriard, L.L. and Coulson, J.H., 1980. "TVA Blast Vibration


Criteria for Mass Concrete. Minimizing Detrimental
Construction Vibrations." ASCE, Preprint 80 - 175, pp. 103 -
23.

Oriard, L.L., 1971. "Blasting effects and their control in open


pit mining." Proc. 2nd Intnl. Conf. on Stability in Open Pit
Mining. Vancouver 1971. Published by AIME, New York, 1972,
pages 197-222.

Persson, P.A., 1975. "Bench drilling - An important first step


in the rock fragmentation Process." Atlas Copco Bench Drilling
Symposium. Stockholm.

Sheahan R.M. and Minchinton A., 1988. "Non-ideal Explosive


Performance Prediction Using The CPEX Model." Explosives in
Mining Workshop, Melbourne, November.

Siskind, D.E., Stachura, V.J. and Raddiffe, K.S. ,1976. "Noise


and Vibrations in Residential Structures from Quarry
Production Blasting." US Bureau of Mines, Report of
Investigations 8168.

United States Department of the Interior, 1976. "Blasting


vibrations and their effect on structures."Bureau of Mines,
Bulletin 656.

United States Department of the Interior, Bureau of Mines,


1980. "Structure Response and Damage Produced by Ground
Vibration from Surface Mine Blasting." Report of
Investigations, 8507.

Winzer, S.R. and Ritter, A., 1979." A field test of Du Pont new
technology MS delay electric blasting caps." Martin Marietta
Laboratories MML TR 79-44C.

Worker's Compensation Board of British Columbia. " Industrial


Health and Safety Regulations." Section 46, Amended 1980.
YAMAMOTO, M. et. al.,1995. "Smooth blasting with the
electronic delay detonator." Proc. 11th Annual Sym. on
Explosives and Blasting Research, ISEE, Nashville.

You might also like