Você está na página 1de 14

Chapter 2.

Blasting Effects and Their Control

INTRODUCTION

In recent years, there has been a trend in the direction of larger drilling equipment and larger diameter
blastholes. Although this change has improved the efficiencies and reduced the costs in many operations, it has
increased the potential for damage to underground openings. In addition, in many instances one now finds more
sophisticated delicate instruments, automated control facilities, and a large variety of structures in proximity to
blasting activity. The combined effect of larger-scale
blasting activity and its proximity to various features of
interest is such that there is an increased need for a more
refined analysis of blasting effects and their control.
BLASTING EFFECTS ON ROCK SURFACES
The Breakage Mechanism
In order to develop techniques for controlled blasting, one must first understand the features of the mechanisms by which blasting causes rock breakage to occur.
These features have not been easy to demonstrate,
mostly due to the difficulty in making tests and observations at the high stress levels and short time durations
involved.
When an explosive charge is detonated, the material
surrounding the charge is subjected to a nearly instantaneous, very high pressure [on the order of 1.4 to 13.8
GPa (0.2 to 2.0 X lo6 psi), depending on the explosive].
If the charge is coupled to "average" rock, this pressure
will pulverize the surrounding rock for a distance on the
order of 1 to 3 charge radii in hard rock, and to a greater
distance in softer rock (this is also dependent on the type
of explosive). As the pressure wave passes into the rock,
high tangential stresses cause radial cracks to appear,
and the nearly discontinuous radial stress zones generated by the shock front may cause tangential cracks to
appear. he extent of these-cracks depends on the energy available in the explosive, how quickly the energy
is transmitted to the rock, and the strength properties of
the rock. The discontinuous shock front is quickly dissipated, but the expanding gases generate a longer-acting
pressure. A compressive pulse travels to the nearest face
or internal rock boundary where it is reflected in tension.
The tensile strengths of most rocks are roughly I/loto Mo
of their compressive strengths, so the rock may now fail
in tension whereas it may have been able to support the
diminished compressive phase without failure. The tensile deflection typically produces a failure described as
tensile slabbing or scabbing.
Laboratory experiments and field experience have
pretty well established that several mechanisms are involved. ~ h e s einclude ( 1 ) the classical case of tensile
parallel slabbing when the pressure pulse is reflected at
a free surface; (2) failure under quasi-static compressive
loading (the shape is normally irregular due to discontinuities in the rock); ( 3 ) radial cracking under the
action of tangential stresses at the periphery of the expanding pressure pulse; ( 4 ) peripheral cracking at the
discontinuous shock front which is quickly dissipated;
and (5) additional mass shifting due to the venting of

the explosive gases. The first three items have received


much attention in the laboratory and the literature. The
complex effects of gas venting are difficult to test in the
laboratory because of the difficulty in reproducing the
many weak planes and discontinuities typical of most
field conditions, which play such a prominent role in
determining the behavior of the rock mass subjected to
blasting. Unfortunately, gas venting effects can be projected to significant distances under certain field conditions, and are sometimes difficult to control. It is
not unusual for gas venting to be the overriding factor in
determining the final geometric shape and physical condition of the finished excavation.
Sources of Damage
For the purposes of this discussion, damage includes
not only the breaking and rupturing of rock beyond the
desired limits of excavation but also an unwanted loosening, dislocation, and disturbance of the rock mass the integrity of which one wishes to preserve (such as mine
pillars, underground openings, etc.). The sources of
damage include, of course, all those physical features of
the rock breakage mechanism. Each of these effects
must be limited to the desired zone of breakage and
excavation if the integrity of the remaining rock mass
is to remain undiminished. The primary zone of rock
breakage usually can be controlled in the normal process
of field experimentation to determine proper charge sizes
and location for primary excavation. However, it frequently happens that there is damage from sources
which are more difficult to account for in the design
process, which are often overlooked. These are ( 1) the
overbreak due to poor drilling control, (2) dislocation of
rock (mass shifting) due to venting of explosive gases,
and ( 3 ) loosening or dislocation due to the influence of
seismic waves (ground
vibrations).
CONTROL OF ROCK BREAKAGE
Importance
In studying the rock mass and blasting design considerations, it is important to keep in mind the geometric
relationships among charge size, shape, and position, and
the physical features of the rock mass to be preserved.
The features of principal interest are the external shape
and position of the rock mass relative to blasting, and
the position and attitude of any weak planes in the rock
mass.
~h~ sequenceof ~ l ~and~ Excavation
t i ~ Events
~
Unfortunately, there are too many times when the
task of preserving delicate rock is considered hopeless,
and because of this attitude, no further effort is expended towards caution or control. In such cases there
is often a failure to recognize the importance of the sequence of the procedures. Attention t~ this can greatly
reduce unwanted effects at minimum cost.
Perimeter Control
The requirements for perimeter control are highly
dependent on the special needs of each particular project. The desirable degree of control is a highly variable

1590

BLASTING
item. Nevertheless, such dramatic improvements can be
made over the simple expedient of terminating pattern
blasting at the perimeter that it is rare today to find a
major project in which some form of pattern modification is not applied at the perimeter to improve results.
Before deciding on the degree of caution to exercise, one
must evaluate the cost and time of the work compared to
the needs. Will concrete replace the overexcavated rock
beyond a prescribed perimeter? How carefully must one
preserve the integrity of the remaining rock? Can it be
allowed to ravel or fail? Is the geometric shape of the
perimeter of importance? The answers to such questions
will help to determine the approach to the work.
If a high degree of control is needed, the two most
common methods for controlling perimeter breakage are
the presplitting or preshearing method, and the smooth
blasting or smooth wall blasting method. A variation of
the latter is usually called cushion blasting, a term which
is even older in its usage.
In any method which is designed to produce precise
control of the perimeter, it is extremely important to require careful drilling. The final results cannot possibly
be better than the drilling. Poor drilling probably accounts for more overbreak in underground excavations
than does poor blasting. Extra time and effort are usually needed at the beginning of a project, since drilling
is an art, and there is usually a noticeable learning curve
as the work gets under way.
Presplitting
Presplitting or preshearing is a method of generating
a crack in the rock along the desired limit of breakage
in advance of the pattern blasting. In this method, holes
are drilled just beyond the desired perimeter [usually 76
to 152 mm (3 to 6 in.) for shallow holes, 305 mm
(12 in.) for deep holes]. Small-diameter explosive
charges are loaded into these holes and detonated simultaneously ahead of primary blasting, generating a
crack or shear along the perimeter. Size and spacing of
the holes and charges are dependent on the rock characteristics and the need for smoothness and soundness of
the final surface.
If an explosive charge is in full contact with the walls
of the drill hole, it is said to be fully coupled. When such
a charge is detonated, a very high pressure shock wave
strikes the walls, usually crushing the rock for a distance
of 1 to 3 charge radii. The expanding gases try to expand the hole and cause radial cracks to be transmitted
into the rock as the perimeter is placed in tension.
In the presplitting method, the shattering is eliminated or greatly reduced by decoupling the charge, i.e.,
an annular ring of air surrounds the cartridge. Ideally,
the charge does not touch the rock. Although the damaging effect of the shock wave is thus largely eliminated,
the expanding gases continue to work on the rock. And
if two adjacent holes are detonated simultaneously, there
is a preferential growth of the radial crack connecting
the two holes, in preference to other directions. For an
illustration, refer to Fig. 1. Assume that the decoupled
charges A and B detonate simultaneously. Stresses are
developed at particle locations x and y. At location y,
the stresses a (radiating from charge A ) oppose the
stresses from B, such that a crack does not develop. At
location x, the stresses from A enhance those from B.
The compressive stresses are not sufficient to cause the

ENHANCED

COMPRESSION
8

Fig. 1. Presplitting stress distribution. Crack propagation


enhanced in rock web between holes.

rock to fail. However, the rock is much weaker in tension, and a tensile crack between the two charges is preferred over any other direction. In addition, this crack
is given further preference after it begins to form, since
less energy is required to extend a crack than to develop
a new one.
In the design of patterns for presplitting, the explosives concentration is a function of the ratio of the hole
diameter and charge diameter and the surface area of
the presplit plane (excavation boundary). With typical
15.87- to 19.05-mm (%- to %-in.) charges in a 63.5- to
76.2-mm (2%- to 3-in.) hole, the charge concentration
is of the order of 3.4 to 5.3 Pa (0.07 to 0.11 lb per
sq ft) of perimeter surface area.
The most difficult portion of a perimeter to preserve
in the desired condition is the shoulder formed at the
perimeter in the collar zone of the presplit holes. Quite
often, extra holes are drilled to shallow depths in this
zone to assist in the formation of rectangular corners in
the rock. These holes may be loaded or left unloaded,
according to the circumstances. When not loaded, they
are usually referred to as guide holes. These work best
when drilled within about five diameters of loaded holes.
Beyond that distance, the crack propagation may not be
noticeably influenced by the existence of the guide holes.
When shoulder rock is being broken or shifted by presplit blasting, it may be necessary to increase the depth
of stemming.
Existing stress fields in the rock have an important
influence on crack propagation during presplitting. If
these in-situ stresses are oriented away from the presplit
plane, cracks may be favored in that direction, and the
presplitting results will be unsatisfactory. Under these
conditions, it may be necessary to place the holes closer
together or to use a different technique for the perimeter
blasting, such as smooth blasting or smooth wall blasting
(this will be discussed later).
It is desirable to have the presplit holes detonate simultaneously. For this reason, it is common practice to
connect the holes with detonating cord to insure simultaneous detonation. If each hole is initiated with a separate blasting cap, a certain amount of timing scatter can
be expected, depending on the cap design. However,
satisfactory results usually are obtained with this method
even though the timing scatter is a departure from
theoretical conditions. Similarly, satisfactory results are
often obtained when detonation is restricted to only 3 or
4 holes per delay due to vibration controls.

UIVDERGROUND MINING METHODS HANDBOOK


There has been much discussion over the benefits of
presplitting as a method to bring about vibration isolation, i.e., to prevent vibrations from passing effectively
beyond the presplit plane. Such assumed benefits may
not in fact exist, and the concept should be regarded
with conservatism. In most cases, a semi-infinite burden
exists, and/or horizontal in-situ stresses are present. In
such cases, the presplit crack immediately closes and
does not present a significant barrier to vibrations, being
similar to a typical joint plane in the rock. Typical
stresses found in underground workings will cause such
cracks or joints to close tightly. This may not occur
where a limited burden exists, where a different type of
problem would then exist. Sometimes a large presplit
shot will cause the permanent displacement of a large
mass of rock, leaving a large open crack and a disturbed
rock mass. Further drilling and blasting may be restricted because of this disturbance. The open crack will
serve as an effective vibration isolator only for the upper
part of the small zone adjacent to the crack on the opposite side from future blasting. Charges placed below the
level of the bottom of the crack will generate vibrations
which pass undiminished below the crack through the
rock mass in a normal fashion.
Pf it is desired that presplitting be used to develop
a vibration barrier in a rock mass, it is possible to blast
with heavier charges than normal or use more than one
row of charges to produce a fractured zone. Again, it
should be remembered that seismic waves will be diffracted around this zone so that its influence is limited.
In connection with vibration problems, it should not
be overlooked that the presplit blast itself is capable of
causing an unacceptable vibration. Because of the excessive confinement, the presplit blast will usually generate the largest amount of vibration for a given amount
of explosive.
In many instances, the balance between costs and
benefits would suggest that some form of modified presplitting be put into use. Depending on the requirements
of the particular project, the diameter and spacing of
holes could be increased until the perimeter condition
reached its limit of acceptability. Table 1 lists suggestions for designing such presplitting shots as a first-order
cut.
Smooth Blasting
Smooth blasting is similar in concept to presplitting,
except that the charges are detonated after the primary
blasting. In the ideal case, the primary blasting has been
done and the rock excavated to the last row of holes.
Then a separate blast is detonated for the last row. A
common alternative is to detonate the perimeter charges
as the last delay in a larger blast. If the latter technique
Table 1. Recommendations for
Presplit Blasting

Hole
Diameter
in.
21/2-3
4
6
8

rnm
64-76
102
152
203

Spacing

Charge
Concentration,

ft

Ib per ft

2-3'15
3-4
4-6
6-8

0.6-1.1
0.9-1.2
1.2-1.8
1.8-2.4

0.18-0.25
0.25-0.50
0.35-0.75
0.75-1.50

(kg/m)
(0.02-0.035)
(0.035-0.50)
(0.05-0.10)
(0.10-0.20)

is used, there should be some adjustment in powder factor and timing to give the greatest amount of free movement to the perimeter blasting. In smooth blasting, it is
customary to reduce the spacing between holes to a p
proximately 80% of the burden. Holes are fired simultaneously, or in groups if a vibration problem exists.
Charges can be reduced slightly below those used for
presplitting if a free face exists.
Both presplitting and smooth blasting usually produce good results in massive rock. Smooth blasting usually is capable of reducing venting damage in highly
jointed or fractured rock.
Smooth blasting has a benefit if strong in-situ stress
fields are causing presplit cracks to travel in the wrong
directions. The primary blasting removes the burden
and relieves most of the in-situ stress, so that the smooth
blasting no longer has the same unfavorable conditions
imposed.
If smooth blasting is taken to mean a completely
separate blast fired after the primary round, it is a costly
procedure for tunnel practice, and is not often used. In
hard massive rock, no special technique is usually
needed. In loose jointed rock, the stand-up time of an
exposed roof is a problem, so it is not desirable to drill
and blast a separate round for the perimeter. A compromise is to fire sections of the perimeter simultaneously
on the last delay intervals.
Modifications to Perimeter Blasting Techniques
It would be misleading to propose that presplitting or
smooth blasting methods must follow certain prescribed
patterns. There are as many variations as the rock conditions and the imagination of the explosives engineer
will allow. The plans should be tailored to the conditions
and the purpose of the blasting.
The perimeter charges may be modified by diameter,
length, position, density, strength, etc. Conditions can
vary even within a single hole. One quick approach to
modified perimeter blasting is to use a very low-density
bulk blasting agent, consisting of a mixture of ANFO
(ammonium nitrate-fuel oil) and expanded plastics. If
higher density is satisfactory or needed, requirements
may be met with low-density slurries or untamped cartridges. Going to the other extreme, one may find that
conventional presplit cartridges are too large where extreme caution is required. In such cases, holes can be
drilled closer together, leaving a narrow web to be
broken with detonating cord.
Fracture Control in Blasting
Another interesting modification to conventional
blasting is that where notches or slots are scribed in the
walls of the drill holes to enhance the growth of cracks
in the preferred plane. The scribing or slotting can be
done with mechanical tools or by means of high-pressure
water jets. These techniques offer the following advantages: (1) a perimeter fracture plane that is more
sharply defined, (2) less shattering effect on final surfaces, (3) lower vibration levels, (4) greater spacing
between holes, and ( 5 ) better extraction of a cut zone
if used in a tunnel round (Oriard, 1981) .
Research conducted at the University of Maryland
suggests that holes may be spaced up to 25 to 50 diam
and that charges may be reduced to about 1/40 or less of
the normal concentration (Barker, Fourney, and Dally,
1977). Plewman has demonstrated a fracture extension

BLASTING
of 50 diam (Plewman, 1968). The writer is familiar
with a dimension-stone auarrv in which even greater
fracture lengths have bee; achieved. As of this writing,
the technique has been tried at two rock sites and on a
project requiring partial demolition of an old concrete
structure which required extremely delicate work.
One of the rock sites is a construction site in relatively massive and competent limestone. In experimental
blasting for fracture control at this site, vee-shaped
notches were scribed mechanically in vertical holes [diam
of 63.5 mm (2% in.)]. It was found possible to generate preshear fractures in this limestone with approximately ?hto ?4 normal loading per unit of surface area
of perimeter wall when the holes were scribed. The rock
did not fracture at loadings or spacings equivalent to
those used successfully in laboratory experiments. However, the method was considered successful at this site
and proved that it could be used under typical field
conditions for large-scale blasting.
The second site was a research chamber in highly
foliated rock appended to the Peachtree Center subway
station in Atlanta, Georgia. As expected, it was far
easier to develop fractures parallel to foliation than it
was to develop them perpendicular to foliation. As before, fractures developed at reduced loadings. However,
in this highly anisotropic rock, approximately equal
results were obtained at the same spacing with the use
of heavier charges detonated conventionally without
scribing (Oriard, 1979).
A third case involved the use of explosives for partial demolition of the old concrete lock walls at Lock
and Dam No. 1, Minneapolis, MN. Field tests were
very successful in demonstrating that explosives charges
in notched holes could be reduced to about ?4 of those
in holes that were not notched. Good stemming was
required to accomplish this reduction.
Mechanical scribing tools were developed at each of
these sites. The writer has not yet had the opportunity
to test high-pressure water jet scribing, but feels that it
has the potential for producing slightly better results
than mechanical scribing.
It is probable that field results will not consistently
equal those obtained in the laboratory. There may be
need for more field data before we can scale up the
laboratory data successfully. In this writer's opinion, the
question of the fracture toughness of a given material is
complicated by another factor which we might call the
beam strength. A relatively small block of material is
more easily broken in the laboratory than the same
material in a semi-infinite mass of rock in the field.
For the enhancement of fracture growth in a presplitting operation, the ideal explosive would be a slow
explosive which produces a large volume of gas, where
the gas expansion can be maintained for a relatively long
period of time. Confining the gases is very important.
One must use good stemming so that the gases are p r e
vented from escaping before the fracture growth is w m pleted. This was dramatically demonstrated at the limestone site previously mentioned. After a given shot
failed to develop a presplit fracture, the stemming was
changed and the shot was then repeated successfully
because the stemming was not ejected the second time.
Similarly, the test blasting in concrete demonstrated the
importance of good stemming as an adjunct to good

fracture proportion (Oriard, 1980; Tart, Oriard, and


plum^. 1980).
Grade Control
In underground bench type blasting, usually there
are two aspects of grade wntrol that are of primary
interest. One of these is the depth of the rock breakage;
the other is the smoothness of the surface breakage.
In homogeneous isotropic rock, each charge in a
blasting round tends to form a crude crater, breaking
upward and outward from the bottom of the charge.
The size and shape of this crude crater are a function of
the charge concentration and the character of the rock.
Because of this tendency to form a crater, it is necessary
to drill the blastholes deeper than the desired grade level;
otherwise high spots in the rock surface will protrude
above the grade level. This additional drilling is usually
called subgrade drilling. As a first-order rule of thumb,
the depth of subgrade drilling is approximately '/3 the
dimension of the hole burden (distance to free face).
However, the range of subgrade drilling can extend from
less than zero to as much as half the burden. If a soft
layer is encountered at grade level with hard rock above,
the breakage may actually extend below the level of the
charges. If a soft layer is found a few feet above grade
level with hard rock at grade, there will be a tendency
for explosive energy to be dissipated in the soft layer,
and deeper drilling will be needed to break to grade.
The smoothness of the grade surface is also determined strongly by the character of the rock. When blasting in horizontally bedded rock with prominent horizontal parting planes, one can achieve smooth planar bottom
surfaces. However, when blasting in massive unbedded
rock, or rock with weak planes at some attitude other
than horizontal, the smoothness of the grade surface is
strongly dependent on the spacing between the holes.
Smaller charges at closer spacings will develop a
smoother surface. For planning purposes, the predicted
surface can be modeled as contiguous craters whose
bottoms are ?L3 times burden below grade.
VIBRATIONS (ELASTIC WAVES)
After the primary shock front or pressure pulse has
passed beyond the zone in which shattering or fracturing of the rock occurs, it passes through the rock in the
form of vibrations or elastic waves. As this energy
passes through the rock, it takes on different forms which
travel at different velocities and cause different types of
deformation to occur in the rock. The fastest traveling
wave originally was given the name primary or P-wave.
This is a compressional wave, sometimes called a radial
wave or longitudinal wave, because the rock is deformed
in the radial direction from the energy source. Following the P-wave is a slower traveling wave which was
originally called a secondary wave or S-wave. This is a
shear wave, sometimes called a transverse wave. Although this wave travels in the same direction as the
P-wave, the deformation of the rock is at right angles
(or transverse) to the direction of the wave travel. The
P-wave and S-wave move through the main mass of
the rock and have the general name body waves.
As the body waves arrive at the ground surface, new
forms of waves are generated. One group of these waves
is known as surface waves because they travel along the
ground surface. Their motion is quite different from that

1594

UNDERGROLIND MINING METHODS HANDBOOK

of the body waves, being characterized by larger amplitudes, lower frequencies, and a lower propagation velocity. Structures which rest on the ground surface are
usually located far enough from the blasting for the
surface waves to develop, and they receive the strongest
part of the motion from these waves. In underground
measurements, the body waves tend to be the most
significant.
If one makes the usual assumption that there is an
elastic half space that is homogeneous and isotropic,
elastic wave theory describes the wave motions that can
be anticipated. In practice, it is simpler and more reliable to determine particle motions by means of field
measurements rather than through theoretical calculations. However, it is important to remember that the
different forms of energy are propagated at different velocities. The compressional or dilational wave is propagated with the velocity

c,= [E(1 - - p ) / p ( l - - 2 p ) ( l
= [(A

+ 2G)/pI1l2

+ p)1112

where

and

E is the modulus of elasticity; p is mass density; and


p is Poisson's ratio. The constants A and G are known as
Lame's constants. G is also known as the shear modulus.
Compressional wave transmission velocities for most
rock types fall in the range of 1524 m/s (5000 fps)
to around 6096 m/s (20,000 fps), correspondingly less
for weathered or decomposed rock. Most soils fall in
the range of about 152 m / s (500 fps) to about 1220 m/s
(4000 fps).
The shear wave travels at the velocity

The ratio of compressional and shear velocities is


Poisson's ratio for most rock materials is very nearly
0.25. Thus, the velocity ratio C p / C , is often very nearly
0 = 1.73.
The Rayleigh wave is named after Lord Rayleigh
who was the first to examine the case of this seismic
wave traveling along the boundary of a free surface.
This wave is characterized by particle motion that is
polarized in a vertical plane parallel to the direction of
the wave propagation, and the particle motion is elliptical retrograde. When Poisson's ratio is equal to 0.25,
the velocity of the Rayleigh wave is 0.92 times the velocity of the shear wave.
Not only do these different wave forms travel at different velocities, but they have the additional characteristic of attenuating at different rates. In the case of
spherical symmetry in a nondispersive medium, such as
the outward-advancing body wave, elastic theory shows
that the amplitude is inversely proportional to the distance. In contrast, surface waves have an amplitude
which is inversely proportional to the square root of the
distance.

Therefore, when the point of observation is close to


the energy source, there will be a complex combination
of several different wave forms. However, as one moves
farther from the source, the wave forms become separated, arriving at different times and producing different
types of particle motion. The more distant the point of
observation is from the source, the more prominent will
the surface waves be compared to the body waves.
Both theory and observation suggest that the particle motion transmitted to a free surface is more prominent than for the same wave within the body of the solid.
For a wave arriving at normal incidence, the particle
amplitude may be doubled. This is of interest when
considering underground structures.
Kinetics of Wave Motion
The displacement or amplitude of the ground wave is
the distance from a particle at rest to its peak or trough
as the wave passes. Typical displacements for blasting
vibrations fall in the range from 0.025 to 2.5 mm (0.001
to 0.1 in.). The term amplitude is used also to refer to
the trace amplitude on the seismogram (recording of the
motion).
The frequency of a vibration is the number of cycles
that pass a given point in unit time, usually expressed as
cycles per second or hertz. Frequencies of interest for
blasting vibrations usually fall in the range from 1 to
500 Hz, and most are in the range from 5 to 100 Hz.
The period of a vibration is the length of time required for one complete cycle to pass a given point,
usually expressed in seconds. Period is the inverse of
frequency.
Particle velocity is the time rate of change of particle displacement. It is the velocity of the motion of a
particle during the passage of the seismic wave. Particle velocity is not to be confused with propagation
velocity.
Propagation velocity is the velocity with which a
wave travels through a given medium. The propagation
velocity varies widely according to the elastic properties
of the medium. Typical P-wave velocities in rock range
from about 1500 m/s (5000 fps) to about 6000 m/s
(20,000 fps). The term velocity will denote particle
velocity unless otherwise specified.
Acceleration is the time rate of change of particle
velocity. It refers to the acceleration of a particle as the
seismic wave passes this particle. For simple harmonic
motion, the following relationships apply: x is displacement at time t; A is maximum value of x which is equal
to the amplitude (zero to peak); f is frequency; v is
particle velocity; a is acceleration; and w is angular
frequency.
Some useful formulae are:
w = 2-f
x = A sin wt
Xmax. = A
v = dx/dt = w cos wt
= w sin wt 7r/2)
v,,,, = 2 ~ f A
a = d2x/dt2 = -w2A sin wt
= w Z Asin (wt m )
a,,,, = 47r2f2A.

Damage Criteria Parameters


Over the years there have been many attempts to

BLASTING
select suitable criteria for limiting vibrations or for representing the ability of some entity (structure, rock slope,
etc.) to withstand vibrations. The two parameters which
have been used most often to express the intensity of a
vibration are acceleration and particle velocity. Displacement has received somewhat less emphasis. The
purpose of these various researches generally has been
to find a single value of some vibration parameter which
can be used to express damage potential. For a specific
type of vibration and an identified structure or response
system, the problem can often be solved theoretically
through the use of response spectra.
In attempting to find a simplified approach to blasting vibrations, a number of investigators (Duvall, et al.,
1963; Devine, et al., 1966; Oriard, 1970; Nichols, et al.,
1971; Hendron and Oriard, 1972; and others) have
found it practical to use values of particle velocity as
criteria in preference to other single-valued parameters
because particle velocity appears to have the best correlation in the frequency range encompassed by most
blasting vibrations. Nevertheless, it is this writer's conclusion that for most structures a single-valued velocity
criterion is less conservative at low frequencies and more
conservative at high frequencies. One reason for this is
the larger response that occurs in most structures at low
frequencies. Even without an enhanced response, distortion (strain) of the structure plays an important role in
determining the extent of damage. Thus, large displace
ments and low frequencies tend to be more harmful than
small displacements and high frequencies, even if the
assumed criterion parameter remains constant.
Therefore one should be aware of the time history of
the motion, as well as any single value used to express
intensity. Ideally, one should look also at the response
time of the structure and compare that to the input signal. There can be an important advantage towards
greater liberalism if the response time of the structure is
large compared to the rise time or frequency of the input
signal. For example, it is not ordinarily suitable to
establish a criterion for shock waves on the basis of peak
pressure. For such short-duration transients, the damage
potential is more related to impulse, which takes the
time history into account. The time histories of both
shock wave and structure are important. The total
duration of the motion is also significant in instances
where the motion reaches or exceeds a certain threshold
level.
In the case of underground openings in rock, the
span of the opening becomes an item of considerable
importance, not only because the span has an important
bearing on the static stability of the opening but also
because it has an important bearing on the amount of
seismic energy that is reflected at the surface of the opening, hence the dynamic stability. This reflected energy is
a function of the span of the opening, the wave length
of the incoming seismic signal, and the angle of incidence of the signal.
In spite of the technical difficulties in specifying a
single value of a given parameter, it turns out in practice
that one can select a threshold value that is restricted to
certain types of vibrations and certain types of structures
(or whatever might be the entity of concern). One can
be as conservative as desired in selecting that value.
Therefore, particle velocity can serve as a useful parameter for describing the damage potential of b!asting vibra-

1595

tions, even though the degree of conservatism may


change with frequency. In other words, there may be
a different limiting value of particle velocity for 100- to
200-Hz vibration from tunnel blasting than there would
be for a 2- to 20-Hz vibration from a bench blast. Recognizing this, one can refine the analyses and still avoid
the complications of developing response spectra. However, with advances in instrumentation technology, it is
becoming ever easier to determine the spectral content
of vibrations, and it is expected that there will be an
increased use of criteria which take this into account.
Data Scaling
In order to compare blasts of different sizes at different distances, it is customary to scale the distance factor
by some function of the explosive weight. Two methods
are currently in popular use in the US. One of these
scales distance by the ?4 power of the charge weight per
delay, the other by the I/? power. Ambraseys and
Hendron (1968) have suggested the use of cube root
scaling. Research sponsored by the US Bureau of Mines
has led to the recommendation for square root scaling
(Devine, et al., 1966). This writer uses cube root scaling
for blast waves in water, sometimes for seismic body
waves in rock, but usually uses square root scaling for
surface ground motion data, and emphasizes the need
to consider modifications due to spatial and time distributions of energy (Oriard, 1972). Cube root scaling
demands certain theoretical limitations that are often
not met under conditions found for many types of
construction blasting.
Empirical prediction curves based on cube root scaling are shown in Fig. 2. In this figure, R is range (in
feet), W is charge weight per delay (in pounds), and
V is peak particle velocity (in inches per second).
Similar curves for square root scaling are shown in
Fig. 3. The trend lines on the graph can be expressed
mathematically in the form:

V = H ( D I W"2)-B k , , k p , k3. . .

where V is peak particle velocity; H is velocity at unity


scaled distance, D is distance, W is charge weight per
delay, B is slope of trend line, and k are factors defining
firing time variations, travel time variations, coupling,
confinement, geology, isolation, spatial distribution, etc.
Scaling by the ?4 power is considered to be the preferred method by some who argue that it can be shown
to be dimensionally correct by the Buckingham Pi
theorem. Yet this is only true for a spherical charge or
for a cylindrical charge whose height changes in a specified manner with a change in radius. Although these
conditions are sometimes approached in the field, it often
happens that they are not. In many blasting cases, the
hole depth is a fixed constant, and charge weight is
changed by increasing the hole diameter, thus being proportional to the change in radius squared. One needs to
recognize that theoretical conditions are seldom met perfectly, and the data is empirical. Neither of these scaling
methods accounts perfectly for all the conditions that
may be encountered. Neither accounts for the many
possible variations and complexities in blast design and
geometry. One charge may approach similarity to the
classical case of the concentrated spherical charge; another may be a very long small-diameter cylinder; and
another may be many small charges in many holes deto-

1596

UNDERGROUND MINING METHODS HANDBOOK

SCALED

DISTANCE

(5)

Fig. 3. Peak particle velocity vs. scaled distance accord-

ing to square root scaling (Oriard, 1972).


Fig. 2. Peak particle velocity vs. scaled range according
to cube root scaling (Ambraseys and Hendron, 1968).

nating simultaneously throughout a large mass of rock.


Since correction factors or judgment must be applied for
many such factors, the degree of theoretical correctness
of scaling laws does not solve the problem for the explosives engineer. Additional input is needed.
As was mentioned earlier in the section on elastic
waves, both particle motions and attenuation rates are
different for the different types of waves. When the
point of observation is close to the energy source, there
will be a complex combination of several different wave
forms. As we move farther from the source, the wave
forms become separated, arriving at different times and
producing different particle motions. The more distant
the point of observation is from the source, the more
prominent will be the surface waves compared to the
body waves. Thus, the various scaling techniques for
prediction purposes are merely an effort to express the
average attenuation rate of a given type of blasting in a
given type of geological setting.
For particle velocities below about 0.05 m / s (2.0
ips), there is relatively little significance to differences in
the prediction methods, since the predicted values themselves are not ordinarily significant. For predictions very
close to the source of energy, the methods show a more
significant departure from each other. Unfortunately,
this is the area in which blast-design parameters have an
important effect on the vibration results.
For a prediction of body waves generated for single
charges or greatly concentrated charges, cube root data
would sometimes be the best choice. For a prediction of
surface waves generated by a charge distributed in a

number of holes, square root data would be a better


choice. Both methods might give trend lines that are too
conservative for complicated geometrical arrangements,
e.g., if the charge per delay were to be distributed over
a linear distance that is appreciable compared to the distance from the point of observation, or if there is timing
scatter in the detonation of individual holes. Another
example would be large coyote blasts in which a tunnel
is filled with explosives and detonated. Under a number of such conditions, the trend line will often drop to
a flatter slope at lower values of scaled distance. Failure
to realize this might have the effect of prohibiting work
that would be quite feasible and safe.
It can be an error in judgment to become overly
engrossed with the best fit concept when there are questions about the basic data. For example, it is very common to find errors in the values assigned to charge weight
per delay, and these values have a primary input to the
data plot. In many of the complex firing patterns in use
today, there can only be an estimate of the charge weight
that detonates at any given instant of time because of the
variations in actual firing times from the theoretical or
nominal.
It is important for the person using some prediction
method to realize the differences in conservatism under
different field conditions. Hendron and Oriard (1972)
have compared the effects on quantity-distance predictions using several different methods. Consider the longrange accumulation of data by several investigators as a
basis for predicting a certain particle velocity. Fig. 4
shows a graph of charge per delay vs. range to generate
a particle velocity of 0.05 m / s (2.0 ips) or threshold of
damage predicted by Langefors (1963) for residential
structures, and a conservative prediction used by Devine
in the absence of field data. In this general case, Hen-

BLASTING
1000

-4
-

1 4 3

---

I00

0
w

a
w
w
c)

1
0
,

--

I
10

100

1000

Table 2. Typlcal Values of Allowable Charge


Welght per Delay for Vlbratlon Llmlt of
2.0 In. per Sec*
Distance to
Equipment, ftt

Allowable Charge
per Delay, Ibt

10
15
20
25
50
75
100
150
200
250

0.25
0.5
1.o
1.5
6
14
25
56
100
156

'Oriard

t Metric equivalents: A

x 0.3048 = m; Ib x 0.453 592 4 = kg.

lo,oo~ ested in

RANGE 111.)

Fig. 4. Relation between charge per delay and range for


residential vibration limits.
dron is the most conservative at small distances and the
most liberal at great distances. Langefors is the most
liberal at small distances and the most conservative at
great distances. The Oriard values fall between the other
two. Devine's line falls outside the range of the others
and represents an extreme beyond which Devine does
not expect any data. Table 2 shows the Oriard prediction line presented as a table of allowable charge weight
per delay vs. distance for 0.05 m/s (2.0 ips sec).
It is common for typical blasting to be monitored
with a single seismograph. Examining such data, one
finds a great deal of scatter in the data, and it is usually
very difficult to decide what power function would best
suit the data. An even more important fact to keep in
mind is that the data fit often changes dramatically as a
few more data points are gathered. It would not be rare
to find that a horizontal line or even a line with reverse
slope would best fit a given number of data points, and
that a completely new line would be required after a few
more points are added. As more data is added, it becomes increasingly evident that we are not usually inter-

finding the best fit, but in finding the envelope


or upper bound of the data. Predictions or limits based
on averages can be disastrous. It is not good enough to
say that the average blast won't cause damage, if we
don't want any blast to cause damage.
It is especially difficult to compare differences in the
methods for scaling distance in the middle range. For
example, take the series of points shown in Table 3.
These are points which were selected to fit exactly on a
line drawn for scaling to the Yi power, as in Fig. 5.
Using the same data, but scaling the distance by the
1/4 power, 9'3 power, and 2/3 power of the charge weight
per delay, one can plot three additional lines for the respective scaling factors. Fig. 5 reveals that scaling to a
lower fractional power has the effect of producing a line
which is concave downward, whereas scaling to a higher
fractional power has the effect of producing a line which
is concave upward. The curvature increases with departure from the best fit line. Also, within the data for
any single line, the curvature increases as the line is
extended upward or downward from middle-range
values.
In these hypothetical examples, the data were selected to fit perfectly formed lines. Of course, actual
field data would show a great deal of scatter. Typical
field data in middle-range values would show a convino
ing fit to any of these approaches. Therefore, within that

Table 3. Partlcle Veloclty Vs. Scaled Dlstance for Different Fractional Powers of Dlstance Scallng
(Hypothetical Fleld Data)
Particle
Velocity, ips'

Distance, ft'

Charge per
Delay, Ibs'

Dlw1'4

DIW1' 3

2.0
1.O
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1

500
1000
1100
1200
1300
1400
1500
1600
1700
1800
1900

625.0
1040.6
1138.5
1142.6
1140.2
1085.1
997.2
855.6
673.6
459.2
216.9

100
176
189
206
224
244
267
296
334
389
495

58.5
98.7
105.3
114.8
124.4
136.2
150.1
168.5
193.9
233.3
316.2

'Metric equivalents: ips 0.025 4

mls; A x 0.3048 = m; Ib x 0.453 592 4 = kg.

Dlw1'2

D/W2/~

1598

UNDERGROUND MINING METHODS HANDBOOK


stant of time, ( 3 ) geological environment, and (4)
confinement.
In order to evaluate these variables, we need to look
at the results of extensive field experience because an
acceptable theoretical approach has not yet been developed for calculating ground motions from typical
blasting.
In some instances it suffices to assume maximum
conservative values for all modifying factors and accept
the conservatism that this introduces into predictions or
data treatment. For example, we could assume that all
k factors equal 1.0, and assume a high value for H.
However, there are instances when such high levels of
conservatism introduce unacceptable costs into a project
or make it technically impossible to perform the work.
Clearly, such conditions call for refinement of the approach, and one needs to look again at all significant
variables.
Distance and Geometry
SCALED

DISTANCF

(Fael / Poundlal

Fig. 5. Comparison of data for different scaling powers.

At first glance, there would seem to be no ambiguity


in a term such as distance. However, it happens quite
often that there is indeed some question as to its meaning. One must often try to define it as some effective
distance to multiple charges controlled by the orientation
of the firing sequence, or some effective distance to a
charge with a high ratio of length to diameter, or some
effective distance modification because the surface area
covered by the blast is large compared to the distance to
the point of observation, or effective distance modification because of some geometric isolation. In other
words, spatial distribution can be an important consideration, especially when the distance is small. Modifications are introduced by deviation from the concept of
a point charge.

range, there is very little significance to the best fit


concept.
The important thing is to recognize that scaling is
just a means of comparing blasting data for different distances and different charge quantities. Judgment must
come in recognizing what happens when we wish to manipulate the data for various purposes. Suppose, for
example, that a small-scale test is arranged with small
charge weight and small distance in order to determine
the response characteristics and assume that this test produces the limiting scaled distance. Suppose further that
the final blast will be a larger charge at greater distance,
but will be designed for the same scaled distance so as
to generate the limiting vibration. What is the correct
charge weight? A simple arithmetic manipulation will
show that scaling to the ?hpower gives the more conservative prediction in this instance; scaling to the %
power is less conservative. Take the reverse case for
comparison. Suppose that the limiting-scaled distance
has been determined by a large blast at great distance,
and that the final blast will take place at small distance
and small charge weight. What is the limiting charge
weight? In this case, scaling to the !h power gives
a less conservative prediction and scaling to the %
power is more conservative.
This exercise should reveal to the reader how important it is to recognize where a given test value falls in the
expected range of data scatter. If the test value were
unusually low, extreme conservatism should be used for
predicting future events. If the test value were high, a
more liberal prediction could be made safely. And, of
course, it is more important to know the expected range
of data than it is to use a specific scaling law.

Similarly, there can be ambiguity regarding the term


charge weight per delay. It would be more meaningful
to define the term as charge weight detonating within a
specified interval of time and within a specified distance
interval. In the case of single charges whose dimensions
are small compared to the distance to the point of observation, the effects are controlled by sequential detonation. Variable time periods are available for selection
from the types of initiators on the market. These include electric caps in long-period series (usually % to 1
sec), short-period series (usually 20 to 250 m-sec), intermediate series (usually 1/4 to $5 sec), or combinations
of these. An additional alternative method for elective
firing is the use of a sequential timer, a detonating device which fires a series of electrical circuits with a
preselected time interval between circuits. In nonelectric caps, similar selections are available with the
added feature that the initiators can be used in a combination of surface and down-the-hole initiation for an
unlimited series. However, it is extremely important to
be aware of deviations from nominal firing times.

Factors Which Affect the Vibration


There are a great many variables which have an influence on vibrations generated by blasting. We will
offer a few comments on those variables which most
commonly have the predominant effects. These are
(1) distance and geometry, ( 2 ) charge weight per in-

Geological Environment
The elastic properties of the medium through which
the waves pass will have a strong influence on the character of the waves. The two fundamental characteristics
of the motion which are of most interest to us are the
frequency and the amplitude. As a general comment, we

Charge Weight Per Delay

BLASTING
can state that a soft medium will transmit waves with
lower frequency and larger amplitude, whereas a hard
medium will transmit waves with a higher frequency and
smaller amplitude. Thus, motions in soft saturated soils
would be quite different from those in rock. Similarly,
as the rock becomes harder and more brittle, the motions
would show an increase toward even higher frequencies
and smaller amplitudes.
Distance and charge size also have an influence on
the character of the vibrations. High frequencies are attenuated with distance; and the smaller the size of the
explosive charge, the higher will be the generated
Confinement
The concept of confinement is partly a matter of geological environment and partly a matter of blast design.
The greater the physical confinement of a charge, the
greater will be the vibration generated by the detonation
of that charge, up to the limits which the elastic properties of the confining material will allow. A typical
upper limit in construction practice comes from the
detonation of a presplit blast. Ideally, such a blast has
a semi-infinite (great) burden, the maximum to which
the material is capable of confining the charges, and as
little energy as possible is dissipated in generating more
than a single fracture plane through the rock. Such a
blast represents the normal upper limit of vibration for
a particular geological setting in typical construction or
mining practice. Under some circumstances, presplitting
may generate vibrations which are about twice the typical upper bound of down-hole bench blasting.
Minimum Delay Time for Vibration Control
Currently, there is a very widespread concept that
short time intervals between detonations are not effective in the control of ground vibrations from blasting.
It is unfortunate that this concept has become so deeply
ingrained. It has resulted in some very costly decisions
on major projects.
The following advice is presented for the consideration of the reader: If the question is an important one
for your project, cautiously test the reaction of your geological setting to the desired blasting technique. As a
general guideline, shorter time intervals can be used in
hard, brittle, heterogeneous rock, especially if the charges
are small and in boreholes at relativelv close s~acines.
One may need to use longer time intervals in soft,
elastic, homogeneous materials especially if the charges
are large and placed in widely spaced boreholes. A case
history of a large-scale blasting operation, along with
further theoretical discussion, has been presented by
Oriard and Emmert (1980).
L

Effects on Natural and Man-Made Structures


Different man-made structures, underground openings, slopes, and other entities will have different strengths
to resist vibration damage. Furthermore, the range is
quite large. Therefore, to be definitive, it is necessary to
separate these entities into appropriate groups. One
common group needing definition is that encompassing
nonengineered structures such as small mine offices, residences, and the like. Although their strengths vary
widely, a great deal of experience has been gained in the
past 40 years in the observation of residential structures,
prompted by the unfavorable reaction of homeowners to

blasting vibrations. It is now generally agreed that a


typical blasting vibration in the middle to upper frequency range with a peak particle velocity less than
0.05 m / s (2.0 ips) will not be harmful to such nonengineered structures in average condition. Most such
structures will not be damaged until the particle velocity
of the ground motion reaches a value near or above
0.1 m / s (4.0 ips). Major damage may occur in some at
about 0.2 m / s (8.0 ips), but others will still not be
harmed at this level. The writer has observed cases
where 50-year old frame residences were subjected to
ground motions of more than 0.25 m/s (10 ips) without
incurring damage.
The reader is expected to have a primary interest in
underground mining, where blasting normally generates
ground vibrations in the middle to upper frequency
range, say 20 to 200 Hz. However, if the vibration is
characterized by a very low frequency, say 1 to 5 HZ,
and the structure is thought to be unusually weak, it
would be prudent to reduce the foregoing values by
about half, depending on conditions.
If surface mining is involved, the reader should
become familiar with the latest regulations covering
that activity. At the time of this writing, it appears
possible that new regulations will be forthcoming.
Explosives users often limit vibrations to lower levels
than necessary because of the adverse response of humans. There is a trend toward giving a greater role to
this aspect of blasting effects, and certain regulations
and specifications are being revised to more conservative
levels for this reason.
Of course, engineered structures have greater
strengths than residential structures and have a similar
range of strengths. Because of their great variety, it is
difficult to define them by categories. However, many
engineered concrete structures have been subjected to
particle velocities in the range of 0.25 to 0.38 m / s (10 to
15 ips) without incurring damage, and the writer has
observed several which were subjected to particle velocities in the range of 0.5 to 0.8 m / s (20 to 30 ips) and
did not incur damage. Tied-down, heavily braced steel
structures may withstand as much as 5 m/s (200 ips),
even at low frequencies.
For rock and soil slopes, a similar scale can be formulated. At the low end of the scale, slope displacements have been observed at low levels of vibration but
were also observed in the absence of vibrations. Usually
it appears that no primary influence is found at velocities
under 0.05 to 0.1 m/s ( 2 to 4 ips), but the time history
may be changed under special circumstances. At 0.05 to
0.1 m/s ( 2 to 4 ips), we may expect the occasional falling of loose stones on slopes. At 0.13 to 0.38 m / s ( 5 to
15 ips), we may expect the falling of partly loosened sections of rock underground and on above-ground slopes,
sections of rock that would otherwise remain in place for
long periods of time. Above 0.6 m / s (25 ips), we would
expect some damage to occur in the relatively unsound
rock types, but we find also that damage may not occur
in sound rock at much higher vibration levels. Much depends on whether the seismic waves merely pass through
a section of confined rock, or whether the waves are reflected at a free boundary at the rock surface. One can
observe sections of sound rock that show no visible signs
of damage even after they have been subjected to particle velocities in excess of 2.5 m / s (100 ips). Thus, it is

1600

UNDERGROUND MINING METHODS HANDBOOK

not merely a question of variations in inherent strength


but also a question of the geometric shapes in relation to
the surfaces of reflection.
Blasting Effects on Concrete
As in the case of rock, concrete can be damaged by
either excessive vibration or by rupture. It has been this
writer's experience to see more damage from rupture
than from excessive vibration. Establishing realistic limits for either of these blasting effects is made more complex by the fact that the ability of concrete to resist
damage is not merely a function of the strength of the
concrete fabric; and the fabric strength itself varies with
time for freshly poured concrete. Most of our knowledge about the ability of concrete to resist vibration
comes from field experience, although there has been
some limited research conducted by such agencies as the
Corps of Engineers and the Portland Cement Association. Such research results were encouraging in that no
damage occurred to concrete specimens of varying ages
of curing which were subjected to moderate levels of
vibration. However, these test conditions were not severe enough to be applicable to the field conditions that
are often encountered in mining and construction. Experience demonstrates that most engineered concrete
structures are able to withstand particle velocities in the
range of 0.5 to 0.8 m/s (20 to 30 ips) at medium
frequencies, say in the range of 20 to 100 Hz, and can
withstand much higher particle velocities in the range of
several thousand hertz.
I n a recent case, this writer conducted experimental
test blasting near freshly poured concrete at a construction site. The concrete was composed of a lowstrength mix and formed in cubes 0.9 m (3 ft) on each
side. These cubes were not subject to the type of deflection one might expect for structural sections. They were
intended to represent mass or fill concrete. When they
had attained a compressive strength of 2.8 MPa (400
psi), they were subjected to particle velocities of about
1.8 m / s (70 ips). When they had attained a compressive strength of about 8.3 MPa (1200 psi), they were
subjected to more than 2.5 m/s (100 ips). This final
blast destroyed the surrounding rock and threw the concrete into the muck pile. However, it was not damaged
by the vibrations it had received (Oriard, 1980; Oriard
and Coulson, 1980).
Even without field experience, there is a justification
for recommending values in the range of 0.25 to 0.38
m/s (10 to 15 ips) for the case of tensile slabbing. This

is based on the assumption that the pressure in a medium


generated by a traveling stress wave can be represented
by the product of the acoustical impedance and the particle velocity. Assuming a tensile strength of concrete of
about 4 i o that in compression, one might expect tensile
slabbing at about 0.25 m/s (10 ips). This figure can be
increased due to the fact that the dynamic strength of
concrete is higher than its static strength. The exact
value is a function of the time history of the incoming
stress pulse. The faster the rise time of the pulse, the
greater will be the strength of the concrete in resisting it.
It is this writer's opinion that it is important to recognize
what order of magnitude of strain can be expected and
to take this into account when selecting limits. An example of the values of particle velocity that would generally be acceptable for engineered concrete structures
is illustrated in Tables 4-6. The values are those which
have been recommended by this writer for use in the
blasting of rock near existing concrete sections of various
ages. Oriard and Coulson (1980) give additional examples of criteria developed for the Tennessee Valley
Authority to provide protection to concrete.
The Distance Factor shown in the tables was a means
used to reduce the allowable particle velocity with increasing distance because of the attenuation of frequenky with distance. Experience demonstrates that
higher particle velocity can be allowed at higher frequencies. At lower frequencies, greater deflections
(hence greater strains) will be induced in those structures which are capable of responding in that fashion.
(There is a special concern for structural walls of
freshly poured concrete). This is not entirely a question of resonance. Rigid systems at resonance are
usually more durable.
The limits are varied according to the age of the
concrete. In the period of 0 to 4 hr, the concrete has not
started to set and it can still tolerate vibration. (Of
course, structural concrete forms can be damaged by
severe shaking at this time). From 4 to 24 hr, the concrete has begun to set, but has very little strength. After
7 days, the concrete has a strength that is approximately
3$ of the ultimate (28-day) strength.
Although these recommended limits are far more liberal than some which have appeared in project specifications in the past, experience has shown that they are
acceptable for many engineered concrete structures, and
will often save very large sums of money by allowing the
construction work to proceed a t a much higher rate.
The reader should not apply these criteria indiscrimi-

Table 4. An lllustratlon of Particle Veloclty and Distance Crlterla for Blasting Near Concrete
(Case History)'
Time From Batching

0- 4 hr
4-24 hr
1- 3 days
3- 7 days
7-10 days
over 1 0 days

Nonstructural Fill
and Mass Concrete

4 ips x D F t
1 ips x D F
1.5 ips x D F
3 ips x D F
8 ips x D F
15 ips x D F

Structural Concrete Walls,


Structural Slabs, etc.

2 ips x DFt
1/4 ips x DF
1 ips x DF
2 ips x DF
5 ips x DF
1 0 ips x DF

Unspecified Electrical
Equipment, ips

2
2
2
2
2
2

'Intended as an illustration of a case history, not a general recommendation for all cases. (No allowance for form bracing, which would provide
additional strength.)
tDistance factor: 0-50 ft*, multiply x 1.0; -150 ft, multiply x 0.8; 150-250 ft, multiply x 0.7; and over 250 ft, multiply x 0.6.
*Metric equivalent: ips x 0.0254 = mls.

Table 5. An lllustratlon of Allowable Charge Welght per Delay


tor Nonstructural Flll and Mass Concrete'
Age of Concrete (From Batching Time)
Distance to
Concrete, ftt

0 to 4 hr

4 to 24 hr

1 to 3 Days

10
15
20
25
50
75
100
150
200
250

0.6
1.3
2.4
3.7
11
26
47
93
165
221

0.1
0.2
0.4
0.6
2.6
4
8
16
29
39

0.17
0.4
0.7
1
3
7
13
27
48
65

3 to 7 Days

7 to 10 Days

Over 10 Days, Ibt

1.4
3
5
8
28
63
113

3
7
12
19

0.4
0.9
1.7
2.6
8
18
33
65
116
156

223
396

500

63
143
255

500
500
500

'Intended as an illustration of a case history, not a general recommendation for all cases. (No allowance for form bracing, which would provide
additional strength.)
tMetric equivalents: f l x 0.3048 = m; Ib x 0.453 592 4 = kg.

nately to all concrete structures. They are presented as


an illustration of the fact that past criteria have often
been unnecessarily conservative. This liberalization
should not be attempted without an understanding of
the time-history of the input motion and time-history of
the response. This understanding is critical to the success of establishing criteria. Its importance cannot be
overemphasized.
When assessing the effects of blasting on freshly
poured concrete, one should also consider the strength of
the forms which will hold the concrete in the early stages
of its curing. Before the concrete has attained a strength
of its own, and especially when it is still in a semiliquid
state, the forms must be capable of resisting any shaking
that might be generated.
If forms are es~eciallvstrong- and well-braced. it is
possible to allow even higher levels of vibration than
those shown in the tables. The writer is involved in
other projects where substantially higher particle velocities were allowed, but only under special circumstances.
rn addition to tensile slabbing or structural deflection, concrete may also be damaged through dislocation,
i.e.., a ~ e r m a n e n tdisplacement. This mav come about
through vibration, rupture, or venting of explosive gases.
More commonly, this type of damage comes about be-

cause the rock supporting the concrete has been shifted


or ruptured, rather than the concrete alone.
There is really no acceptable basis for making a
calculation to control this type of damage because of
the many variations that can exist in the rock and in the
blast designs. Someone in the field must examine the
rock conditions and estimate the manner in which the
rock could be broken by the blasts being considered,
then form a judgment as to whether or not it is safe to
proceed. The writer has developed criteria to prevent
rupture damage under various rock and soil conditions
(Oriard, 1973 and 1980). In many cases the field conditions justify the use of more liberal criteria than those
cited here. A judgment must be made on the basis of
the vibration characteristics, the structural response
characteristics, and the possibilities of block motion or
ground rupture.
Mechanical and Electrical
There is such a wide range in response characteristics of mechanical and electrical equipment that it is
Very difficult, and probably misleading, to give specific
recommendations for vibration limits. Even within a
narrow category, such as electrical switches or relays,
the sensitivity to vibration will often vary by at least one

Table 6. An lllustratlon of Allowable Charge Welght per Delay


tor Structural Concrete Walls, Structural Slabs, etc.*
Age of Concrete (From Batching Time)
Distance to
Concrete, ftt

0 to 4 hr

10
15
20
25
50
75
100
150
200
250

0.25
0.5
1.O
1.5
6
11
20
39
70
93

4 to 24 hr

0.1
0.4
0.8
1.5
3
5
7

1 to 3 Days

3 to 7 Days

7 to 10 Days

Over 10 Days, Ibt

0.1
0.2
0.4
0.6
2.6
4
8
16
29
39

0.25
0.5
1.O
1.5
6
11
20
39
70
93

0.7
1.7
3.1
4.8
15
35
62
123
219
293

1.9
4
7
11
37

84
150
295

500
500

'Intended as an illustration of a case history, not a general recommendation for all cases. (No allowance for form bracing, which would provide
additional strength.)
tMetric equivalents: ft x 0.3048 = m; Ib x 0.453 592 4 = kg.

1602

UNDERGROUND MINING METHODS HANDBOOK

order of magnitude. Some switches can be tripped at


particle velocities of as low as 0.025 m/s (1 ips) or less.
Others might not be affected at 0.5 m / s (20 ips).
Just as with structures, many electrical and mechanical
devices will have sensitivities that are frequency-dependent, i.e., their sensitivities are greater in certain frequency ranges.
Similar to electrical switches, mechanical devices and
controls have a wide range of sensitivities. Some are so
durable that the structure containing them may collapse
before damage is incurred. Others are highly sensitive.
Probably the first step in the judgment process is to
determine what will be the consequence if a switch is
tripped or if some other device is activated, affected, or
broken by vibration. If the results are inconsequential,
the question is academic, and there may be no need to
determine a limit. If the results induced by the vibration cannot be tolerated, two alternatives may be considered. One is, of course, to keep vibration levels below the critical level. Another approach is often more
attractive, but often overlooked, i.e., either to isolate the
device from vibration or to lock out its action, e.g., tying
down a relay or bypassing electrically an automatic
switch.
Except for such items as sensitive switches, most
electronic and electrical equipment will not be physically
damaged by particle velocities of the order of 0.05 m/s
(2.0 ips), and most mechanical equipment is far less
sensitive. The frequently quoted limit of 0.05 m/s (2.0
ips) is, therefore, appropriate to limit physical damage
to most of this equipment, but often inappropriate to
prevent activation of automatic devices.
AIR WAVES
Air Waves Above Ground

It is expected that the reader will be interested primarily in underground work. For this reason, only limited comment will be made about above-ground effects.
Occasionally it may happen that air waves from shaft or
tunnel blasting could be of concern to above-ground
structures or people, especially if the work is taking
place in an urban setting, or in proximity to mine offices
or processing facilities.
An overpressure of the order of 7 kPa (1.0 psi) can
be expected to cause widespread window damage. It
may also cause minor architectural damage to unusually
weak structures. Overpressures of the order of 0.7 kPa
(0.1 psi) are not expected to cause damage except under
unusual circumstances. An example of design criteria
for window glass subjected to sonic boom is shown in
Fig. 6 (Pittsburgh Plate Glass Industries, 1969). A factor of safety of 2.5 is included in the recommendations.
The reader is advised to review applicable regulations,
laws, and ordinances pertinent to his location and type
of operation. For example, at the time of this writing,
the Office of Surface Mining requires compliance to an
airblast overpressure limit of 128 dB (0.0073 psi, 0.05
kPa) for residences.
There are several factors which are important in the
control of air waves. Atmospheric conditions are very
important in the above-ground transmission of air overpressures. In an area of concern, it is desirable to avoid
blasting during times when refraction, reflection, and fo-

Fig. 6. Design criteria for window glass subjected to


sonic boom.

cusing of air waves could build up air overpressures to


potentially harmful or annoying. values. If air waves are
a problem, isotherms, temperature inversions, wind
shears, and unfavorable wind directions should be
avoided. If human response is a matter of interest, it is
desirable to avoid nighttime and other quiet hours, and
to avoid weekend and holiday blasting. The choice of
products can influence the results also. Down-hole
electric detonators will generate lower overpressures
than surface connectors and detonating fuse. The use
of stemming can be used to reduce overpressures. In
many instances, there is a feasibility to using some sort
of barrier or isolation device, such as a shaft cover.
Air Waves Underground

The behavior of blast waves in air and in underground chambers and openings is far more complicated
than the behavior of ground vibrations. Through complex reflection and focusing, air overpressures may be
maintained at high values in an underground environment. To illustrate overpressures developed in a typical room-and-pillar mining operation, Olson and Fletcher
(1971) recorded air waves from three mine production
blasts initiated with conventional ?h-sec delay detonators. For that particular location and blasting system,
the overpressures could be expressed by the equation:
P = 4.9 X 10"Dl

w'3)-2.15

where P is overpressure in pounds per square inch, D is


distance from the blast in feet, and W is zero-delay
charge weight in pounds.
This writer has also found it useful to follow the concept of overpressures developed in vented chambers according to the volume of the chamber. Measurements
at the NORAD facilities (Smart, 1971) showed good
agreement with tests in partially closed chambers
(Weibull, 1968). See Fig. 7.

BLASTING

CHARGE -VOLUME

RATIO.

KG / ~3

Fig. 7 . NORAD overpressure measurements in comparison with vented chamber tests.

REFERENCES AND BIBLIOGRAPHY


Ambraseys, N.R., and Hendron, A.J., Jr., 1968, "Dynamic
Behavior of Rock Masses," Rock Mechanics in Engineering Practice, K.G. Stagg and O.C. Zienkiewicz, eds., John
Wiley and Sons, pp. 203-227.
Barber, D.B., Fourney, W.L., and Dally, J.W., 1979, "Fracture Control in Tunnel Blasting," University of Maryland.
Aug.
Barber, D.B., Fourney, W.L., and Dally, J.W., 1978, "Blasting Parallel Hole Cuts with Fracture Plane Control,"
Tunnels and Tunneling, May.
Devine, J.F., et al., 1966, "Effect of Charge Weight on
Vibration Levels from Quarry Blasting," RI 6774, US
Bureau of Mines.
Duvall, W.I., et al., 1963, "Vibrations from Instantaneous
and Millisecond-Delayed Quarry Blasts," RI 6151, US
Bureau of Mines.
Hendron, A.J., and Oriard, L.L., 1972, "Specifications for
Controlled Blasting in Civil Engineering Projects," Proceedings, First North American Rapid Excavation and
Tunneling Conference, Vol. 2, AIME, New York.
Langefors, U., and Kihlstrom, B., 1963, The Modern Technique o f Rock Blasting, Almqvist and Wiksell, Stockholm,
Sweden.
Nicholls, H.R., Johnson, C.F., and Duvall, W.I., 1971,
"Blasting Vibrations and Their Effects on Structures,"
Bulletin 656, US Bureau of Mines.

1603

Olson, J.J., and Fletcher, L.R., 1971, "Airblast Overpressure


Levels from Confined Underground Production Blasts,"
RI 7574, US Bureau of Mines.
Oriard, L.L., 1972, "Blasting Operations in the Urban Environment," Bulletin, Association of Engineering Geologists, Vol. 9, No. 1, Oct.
Oriard, L.L., 1972, "Blasting Effects and Their Control in
Open Pit Mining," Proceedings, 2nd International Conference on Stability in Own Pit Mining, AIME, New
York.
Oriard, L.L., 1973, "A Guide to Evaluate Damage Potential
to Pipelines from Nearby Construction Blasting," unpublished report to Alyeska Pipeline Service Co., Houston.
Oriard, L.L., 1979, "Blasting and Excavation," The Atlanta
Research Chamber, Report No. UMTA-GA-06-0007-79-1,
Dept. of Transportation, Washington, June.
Oriard, L.L., 1980, "Observations On the Performance of
Concrete at High Stress Levels from Blasting," Proceedings, Sixth Conference on Explosives and Blasting Techniques, Society of Explosives Engineers, Montville, Ohio.
Oriard, L.L., and Coulson, J.H., 1980, "TVA Blast Vibration
Criteria for Mass Concrete," Minimizing Detrimental
Construction Vibrations, Preprint No. 80-175, American
Society of Civil Engineers, New York.
Oriard, L.L., and Emmert, M.W., 1980, "Short-Delay Blasting at Anaconda's Berkeley Open-Pit Mine, Montana,"
SME Preprint 80-60, AIME Annual Meeting, Las Vegas,
NV, February.
Oriard, L.L., 1981, "Field Tests with Fracture-Control Blasting Techniques," 1981 RETC Proceedings, AIME, New
York, NY.
Oriard, L.L., 1981b, "Influence of Blasting on Slope Stability: State of the Art," Proceedings, Stability in Surface
Mining, Vol. 3, AIME, New York.
PPG Industries, 1969, "Glass Product Recommendations,"
Technical Service Report No. 101, Pittsburgh, PA.
Rathbone, T.C., 1963, "Human Sensitivity to Product Vibration," Product Engineering.
Reiner, H., and Meister, F.J., 1931, "Susceptibility of Human
Beings to Vibration," Engineering News Record, Mar.,
p. 470.
Smart, J.D., 1971, "Measurement of Blast Pressures in the
NORAD Cheyenne Mountain Complex During Excavation for the Expansion Project," Technical Report No. 14,
US Army Corps of Engineers.
Tart, R.G., Oriard, L.L., and Plump, J.H., "Blast Damage
Criteria for a Massive Concrete Structure," Minimizing
Detrimental Construction Vibrations, Preprint 80-175,
American Society of Civil Engineers, New York.
Weibull, H.W.R., 1968, "Pressures Recorded in Partially
Closed Chambers at Explosion of TNT Charges," Annals
o f the New York Academy o f Sciences, Vol. 152, Art. 1,
Oct.

Você também pode gostar