You are on page 1of 12

Signature Hole Blast Vibration Control – Twenty Years Hence and Beyond

Douglas A. Anderson, PhD, Parsons Brinckerhoff

Abstract

In the early 1980s, a method for controlling blast vibrations other than by modifying the Scaled Distance
came into use. Research studies had indicated that blast vibration could be simulated by detonating a
signature hole with the vibration monitored at critical locations, and then using a computer to superpose
the waveforms with varying delays. By choosing delay times that create destructive interference at
frequencies that are favored by the local geology, the “ringing” vibration that excites structural elements
in houses and annoys neighbors could be reduced.

At that time, pyrotechnic detonator firing times were less accurate than they are today, and electronic
detonators were only in the early stages of development. Since accurate delay times are crucial to
effective vibration control, scatter in the firing times limited the method severely.

As time has gone on, both the tools and the techniques have improved. Electronic detonators now have
scatter less than a millisecond. Computers, of course have increased in speed phenomenally, and
software keeps pace. In light of all these changes, many researchers and practitioners have extended and
modified the original work, finding both limitations and new potential.

This paper will review the history of the method, the assumptions made, the impact of increased
detonator accuracy, and successes and failures of the method. Implications for future control methods,
for both surface and underground blasting, will be discussed, as well as using vibrations to assess the
effectiveness of the explosives.

Copyright © 2008 International Society of Explosives Engineers


2008G Volume 2 - Signature Hole Blast Vibration Control - Twenty Years Hence and Beyond 1 of 12
Introduction
For as long as the ISEE (and its predecessor SEE) has been around, blast-generated vibration and
airblast have captured a lot of attention. Long before the Society’s inception, research and development
addressed the fundamental problem of controlling the wasted explosive energy that goes into vibration.
Early efforts focused on determining what “level” of vibration would cause annoyance or damage, and
what methods could be used to assure a blaster that he or she could work effectively without worrying
about neighbor problems.

Simple solutions to engineering problems are generally preferred; to reach a simple solution, it helps to
first simplify the problem. In the US, research by the Bureau of Mines from the 1940s and later
developed simple methods to classify and address the problem of damaging vibrations. As we all know,
peak particle velocity (PPV) was chosen as a descriptor that could be related to damage occurrences.
Research elsewhere came to similar conclusions, though the details varied. Regulations followed from
recommendations, which in turn followed from the research.

The research showed that by knowing distance D and charge weight per delay W, (Scaled
Distance = D / W ), one could reasonably predict the resultant vibration through a power law relation:

PPV = A ∗ SD − B

where A and B are constants. These “constants” tend to be site-specific, though they often don’t vary
significantly. Cumulative plots typically had (and still often have) scatter of an order of magnitude at a
given Scaled Distance.

Scaled Distance is a simple solution, but unfortunately it is also simplistic. The charge weight parameter
assumes that all explosives generate the same amount of vibration, irrespective of either their strength or
how they interact with the rock. The distance parameter assumes that the only effect of distance is to
produce decay or attenuation, with no effect of the geological path on the wave. These assumptions are
clearly not realistic, and are undoubtedly part of the reason for the scatter. The shape or duration of the
vibration wave (and thus its frequency spectrum) is not addressed by Scaled Distance. The frequency
spectrum is important because the vibration is not a scalar like temperature, but a propagating wave,
extended in space, moving in time, and changing as it propagates.

Briefly, when an impulse is applied to rock by an explosive charge, the impulse (on the order of a few
milliseconds long) propagates as waves away from the site of the detonation. These waves travel in
many directions, reflecting and refracting at the internal boundaries that permeate any real material.
These boundaries, (joints, fractures, bedding planes) also serve as new sites of wave initiation. At a
given location away from the blast, waves traveling a variety of paths from the initial source arrive and
add up, or superpose. The local geology, and not the Scaled Distance, is the primary factor, but not the
only one, affecting the final waveform amplitude and shape.

Until the 1970s, peak particle velocity as a criterion and Scaled Distance as a means of reducing peak
particle velocity were the main methods for controlling blast vibration. When vibrations that satisfied
the regulations still created substantial annoyance for neighbors of coal mines in Illinois and Indiana, the
Bureau conducted further research that showed that the frequency of vibration was also a factor, in both

Copyright © 2008 International Society of Explosives Engineers


2008G Volume 2 - Signature Hole Blast Vibration Control - Twenty Years Hence and Beyond 2 of 12
damage potential and annoyance. This research resulted in the famous US Bureau of Mines RI 8507
(Siskind, et al., 1980), with the “Z-curve” of peak particle velocity as a function of frequency. The
recommendations of RI 8507 have been the basis of regulations around the world.

In the years immediately following RI 8507, there were many efforts to modify blast vibration so that it
would satisfy the Z-curve. One path led to investigations of modifying the millisecond delay sequences
to affect vibration frequency, and ultimately, the amplitude.

My own introduction to this was in my first year at Martin Marietta Laboratories (1980), working with
Steve Winzer. I recall seeing vibration traces and spectra from North Carolina quarry blast
seismograms. Blasts with delays of 100 milliseconds (ms) between rows produced spectra that looked
1
like Mount Fuji, with a strong peak at 10 Hertz. 10 Hertz = . It was impossible to miss the
100ms
correlation.

I also recall that at this time I heard an SEE paper where it was stated that “every explosively loaded
hole is different.” I thought that this is certainly true in detail, but shouldn’t there be overall similarities?
At Martin Marietta Laboratories, we conducted tests of single hole shots measured at various locations,
and at different quarries. We found that that at fairly low frequencies, less than 30 Hz, the waveforms
were quite reproducible.

Others at the same time had similar ideas (Al Andrews, Otto Crenwelge, Klaus Hinzen), and the notion
of a signature waveform and linear superposition was born. I’ll discuss how this idea progressed
shortly, but first, will review what eventually developed as the “signature-hole” method, sometimes
called the “seed waveform” method.

Description of the Signature-hole Method


The underlying principle for the signature-hole method is that each hole in a blast will generate
essentially the same vibration at a given location. In effect, this assumes the following:

• All holes are detonated at the same location, so that the path traveled by the waves is identical.
• All holes have the same explosive charge type and weight.
• All holes have the same explosive-rock interaction, so that the source pulse is the same.

Of course these assumptions are never justified in reality – one would have to be blasting the same rock
over and over, a kind of Blaster’s “Groundhog Day”. However, the way most blasts are detonated
results in these assumptions being good enough in practice, as long as the distance from the blast is
fairly large relative to the area of the blast.

Once a signature has been obtained from a single-hole blast, the method uses time-lagged linear
superposition. A suite of possible delays is chosen, and for each of these delays, the signature is
repeated with the delay and summed for as many times as there are holes in the shot. This much is
familiar to most, and fairly standard. What varies is the way in which the resultant waveform is
analyzed.

Copyright © 2008 International Society of Explosives Engineers


2008G Volume 2 - Signature Hole Blast Vibration Control - Twenty Years Hence and Beyond 3 of 12
Often the desired result is to find delays that will result in the minimum peak particle velocity, given a
range of practical delays. This is the simplest of the analytical techniques, comparing the peaks on the
synthetic seismograms and finding which delay (or delays) produced the minimum.

As noted earlier, problems with vibration annoyance are not always related to high peak particle
velocities, but rather to a rolling waveform with a particular frequency (Siskind, et. al., 1992). Some of
the analysis methods focus on controlling the proportion of frequencies in the resultant spectrum. To do
this, a common approach is to use a 3D plot (like a topographic map) of delay versus frequency, with
vibration amplitude as the “elevation”. The aim, then, is to choose a delay that would avoid “ringing the
bell” created by the local geology.

With this understanding, I will discuss the development of the signature hole method.

Some History
The idea of using delays to control vibration is almost as old as millisecond delays themselves. In
Bureau of Mines Bulletin 442, Thoenen and Windes (1942) showed that delay times could be chosen to
cancel vibration waves from a test production shot. However, they concluded that:

It is difficult to decrease vibration by delay blasting because often the vibration has no
regular frequency, elimination of one component frequency does not appreciably affect
the resultant amplitude, and the timing method must be quite flexible, yet accurate.

It is concluded, therefore, that delay shooting of the type described in this paper is not
practical for reducing vibration in commercial practice.

One of the reasons Thoenen and Windes concluded “the vibration has no regular frequency” was due to
the scatter in the firing times. This scatter was so large that the resultant spectrum would change from
blast to blast, even with the same nominal firing times. The resultant amplitude couldn’t be changed
either, because the highest amplitude is often related to high-frequency vibration. Flexible and truly
accurate firing times were more than fifty years over the horizon.

Nevertheless, the correlation between firing times and the resultant frequencies was occasionally noted.
A decade after Thoenen and Windes’ work, Fish (1951) showed that the resultant vibration from small
test blasts could be modified by destructive interference. Another decade later, Frantti (1963) and
Pollack (1963) further documented the effect of delay on the resultant frequency of ground vibration. It
was clear that there was an association between the delay times and the frequencies. Greenhalgh (1980)
showed that this effect was persistent, being observable at 170 kilometers!

Obviously, some mechanism related delays and blast vibration. The goal was then to find a way to use
this feature to help design blasts “in commercial practice.”

In the 1980s, several studies developed the basic signature hole technique using linear superposition
(Anderson, et al. (1983, 1985), Andrews (1981), Crenwelge (1988), Hinzen, et al. (1987)). At that time,
the inaccuracy of the detonators, while less severe than in the days of Thoenen and Windes, still plagued
attempts to control vibration. Anderson and Reil (1987) discussed what was “good enough” to control
vibration, though this meant only affecting vibration with frequencies less than about 30 Hz. Altering

Copyright © 2008 International Society of Explosives Engineers


2008G Volume 2 - Signature Hole Blast Vibration Control - Twenty Years Hence and Beyond 4 of 12
the peak particle velocity was rather fortuitous, achieved regularly only if primarily low frequency
vibration was present.

Several commercial products were developed that relied on a signature hole as the underlying basis. As
discussed earlier, the key differences are related to the way in which the optimum delay sequence is
chosen – what constitutes the “best” vibration? Each developed their own distinctive approach, some
attempting to minimize peak particle velocity, others reducing the input of the dominant frequencies,
and still others using a combination of the two.

What has happened since then has been refinement of the technique, with the added benefit in recent
years of the control available with electronic detonators. A substantial literature has developed on the
topic.

Support and Skepticism


In researching the literature for this paper, I found about sixty papers from ISEE proceedings that
mentioned “signature”, “seed”, or “single-hole” waveforms for vibration analysis, as well as papers from
other publications. Most of the papers do not question the validity of the method per se, but rather they
point out modifications that might improve the method. The discussion has focused on three areas:

• How valid are the assumptions, such as reproducibility of signature shots?


• How practical is the application of the method?
• What are the physical principles behind the method?

As mentioned earlier, in the initial work I was involved in at Martin Marietta, we noted that the
waveforms from single hole shots were quite similar, similar enough that the gross variations did not
preclude use of linear superposition. Subsequently, the reproducibility has been studied in more detail
(Stump and Reinke, 1988; Yang et al., 1999). Yuill and Farnfield (2001) showed that there are small
differences in the waveforms from shot to shot, and that spectra derived from linear superposition match
those predicted, but peak levels are not adequately matched. A further paper (Yuill, et al., 2002) tried to
obtain the waveform by simple subtraction (the opposite of superposition) and the resultant waveform is
similar to the inverse of the signature. No clear explanation is available for this at present.

Practical applications often report case histories (Reamer, et al., 1993), and several have discussed how
low frequency vibration in particular could be modified (Siskind and Stagg, 2001; Siskind et al., 1992;
Roy et al., 2005). Others have cautionary tales having to do with application of the 8-millisecond
criterion to signature waveform analysis (Anderson, 1989; Moore and Richards, 2002; Spathis, 2006).

Also related to practical application, one of the objections is that it takes too much time to get from
signature shot to production blast. Wheeler (1989) and Harrison et al. (2005) investigated faster
methods of getting from signature hole to blast design. Because of the time delay, the method is
considered inappropriate for some construction blasting (e.g., trench blasting) where the blast location
changes often. Harrison et al. (2005) also investigated this problem.

Though most of the applications have been for surface mines or quarries, Murashita et al. (1998) have
investigated the use of the method for tunneling, where delay sequences are used that employ some

Copyright © 2008 International Society of Explosives Engineers


2008G Volume 2 - Signature Hole Blast Vibration Control - Twenty Years Hence and Beyond 5 of 12
long-period delays. Tunnel blasting (which is one of my main areas of work at Parsons Brinckerhoff)
and other specialized blast designs are still rather uncommon applications of the signature hole method.

Although many papers are rather practical, a few basically start from first physical principles (Yang,
1998; Tibuleac, et al., 2005), or discuss superposition methods from a more theoretical bent (Stump and
Reinke, 1988). These are often by researchers interested in other topics, such as discrimination of mine
blasts from nuclear tests, but have applications to the explosive industry.

A somewhat different approach has been used to look at traditional Scaled Distance relations in the light
of superposition studies. The work of Blair (2004) and Spathis (2006) on charge weight scaling takes
into account the effect of neighboring boreholes on each other in modifying the “signature” from each of
the holes. What constitutes independent firings is a key to such approaches, and the 8-millisecond
criterion still looms heavily.

All in all, this is clearly a fertile area of study from a wide variety of perspectives. What really has
created the possibility of effective use, though, is the advent of electronic detonators in commercial
practice.

Impact of Electronic Detonators


In this millennium, it has become clear that the advent of electronic detonators will revise many aspects
of blasting. Precise interaction of the stress waves generated by explosives will allow blasters to
understand how changes in other parameters, such as burden, spacing, explosive type, and delay
sequence will affect overall blast performance.

Not the least of these effects is the ability to reproduce and control the vibration from blasts, as has been
demonstrated repeatedly (Wheeler, 1991). High-speed cinematography also was used to show the
importance of accuracy in control of vibration using the signature shot method (Floyd and Conn, 1990).

Well before commercially available electronic detonators came into general use, there have been several
papers that used early versions of electronic detonators with signature hole technique for controlling
vibration (Chiappetta, et al., 1986; Hinzen and Lüdeling (1987); Sakamota et al., 1989; Pruss, 1989;
Konya, 1989). More recently, since electronic detonators have come into use, the usefulness in
controlling vibration using the signature hole technique has been discussed (McKinstry et al., 2002).

What has not been adequately discussed thus far is the necessity of using some kind of vibration control
technique when electronic detonators are used. At one time, the scatter in the initiator firing time was
used to assure that individual holes with the same delays did not actually fire at the same time. Such
was a false assurance, though, because one could not guarantee a particular scatter pattern, and it was by
definition not reproducible. This was particularly used in underground applications, where, for example,
one or two delays would be used for all perimeter holes. The long period delays almost guarantee non-
overlap of firing times, but "almost" is very imprecise.

Now with electronic detonators finding applications in many operations, a new approach is needed.
Since there is no noticeable scatter, electronic detonator blasts need to be redesigned so that multiple
holes do not detonate simultaneously. For example, Higgins and Riihioja (2004) discuss the software
needed for modeling with electronic detonators.

Copyright © 2008 International Society of Explosives Engineers


2008G Volume 2 - Signature Hole Blast Vibration Control - Twenty Years Hence and Beyond 6 of 12
An extension of the simultaneous detonation problem is when a resonance is enhanced by electronic
detonators when the wrong delay is chosen. Anderson and Reil (1987) showed that a blast with a delay
that had precisely the wrong period for the specific local conditions would often be worse than one with
some scatter, though not always. Signature hole vibration control may become a necessity with
electronic detonators, rather than an option.

Prospectus
In anticipating future developments, three major aspects come to mind

• Refinement of the overall signature hole analysis process


• Application of the process to new situations, such as tunneling.
• Using the vibration to assess blast performance.

The first two aspects have been discussed earlier, and work continues on them. One thing that has not
been looked at in detail is the use of vibration to assess how a blast performs.

The vibration generated by a blast tells a story in three parts:

• Detonation of the explosive, producing shock waves and gas pressure


• Interaction of the explosive shock wave with the rock around it, fragmenting some of the rock
• Propagation of the energy not absorbed in fragmentation as elastic waves.

The characteristics of the resultant vibration are due to the amplitude and shape of the pulse from each
of the holes, the time delay between holes, and the reflections and refractions as the wave propagates
away from the source. If you change any part of this, the resultant wave changes. There is thus a record
of the explosive interaction with the rock in each of the seismic records from the blast.

As we have discussed, the superposition is done by repeatedly time-lagging the signature waveform with
the appropriate delay, and adding the time-lagged waveform to the original. Mathematically, this is
equivalent to a process called convolution. Convolution of two functions involves multiplication of the
elements of one array with elements of a second and then summing the results. In this case, one array is
what we will call a “comb function”, and the other array is the signature waveform.

The comb function is a time sequence like the signature waveform. For starters, we will assume that the
function has values of 1 at the firing times, and 0 at all other times. An example is shown in the Figure
1, for eight holes with a 50 ms delay (note that there is a point at 0 time). Changing the firing times
Figure 1. Comb Function Figure 2. Signature Wave

1.2 10

8
1
6
Particle Velocity (in/sec)

0.8 4

2
Weight

0.6
0

-2
0.4
-4

0.2 -6

-8
0
0.000 0.050 0.100 0.150 0.200 0.250 0.300 0.350 0.400 0.450 0.500
-10
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Time (sec)
Time (sec)

Copyright © 2008 International Society of Explosives Engineers


2008G Volume 2 - Signature Hole Blast Vibration Control - Twenty Years Hence and Beyond 7 of 12
moves the teeth of the comb left or right, and changing the relative weight of the contribution of each of
the holes changes the height of the teeth. Let’s further assume we have a signature waveform as in
Figure 2.

Convolving the comb function with the signature wave (where “ ⊗ ” denotes convolution),

Wave = Comb ⊗ Signature

generates a waveform as shown in Figure 3.


Figure 3. Summed Wave

10

6
Particle Velocity (in/sec)

-2

-4

-6

-8

-10
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Time (sec)

Let’s assume that one hole does not fire – one of the “teeth” in the comb is missing (Figure 4). The
resultant waveform is as shown in Figure 5.
Figure 4. Comb Function -- One Tooth (Hole) Missing Figure 5. Summed Wave -- One Tooth (Hole) Missing

1.2 10

8
1
6
Particle Velocity (in/sec)

4
0.8
2
Weight

0.6 0

-2
0.4
-4

-6
0.2
-8

0 -10
0.000 0.050 0.100 0.150 0.200 0.250 0.300 0.350 0.400 0.450 0.500 0 0.2 0.4 0.6 0.8 1 1.2 1.4
Time (sec) Time (sec)

There is clearly a substantial difference from the original summed waveform. Visually (as some blasters
do) you can see that something is wrong in the middle of the shot, but what actually is wrong is not
clear. A second example is a hole firing out of sequence – the timing off, and the amplitude higher
because of lack of adequate relief (Figures 6 and 7).

Copyright © 2008 International Society of Explosives Engineers


2008G Volume 2 - Signature Hole Blast Vibration Control - Twenty Years Hence and Beyond 8 of 12
Figure 6. Comb Function -- One Tooth (Hole) Shifted and Higher Figure 7. Summed Wave -- One Tooth (Hole) Shifted and Higher

1.6 10

1.4 8

6
1.2

Particle Velocity (in/sec)


4
1
2
Weight

0.8 0

-2
0.6
-4
0.4
-6
0.2 -8

0 -10
0.000 0.050 0.100 0.150 0.200 0.250 0.300 0.350 0.400 0.450 0.500 0 0.2 0.4 0.6 0.8 1 1.2 1.4
Time (sec) Time (sec)

In this case, it is not immediately apparent that something is wrong, but the information is there. What
is even more important is that, mathematically, a process called “deconvolution” can recover the comb
function from the waveform, assuming the same seed waveform is valid.

I have played around with this concept a bit. It is not quite as simple as it sounds, because the seed
waveform is not identical for all holes, and of course timing is not usually exact. In principle, though,
one could find out how individual holes in a blast perform from the resultant seismograms. The method
is much more robust if a number of seismograms are available, since they are recording the same
information about the blast performance – the comb function is the same for all of the seismograms,
regardless of the seed waveform shape.

There is great potential in this method of blast analysis, and with the use of electronic detonators, this
potential can now be realized, with some persistent effort.

Conclusion
The signature hole method has become part of the standard repertoire in controlling vibrations from
blasting. The increased use of electronic detonators will certainly facilitate, if not necessitate, the use in
many applications. While primarily used in surface blasting, the use in underground operations will also
likely increase.

The need for caution in its application, of course, is necessary. Assumptions must be understood, and
proper application of recommendations that are consistent with safe blasting

In the future, we can look forward to more use of the signature wave method, perhaps even
understanding and using the information collected in seismograms that are now used only for liability
protection.

Acknowledgments
In a review paper, it is appropriate to acknowledge those whose work has contributed to the overall
development of the subject. I have cited many, but certainly not all, of the work done in this field. Any
papers or contributions missed are not intentionally ignored. Thanks to all for their good work.

Copyright © 2008 International Society of Explosives Engineers


2008G Volume 2 - Signature Hole Blast Vibration Control - Twenty Years Hence and Beyond 9 of 12
I believe that special mention must be made of those in the (former) US Bureau of Mines who worked
diligently to understand the mechanisms of blast-induced ground vibration, and measures to control it.
This paper is dedicated to them.

I would also like to thank Kyle Ott of Parsons Brinckerhoff for a review of this paper.

References
Anderson, D.A., S.W. Winzer, and A.P. Ritter (1983), “Synthetic Delay versus Frequency Plots for
Predicting Ground Vibration from Blasting”, Proceedings of the 3rd International Symposium on
Computer-Aided Seismic Analysis and Discrimination, Washington, DC, 70-74 (1983)

Anderson, D.A., S.W. Winzer, A.P. Ritter, and J.W. Reil (1985),“A Method for Site-Specific Prediction
and Control of Ground Vibration from Blasting”, Proceedings of the 11th Annual Conference on
Explosives and Blasting Techniques, San Diego, CA (1985)

Anderson, D.A., and J.W. Reil (1987), “Effect of Cap Scatter on Blast-Induced Ground Motion”,
Journal of Explosives Engineering, 5, 22-28

Anderson, D.A., and J.W. Reil (1988), “Measuring fragmentation efficiency of a blast using ground
vibration”, Proceedings of the 14th Annual Conference on Explosives and Blasting Techniques,
Anaheim, CA, 60-71.

Andrews, A.B. (1981), “Design criteria for sequential blasting,” Proceedings of the 7th Annual
Conference on Explosives and Blasting Techniques, Phoenix, AZ, 173-192.

Blair, D.P. (2004), “Charge weight scaling laws and the superposition of blast vibration waves,”
Fragblast, 8, 221-239.

Chiappetta, R. F., S.L. Burchell, J.W. Reil, and D.A. Anderson, (1986), "Effects of accurate ms delays
on productivity, energy consumption at the primary crusher, oversize, ground vibrations and airblast",
Proceedings of the 12th Annual Conference on Explosives and Blasting Techniques, Atlanta, GA, 213-
234.

Crenwelge, O.E., Jr. (1988), “Use of Single Charge Vibration Data to Interpret Explosive Excitation and
Ground Transmission Characteristics”, Proceedings of the 14th Annual Conference on Explosives and
Blasting Techniques, Anaheim, CA, 151-160.

Fish, B.G. (1951), “Seismic vibrations from blasting,” Mine and Quarry Eng., 17, 189.

Floyd, J.L., and D.B. Conn, (1990), “PhotoSeis ™”, Proceedings of the 16th Annual Conference on
Explosives and Blasting Techniques, Orlando, FL, 73-82.

Frantti, G.E. (1963), “Spectral energy density for quarry explosions,” Bull. Seismol. Soc. Am., 53, 989-
996.

Copyright © 2008 International Society of Explosives Engineers


2008G Volume 2 - Signature Hole Blast Vibration Control - Twenty Years Hence and Beyond 10 of 12
Greenhalgh, S. (1980), “Effects of delay shooting on the nature of P-wave seismograms,” Bull. Seismol.
Soc. Am., 70, 2037-2050.

Harrison, D., D. Bartholomae, and E.J. Walter, Jr. (2005), “A Study Of The Use Signature Holes For
Smaller & Faster Moving Construction Blasting Operations - Can It Work & Can It Fit Your Current
Magazine?”, Proceedings of the 31st Annual Conference on Explosives and Blasting Techniques,
Orlando, FL.

Higgins, M., and K. Riihioja (2004), “Design software for electronic detonators”, Proceedings of the
30th Annual Conference on Explosives and Blasting Techniques, New Orleans, LA.

Hinzen, K.-G., and R. Lüdeling (1987), “A new approach to predict and reduce blast vibration by
modeling of seismograms and using a new electronic initiation system,”, Proceedings of the 13th Annual
Conference on Explosives and Blasting Techniques, Miami, FL, 144-161.

Konya, C.J. (1989) “High precision cap accuracy – an independent study”, Proceedings of the 15th
Annual Conference on Explosives and Blasting Techniques, New Orleans, LA, 419-428.

McKinstry, R., J. Floyd, and D. Bartley (2002), “Electronic detonator performance evaluation,”
Proceedings of the 28th Annual Conference on Explosives and Blasting Techniques, Las Vegas, NV,
123-130.

Moore, A.J., and A.B. Richards (2002), “Time window vibration control techniques – Cautionary tales
for explosives engineers”, Proceedings of the 28th Annual Conference on Explosives and Blasting
Techniques, Las Vegas, NV, 363-380

Murashita, T., N. Sakuma and M. Sakamoto (1998), “Blasting in tunneling”, Proceedings of the 24th
Annual Conference on Explosives and Blasting Techniques, New Orleans, LA, 347-357.

Pollack, R.N. (1963), “Effect of delay time and number of delays on the spectra of ripple-fired shots,”
Earthquake Notes, 34, 1-12.

Pruss, K. (1989), “‘Frequency adjustment’ with high accuracy detonators” Proceedings of the 15th
Annual Conference on Explosives and Blasting Techniques, New Orleans, LA, 123-130.

Reamer, S.K., K-G. Hinzen, and Y Sifre (1993), “Case Studies in the Application of Firing Time
Optimization,” Proceedings of the 19th Annual Conference on Explosives and Blasting Techniques, San
Diego, CA, 281-292.

Roy, M. P., A. K. Sirveiya & P. K. Singh (2005), “Low frequency long duration blast vibrations and
their effect on residential structures”, Proceedings of the 31st Annual Conference on Explosives and
Blasting Techniques, Orlando, FL.

Sakamoto, M., M. Yamamoto, K. Aikou, E. Suzuki, H. Fukui, and K. Ichikawa (1989), “A study on
high accuracy delay detonator”, Proceedings of the 15th Annual Conference on Explosives and Blasting
Techniques, New Orleans, LA, 185-195.

Copyright © 2008 International Society of Explosives Engineers


2008G Volume 2 - Signature Hole Blast Vibration Control - Twenty Years Hence and Beyond 11 of 12
Siskind, D.E., Stagg, M.S., Kopp, J.W., and Dowding, C.H. (1980), “Structure response and damage
produced by ground vibration from surface mine blasting,” U.S. Bureau of Mines RI 8507.

Siskind, D.E., V.J. Stachura, and M.J. Nutting, (1992), “Low-frequency vibrations produced by surface
mine blasting over abandoned underground mines”, RI9078, US Bureau of Mines, 54 pp.

Siskind, D.E., and M.S. Stagg (2001), “Vibration and structure response from Dade County quarry
blasting”, Proceedings of the 27th Annual Conference on Explosives and Blasting Techniques, Orlando,
FL, 185-197.

Spathis, A.B. (2006), “A scaled charge weight superposition model for rapid vibration estimation,”
Fragblast, 10, 9-31.

Stump, B. W. and R. E. Reinke, 1988, “Experimental confirmation of superposition from small-scale


explosions”, Bull. Seism. Soc. Am. 78, 1059-1073.

Thoenen, J.R., and S.L. Windes (1942), ``Seismic effects of quarry blasting,'' U.S. Bureau of Mines
Bulletin 442, 83 pp.

Tibuleac, I.M., D.H. Thompson, and J.L. Bonner (2005) “Modeling ground motion in 3D geologic
media from fragmentation explosions: Preliminary results”, Proceedings of the 31st Annual Conference
on Explosives and Blasting Techniques, Orlando, FL.

Wheeler, R.M. (1989), “Controlling Blast Vibration Effects With On-Site Analysis of Single Hole
Signatures ‘A New Approach’”, Proceedings of the 15th Annual Conference on Explosives and Blasting
Techniques, New Orleans, LA, 123-130.

Yang, X., 1998, “MineSeis – A MATLAB GUI Program to Calculate Synthetic Seismograms from a
Linear, Multi-shot Blast Source Model”, LAUR-98-1486.

Yang. X., B.W. Stump, and J.D. Smith, “Ground vibration from single-hole cast blasts,” Journal of
Explosives Engineering, 16, 36-41.

Yuill, G., and R. Farnfield (2001), “Variations in Vibration Signals from single-hole quarry blasts,”
Proceedings of the 27th Annual Conference on Explosives and Blasting Techniques, Orlando, FL, 309-
317.

Yuill, G., W. Birch, and R. Farnfield (2002), “Isolating Single Hole Vibration Signals from Multi-Hole
Shots,” Proceedings of the 28th Annual Conference on Explosives and Blasting Techniques, Las Vegas,
NV, 231-242

Copyright © 2008 International Society of Explosives Engineers


2008G Volume 2 - Signature Hole Blast Vibration Control - Twenty Years Hence and Beyond 12 of 12

You might also like