Está en la página 1de 13

Blasting In Environmentally Sensitive Areas A workshop to discuss and review current and future practice

Fallacies in blast vibration analysis


A T Spathis Orica Australia PO Box 196 Kurri Kurri NSW Australia 2327 Abstract I contend that there are a number of fallacies applied in blast vibration analysis. Eight fallacies are demonstrated by counter examples ranging from the simple to the more subtle and complex. In most cases, the fallacies have an intuitive appeal but lack the rigour of sound analysis. Some of the fallacies may provide a focal point for future research in blast vibration analysis. Introduction Blast vibration analysis has a long history. Much of the research has been associated with the U.S Bureau of Mines (USBM). Siskind (2000), in a concise and comprehensive summary of that research, provides a valuable insight to the approach taken by the USBM. The USBM work was primarily concerned with taking measurements of blast vibrations generated from mines, quarries and construction sites. The charge weight scaling 'laws' used to fit the wide range of data have become entrenched as a practical, first pass prediction method for blast vibrations. These laws are used for both ground vibration and airblast predictions. Furthermore, the USBM work on assessing blast-induced damage, resulted in frequency dependent criteria using a combination of particle velocity and particle displacement for safe blast vibration limits. These safe levels were designed to prevent even minor threshold cracking in low level buildings and residences. A number of workers have studied various aspects of blast vibration. Blair (1987) established an agenda for a number of important aspects including: detector calibration and ground coupling, a Monte Carlo-Fourier method for superposition modelling and the importance of numerical methods such as the dynamic finite element method in establishing the primary characteristics of blast vibrations. The most recent implementation of a superposition model and related issues are discussed in Blair (1999). Other workers have adopted or modified the approaches cited above, or generated their own approaches to blast vibration analysis. Unfortunately, some of that work has either ignored or misunderstood some of the caution required in performing such analyses. The present paper identifies some fallacies, misconceptions or half-truths which persist in blast vibration analysis. Some of these are self-evident to some practitioners and, indeed, have been considered individually. Others are more complex, or at least, delve into engineering matters not commonly experienced by those who have to deal with the consequences, and need a little extra effort for understanding. Here, I present the fallacies and provide a counter example for each one. It is my purpose to bring these together and encourage discussion and debate, and in so doing enable some learning.

28 October 2001 Hunter Valley, Australia

Blasting In Environmentally Sensitive Areas A workshop to discuss and review current and future practice

The Fallacies It is best to commence with a simple definition of what constitutes a fallacy. The Macquarie Concise Dictionary (1988) defines a fallacy as "a deceptive, misleading or false notion [or] belief ". In the present context, it may be part of conventional wisdom and to some extent reflects the need to have a somewhat deeper understanding of the original premise and/or of alternative engineering methods. The following is a list of the fallacies which will be considered in turn. Examples of the fallacies are not presented as such; indeed, in some cases, the fallacies are rare. Essentially, the fallacies are used to provide a focal point for discussion. Each fallacy will be simply stated and a counter example(s) will be provided. The example will be used to elaborate and explain the fallacy further. Before presenting the fallacies, it might do well to consider the following: " The invalid assumption that correlation implies cause is probably among the two or three most serious and common errors of human reasoning." (Gould, 1981). In an effort to extract knowledge from data in blast vibration many of us fall into this trap. Fallacy 1: There is only one peak particle velocity used in blast vibration analysis. The measurement of blast vibrations in the ground involve particle displacement, velocity or acceleration. These quantities are vectors and may be resolved into three orthogonal components in a range of curvilinear coordinate systems including rectangular, cylindrical and spherical. The most common is the rectangular coordinate system and the components are referred to as radial (R) or longitudinal (in the direction of the horizontal line joining the measurement location to the blast source), the transverse (T) (which is horizontal and perpendicular to the radial) and the vertical (V). These may be oriented in a right-handed or left-handed coordinate system. If we consider a point, P, on the ground as a blast vibration passes it, the point, P, will move in a continuous curve dictated by the nature of the seismic wave. At any point in time, the kinematic quantity (say, velocity) may be resolved into the R, T and V components. These components form the standard records with which we are concerned. Each of the components will be a time record which has both positive and negative peaks. The maximum amplitude (positive or negative) measured from rest is called the peak particle velocity (ppv) for that component. Note that, in general, the ppv of each component does not occur at the same time. Furthermore, the vector amplitude may be formed as a function of time as,

vector (t ) = R 2 (t ) + T 2 (t ) + V 2 (t )

(1)

where the calculation is performed at each time step for the complete vibration record. The vector peak particle velocity (vppv) is the peak amplitude of the vector quantity defined in Equation (1). The vppv represents the maximum excursion in particle velocity of the point, P, from its resting position. There is only one vector peak particle velocity. Thus, when we speak of the peak particle velocity we must identify if we are referring to an individual component, the maximum of the peak particle velocity of all the components or the vector. Dowding (1985) also refers to a "maximum vector sum" which is calculated by,

28 October 2001 Hunter Valley, Australia

Blasting In Environmentally Sensitive Areas A workshop to discuss and review current and future practice
2 2 2 mppv = Rmax + Tmax + Vmax

(2)

where the ppv of each component is used to form the maximum vector peak. This is not strictly a vector quantity and is rarely used. Finally, it is worth mentioning that another peak particle velocity measure needs to be considered when there are multiple blast vibration records and the data needs to be considered in aggregate. An approach is to consider envelopes of the complete time distribution of all the vibration data set. Siskind (2000) recommends that vector sums are not used. Presumably this arises from empirical observations that damage and cracking in structures is most often related to the individual components (Dowding, 1985). Most man-made structures are constructed in a rectangular coordinate system and one would imagine that vibrations resolved in those directions should be examined for design and control needs. However, massive rock surrounding a blast will have a failure mechanism and criteria which may require a combination of the vibration components to be used. Fallacy 2: If particle velocity decreases then particle displacement and particle acceleration decrease. The focus on particle velocity in ground vibration analysis arises in the main for two reasons. Firstly, it is relatively simple to measure using either geophones whose output is usually particle velocity or accelerometers whose output is integrated to estimate the particle velocity. Secondly, and perhaps more importantly from an engineering perspective, blast induced damage is often associated with particle velocity. Here, damage refers to the reduced ability of a structure or rock mass to support loads. This reduced capacity is often inferred from particle velocity but is best measured by observing displacements and strains directly. We need to recall the basic relationships between displacement (s), velocity (v) and acceleration (a) where each is a function of time (t):

v=

ds dt

(3)

a=

d 2s dt 2

dv dt

(4)

and these differential equations may be integrated to recover the other quantities, depending on which kinematic quantity is measured. For example, if the particle velocity is measured, the particle acceleration may be calculated using Equation (4) and differentiating the velocity while the displacement may be calculated by integrating the measured velocity as indicated by Equation (3). It is important to realise that these transformations can recover permanent offsets (full body displacement, velocity or acceleration), provided the frequency response of the measurement system extends from zero to the highest frequencies contained in the quantity (while theoretically the upper limit is infinite frequency). Most measurement systems do not have a

28 October 2001 Hunter Valley, Australia

Blasting In Environmentally Sensitive Areas A workshop to discuss and review current and future practice

frequency response which extends down to zero frequency. This is usually not an issue for far-field vibrations where no permanent offsets are observed due to the (almost) elastic behaviour of the ground. At this stage it is worthwhile using Equations (3) and (4) for the special case of a pure harmonic vibration. This has a bearing on the present fallacy, but also on Fallacy 3. Suppose the particle displacement is described by,

s = A cos(2 f t )

(5)

where A is the amplitude, f is the frequency and t is the time. Then the particle velocity is given by,

v = A(2 f )sin (2 f t )
and the particle acceleration by, a = A 4 2 f 2 cos(2 f t )

(6)

(7)

Again, given one kinematic quantity, the others may be calculated. The relationships are summarised in Table 1 below. The relationships appear quite simple but there is a good deal of care required to recover one quantity from another. We can see that the amplitude of two of the displacement, velocity or acceleration depend on the product of an amplitude and frequency or frequency squared. Suppose the amplitude, A increases but the frequency decreases sufficiently, then the displacement would increase but the velocity and acceleration would both decrease. This is a first counter example.

Table 1. Relationships between particle displacement, velocity and acceleration. The general relationship is given first, followed by the case for a pure harmonic wave ignoring the phase. Given Kinematic Quantity s Calculated Kinematic Quantity v ds

s no change

a
d 2s dt 2 4 2 f 2 s dv

dt 2 f s

v dt
v
v 2 f

no change

dt

2 f v

a dt
a
a 4 2 f
2

a dt
a 2 f

no change

28 October 2001 Hunter Valley, Australia

Blasting In Environmentally Sensitive Areas A workshop to discuss and review current and future practice

It is simple to demonstrate another counter example for Fallacy 2. Consider a particle displacement pulse given by,

s = At exp( bt )

(8)

where A and b are constants. A normalised pulse shape is shown in Figure 1. This waveform has the desirable characteristics of being readily differentiated and integrated analytically. By using Equations (3) and (4), it is possible to derive the analytical expressions for the particle velocity and acceleration, respectively. These are not given here. However, some interesting characteristics of these expressions are provided in Table 2. These are the maxima (maximum positive values) and minima (maximum negative values), the peak amplitude (maximum irrespective of sign), and the values at zero and infinite times.

1.0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 10 20 30 40 50 60

Figure 1. Normalised displacement pulse given by Equation (8)

Table 2. Characteristics of the kinematic quantities using a displacement pulse described by Equation (8) Parameter maximum positive value maximum negative value peak amplitude value at zero time value at infinite time s v a Ab 3 at t = e3 b

A1 be

at t =

1 b
-

A at t = 0

0 at t = 0

A e2

at t = A A 0

2 b

-2Ab at t = 0 2Ab -2Ab 0

A1 be 0 0

28 October 2001 Hunter Valley, Australia

Blasting In Environmentally Sensitive Areas A workshop to discuss and review current and future practice

Suppose a displacement pulse, X, has A=10 and b=2. Then, the corresponding velocity and acceleration peak amplitudes will be 10 and 40, respectively. Now consider a displacement pulse, Y, with A=5 and b=1. Y will have exactly the same displacement peak amplitude as X but the corresponding velocity and acceleration peak amplitudes will be 5 and 10, respectively. So here we have a situation where both the velocity and acceleration peak amplitudes have decreased but the displacement amplitudes have remained identical. Other scenarios can be created based on Table 2 which further demonstrate that increasing or decreasing one of peak particle displacement, velocity or acceleration do not show the same trend for the other quantities. Fallacy 3: You may ignore the frequency content of blast vibrations. The use of charge weight scaling laws ignores the frequency content of vibrations. Usually, those engaged in assessing the impact of vibrations consider the frequency content separately. Most standards endeavour to include aspects of the frequency content of blast vibrations. The frequency content of vibrations may or may not be important in a particular situation. The importance of the frequency content of blast vibrations is amply demonstrated by the relationships given in Table 1 for pure harmonic motion. As stated earlier, field measurements are usually particle velocity or acceleration. When the particle displacement is calculated from these, we see that it involves a division by frequency terms, f and f 2. For earthquakes this is significant as frequencies are often less than 1 Hz and an amplification occurs. Even in the near field of blasts, these effects can be significant: for a peak particle velocity of 1000 mm/s at 10 Hz, the peak particle displacement (for a pure harmonic wave) will be about 16 mm which may damage brittle materials such as concrete and rock. Dowding (1985) quotes the range of typical blast parameters given below in Table 3. Table 3. Range of typical blast parameters (Cording (1975) as quoted by Dowding, 1985). Parameter Particle displacement Particle velocity Particle acceleration Pulse duration Wavelength Frequency Strain Typical Range 10 -4 to 10 mm 10 -4 to 10 3 mm/s 10 to 10 5 mm/s/s 0.5 to 2 s 30 to 1500 m 0.5 to 200 Hz 3 to 5000 microstrain

The frequency content of blast vibrations is important because it dictates the relationship between the various kinematic quantities. Fallacy 2 is also controlled by the frequency content of the vibrations. Low frequencies are observed far from a blast due to selective attenuation of higher frequencies by the ground. Such frequencies are responsible for significant displacement in the ground and of structures surrounding the blast. Many structures have resonant frequencies in the band associated with blasting. Modern electronic detonator systems offer the potential to control the frequencies generated by blasting and hence minimise the risk of inducing resonance in structures.
28 October 2001 Hunter Valley, Australia

Blasting In Environmentally Sensitive Areas A workshop to discuss and review current and future practice

Fallacy 4: It is not necessary to use Fourier analysis to determine the frequency content of blast vibrations. It appears that Fourier analysis is not in the usual armoury of tools used by engineers who need to deal with blast vibrations. Practitioners have tended to advocate simple time-domain measures where the pseudo-period of an assumed pure harmonic wave is used to estimate the frequency (given by the inverse of this measured period). A simple counter example is provided in Figure 2 and was suggested by Blair (2001). A portion of a theoretical periodic wave is shown and the zero crossings are indicated by the transitions of the superimposed trace which has unit peak amplitude. For the theoretical periodic wave, the durations between adjacent zero crossings are: 1.346, 0.368, 1.714, 0.612, 1.102, 1.591, 0.122, 1.714, 0.857. How should one proceed ? Clearly, if we associate these durations with half the period of a pure harmonic wave we obtain many different frequencies. Are these all present in the waveform ? In fact the waveform was generated by simply adding two sinusoids,

z (t ) = sin (3t )+ sin (2t / 3)

(9)

with frequencies equal to 3/2 (0.477) and 1/3 (0.106) and half-periods of 1.047 and 4.712, respectively. Neither of these half-periods appear in the zero crossing analysis of the signal. In fact, it may be shown that the zero crossings arise from frequencies which are sums and differences of the two constituent frequencies.
2.1 1.7 1.3 0.9 0.5 0.1 -0.3 -0.7 -1.1 -1.5 -1.9 0 1 2 3 4 5 6 7 8 9 10

Figure 2. Zero crossings for a theoretical periodic wave. Objections could be raised that this particular example does not have the characteristics usually seen in real blast vibration data. Another simple example should suffice. Again, a theoretical waveform and its zero crossings are shown in Figure 3. The duration between adjacent zero crossings are approximately 29, 31, 3, 17, 107 and 151 milliseconds. Again the calculated frequencies assuming that these durations are the half-period of pure harmonic waves range between 3.3 to 167 Hz. In fact, the waveform is the impulse response of an 8th order Butterworth filter with cutoff frequencies of 3 and 30 Hz. The amplitude spectrum is maximally flat between these two frequencies (Spathis, 1983).

28 October 2001 Hunter Valley, Australia

Blasting In Environmentally Sensitive Areas A workshop to discuss and review current and future practice

1.0 0.8 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 -0.8 -1.0 0 0.1 0.2 0.3 0.4 0.5 0.6

Figure 3. Zero crossings for an idealised blast wave. The two counter examples given here are at two extremes: the first is a signal consisting of two sine waves with just two spectral values while the second is a broad band pulse with a flat spectrum between two frequencies. Real blast vibrations consist of neither of these, but rather are more complex and have a range of spectral values of different amplitude. Fourier analysis is the only secure means for estimating the spectral content of blast vibrations. While such analysis is readily applied using standard techniques, it does require care. Several blast vibration equipment manufacturers can provide software for performing Fourier analysis. One word of caution. If a Fourier analysis of a blast vibration record indicates a single frequency peak, it is wise to consider possible resonances in the mount and/or coupling of the transducer to the ground. It may be that the narrow spectrum is an artefact from poor placement of the vibration transducer. Fallacy 5: Charge weight scaling laws are a good predictor of blast vibrations. Charge weight scaling laws are empirical fits to blast vibration data. Most, if not all, blast vibration data which has been fitted using Equation (10) below, show enormous variation. It is not unusual to observe peak particle velocity or airblast values spanning at least a factor of two and even one order of magnitude (that is, a factor of 10) for a given scaled distance. The reader is referred to Siskind (2000), Dowding (1985) and the annual conferences of the International Society of Explosives Engineers for many examples. The typical form of charge weight scaling laws is,

R z = A 1/ n W

(10)

28 October 2001 Hunter Valley, Australia

Blasting In Environmentally Sensitive Areas A workshop to discuss and review current and future practice

where A, b and n are constants and the measured quantity is z. W is the charge weight per delay and R is the range. Square root scaling applies when n=2 and cube root scaling when n=3. The laws have some foundation based on dimensional analysis as described by Dowding (1985). Langhaar (1951) observes that "The application of dimensional analysis to any particular phenomenon is based on the assumption that certain variables, which are named, are the independent variables of the problem, and that all variables, other than these and the dependent variable, are redundant or irrelevant. This initial step - the naming of the variables - often requires a philosophic insight into natural phenomena." The variables quoted by Dowding (1985) are shown in Table 4. He argues that the variables considered in a dimensional analysis of explosion phenomena can be reduced to the three terms,

pps W , and ppv R R3


where pps and ppv represent peak particle displacement and velocity, respectively. Dowding (1985) indicates that some blasting variables cannot be accounted for by these dimensionless parameters; in particular, the coupling of the explosive to the rock and the wave type. In essence, this is a reflection of the fact charge weight scaling laws are not mechanistic. In reducing the variables to just three, it is assumed that the observed variation in rock density and seismic velocity are much less than the other parameters involved (range and charge weight per delay). Again to quote Langhaar (1951) referring to dimensional analysis, "With little effort, a partial solution of nearly any problem is obtained. On the other hand, a complete solution is not obtained, nor is the inner mechanism of a phenomenon revealed by dimensional reasoning alone." It is worth noting that accepted attenuation behaviour of waves has been shown to be generally inconsistent with charge weight scaling laws (Blair, 1987). Dowding (1985) also supports this observation. Table 4. Variables considered in a dimensional analysis of explosion phenomena (as quoted by Dowding, 1985) Variable
Independent Energy released by the explosion per delay Distance to explosive Seismic velocity of rock mass Density of rock mass Time Dependent Maximum ground displacement Maximum ground velocity Maximum ground acceleration Frequency of ground motion

Symbol
W R c t pps ppv ppa f

Dimension
force x length length length/time mass/volume time length length/time length/time/time 1/time

28 October 2001 Hunter Valley, Australia

Blasting In Environmentally Sensitive Areas A workshop to discuss and review current and future practice

Despite their limitations, charge weight scaling laws are usually used by practitioners as a first starting point for a conservative prediction of the likely vibration from a blast. The USBM work in this area focused on providing upper limits for peak particle velocities and airblast peak values. The large scatter observed in charge weight scaling data suggests a prudent approach is to collect site specific data with relevant range and charge weight. Fallacy 6: A single seed waveform is adequate for use in superposition models which predict blast vibrations. An alternative to charge weight scaling laws for blast vibration prediction is the superposition model. Here, a recorded seed or signature waveform is measured from a typical blasthole. A prediction for the blast vibration emanating from a blast is constructed by summing scaled, suitably delayed, versions of the seed waveform. The blasthole delays and charge weights form the primary input to such superposition models. The model parameters may be calibrated using measured blast vibrations from blasts located in the same ground type. Spathis (1990) shows ground vibration data for five holes of equal charge weight initiated by electronic delay detonators. The amplitudes varied by up to a factor of three. Yuill and Farnfield (2001) show vibration data with an amplitude variation of two for almost identically charged single blastholes. The general variability of charge weight scaling data demonstrates that identically charged blasts are unlikely to generate identical vibration levels. Stump and Reinke (1987) present vibration data measured at different azimuths and ranges from a buried TNT source. The vibration amplitudes varied by factors of 10%, 50% and 330% for the radial, vertical and transverse peak accelerations, respectively. Such variations in single blasthole seed data suggest the need for a statistical approach. Blair (1999) has developed a statistical superposition model based on both Monte-Carlo and Fourier techniques. The initial work examined the implications of accurate delay intervals (Blair, 1987). The present version of the model incorporates many features including statistical scatter in delay times, random fluctuations in individual seed waveform amplitudes, delays due to propagation distance between individual blastholes and the monitoring station, and blasthole screening effects due to earlier firing blastholes situated between the current firing blastholes and the monitoring station. All of these effects have been verified experimentally. Fallacy 7: The 8ms rule applies universally. The application of charge weight scaling laws requires a choice of the charge weight per delay. The early USBM work identified the charge to be used was the maximum charge which was (nominally) initiated within an 8 ms window. Anderson (1989) gives a clear description of the original work and Siskind (2000) clarifies the original USBM work as being a guide and site-specific. Blair (1990) has demonstrated that in the far-field, where the effects of attenuation broaden the vibration waveform, it is the total charge in the blast that needs to be used. Of course, in the context of a superposition model, the 8 ms window has no need for existence (Blair, 1999). While the number of decks or blastholes initiating within a particular window may be estimated based on nominal delay times, the superposition model produces a prediction which accounts for all essential time delays and the particular seed waveform with

28 October 2001 Hunter Valley, Australia

Blasting In Environmentally Sensitive Areas A workshop to discuss and review current and future practice

no recourse to another variable. Indeed, a typical approach is to explore the influence of various delay intervals on the blast vibration at a range of monitoring locations until a minimum is found (Yuill and Farnfield, 2001). Fallacy 8: In ground vibrations, strain is proportional to particle velocity. One of the primary objectives of blast vibration measurement and analysis is to predict possible damage to the surrounding rock mass and structures located near the blast. Damage is associated with displacements and strains which are beyond what the rock or structure can support. It is well known that for plane elastic waves, the strain in the material is directly proportional to the particle velocity (see, for example, Dowding, 1985). This fact has been extended by some to assume that it applies universally. Blair and Minchinton (1996) provide a comprehensive counter example for this fallacy in the case of a cylindrical charge. Here, the key features of an analytical analysis is provided using the theory of Heelan (1953). The elastic strains have been calculated. A summary of the results is given in Table 5 where the primed quantities refer to time derivatives of the function, g, which is itself the derivative of the pressure applied to the wall of the cylinder.

Table 5. Functional dependence of the elastic strains and particle velocities for the Heelan model. Quantity radial particle velocity,

ur t u t

Range and time dependence ignoring constants g'


r g' r g r
'

azimuthal particle velocity,

normal radial strain, rr normal tangential strain, normal azimuthal strain, shear strain radial/tangential, r shear strain radial/azimuthal, r shear strain azimuthal/tangential,

+ g r2 g r2 0

g r2

g' r

+ g r2

g r2

Both particle velocities depend on the second derivative of the pressure time history and are inversely proportional to range. Only the normal radial and the radial/azimuthal shear strain

28 October 2001 Hunter Valley, Australia

Blasting In Environmentally Sensitive Areas A workshop to discuss and review current and future practice

contain such a dependence. However, these and all other non-zero strains contain terms with an inverse square dependence on range. It is clear that, in general, the particle velocities are not proportional to strain for the Heelan model. Blair and Minchinton (1996) indicate that some other models for dynamically loaded cylindrical holes are identical to the Heelan model in the far-field and we should expect the same non-compliance with Fallacy 8. Much work has been done to correlate measured strains in structures with particle velocities. Such studies have an engineering purpose but one should recall that strains can be induced by causes other than blast vibration. The reader is referred to the work by Oriard (1999) for a thorough review. Conclusions Blast vibration analysis relies upon good data and sensible models to aid interpretation. Early workers in the field produced useful approaches which continue to be used today. Some of these methods or techniques have strayed from their original intent, or have been applied too widely. They have, in effect, become fallacies. In this paper, I have put forward eight fallacies in blast vibration analysis. They have ranged from time domain measures, the importance of correct frequency estimation, charge weight scaling law limitations, superposition models and the relationship between particle velocity and strain. Counter examples have demonstrated their failings and indicated a better way, or perhaps some aspects that require further research. It is always a little risky to put forward fallacies only to find them proven at a later time. If the debate concerning these provide an improved view, then my primary aim has been satisfied. Acknowledgements Dane Blair has 'seeded' many of the ideas presented here. For that, I thank him. References Anderson, D A (1989). The 8 millisecond "criterion": Have we delayed too long in questioning it ? Proc. 15th Conf. Explosives and Blasting Technique, New Orleans, 5-10 February. Blair, D P (1987). The measurement, modelling and control of ground vibrations due to blasting. Proc. 2nd Intnl. Symp. Rock Fragmentation by Blasting, Keystone, 23-28 August. Blair, D P (1990). Some problems associated with standard charge weight scaling laws. Proc. 3rd Intnl. Symp. Rock Fragmentation by Blasting, Brisbane, 26-31 August. Blair, D P and A Minchinton (1996). On the damage zone surrounding a single blasthole. Proc. 5th Intnl. Symp. Rock Fragmentation by Blasting, Montreal, 25-29 August. Blair, D P (1999). Statistical models for ground vibration and airblast. Int. J. Blasting and Fragmentation, 3, 335:364. Blair, D P (2001). Personal communication. Gould, S J (1981). The Mismeasure of Man, New York, W W Norton. Heelan, P A (1953). Radiation from a cylindrical source of finite length. Geophysics, 18, 685-696.

28 October 2001 Hunter Valley, Australia

Blasting In Environmentally Sensitive Areas A workshop to discuss and review current and future practice

Langhaar, H L (1951). Dimensional analysis and theory of models. New York, John Wiley and Sons. Macquarie Dictionary (1981). McMahons Point, Macquarie Library. Oriard, L L (1999). The effects of vibrations and environmental forces, Cleveland, International Society of Explosives Engineers. Siskind, D A (2000). Vibrations from blasting, Cleveland, International Society of Explosives Engineers. Spathis, A T (1983). Design data for digital Butterworth filters in geophysical data processing. Bull. Aust. Soc. Explor. Geophys. 14, 63-72. Spathis, A T (1990). A seismic source wavelet with applications to the modelling of blast vibrations. Proc. 3rd Intnl. Symp. Rock Fragmentation by Blasting, Brisbane, 26-31 August. Stump, B W and R E Reinke (1987). Experimental seismology: In situ source experiments. Bull. Seismological Soc. Am, 77, No 4, 1295-1311. Yuill, G and R Farnfield (2001). Variations in vibration signals from single hole quarry blasts. Proc. 27 th General and Research Proc. Explosives and Blasting Technique, New Orleans, 28-31 January.

28 October 2001 Hunter Valley, Australia

También podría gustarte