Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

Field Measurement Techniques and Tests for Dynamic Strain

Measurement from Blasting


Ruilin Yang, Kameron Ray, and Kelly Hansen
Orica USA Inc.

September 2011

Executive Summary

A field measurement technique has been developed recently for assessment of dynamic
strain from blasting. The technique is based on the method proposed in the report due to
Yang (2010) to determine dynamic strain from blast vibration. The equipment hardware
was selected and tested in the field. This report documents the field test procedures, data
analysis, and result discussions.

The various strain quantities derived in the report derived from the measurements, such
as maximum tensile, maximum compressive and maximum shear stains, may be used to
describe potential blast damage to the highwalls or rock slopes. Blast damage may be
better controlled by managing the dynamic strains, rather than focusing purely on Peak
Particle Velocity (PPV). Such quantities are more meaningful to a rock mechanics
engineer than PPV alone.

The analysis of the strain quantities also shows that the commonly used assumption by
the blasting community that the PPV is proportional to the dynamic strain may not be
always true because the dynamic strain is related to the displacement gradients which is
affected by the vibration frequency. The dynamic strain measurement may advance our
capability to control blast vibration and damage.

The dynamic strains determined from blast vibration are within theoretical expectations.
The measurement techniques appear applicable at least to surface coal mines and could
also be offered to other open pit mines. At present, the method may not be applicable to
underground applications since surveying the locations and measuring the orientation
angles of the sensor may be too difficult to accurately obtain with sufficient accuracy.

Summary

A field measurement technique has been developed recently for assessment of dynamic
strain from blasting. Equipment hardware was selected and tested in the field. The
essential piece of equipment for this process is the eDAQ lite (eDAQ) data collecting unit

1
which can simultaneously collect up to 20 channels of data with a time synchronization
error less than 0.45µs. The eDAQ records blast vibration from 6 tri-axial accelerometers
placed around an area of 3 m radius. Effort was made to ensure that no more than three
monitors occupied the same plane to achieve an accurate solution for the dynamic strain.
Each accelerometer was placed in a horizontal plane checked by a bubble level. The
direction of the x-axis of the accelerometer from the north was measured by a Brunton
compass and the three-dimensional location of each accelerometer was surveyed with an
accurate differential GPS system.

Blast vibration waveforms were recorded from the accelerometers as input to a program
(written in-house), which processes the data and calculates the dynamic strains three-
dimensionally. Each strain component as a function of time is calculated from the
accelerometer data. The principal strains, their directional numbers, and other selected
strain quantities, such as the maximum octahedral strains, the maximum shear strain,
extensional strain, were also obtained from the program. The trends and value ranges of
the strain quantities seem to be within expectations from theoretical estimates. The planes
on which each maximum strain acts can be identified from the calculations.

Two strain invariants, the maximum principal strain ε1 (maximum compressive), the
maximum octahedral shear strain, and the maximum shear strain contradict the trend of
the peak particle velocity (PPV), but are consistent with the trend of the peak particle
acceleration (PPA). This finding may reveal that the commonly used assumption by the
blasting community that the PPV is proportional to the dynamic strain may not be always
true since the dynamic strain is related to the displacement gradient at each time instance
and may not be related to PPV directly. Instead, the vibration frequency may affect the
strain; higher vibration frequencies will result in higher displacement gradients, therefore,
the high strains. The finding here may advance our understanding of blast damage from
blast vibration.

The technique is recommended as an offering to surface coal mines or other open pit
applications. Using dynamic strains will provide better understanding and prediction of
blast damage caused by blast vibration.

2
Table of Contents
EXECUTIVE SUMMARY...........................................................................................................................1
SUMMARY ............................................................................................................................................1
INTRODUCTION.....................................................................................................................................4
ACCELEROMETER PLACEMENT...............................................................................................................4
SELECTION OF ACCELERATION MOUNTING AREA...................................................................................................4
LEVELING OF THE ACCELEROMETERS .................................................................................................................4
SURVEY OF ACCELEROMETER LOCATIONS............................................................................................................5
MEASURING THE DIRECTION OF THE X‐AXIS OF THE ACCELEROMETERS ......................................................................6
THE EDAQ DATA RECORDER ..................................................................................................................8
TEST BLASTS..........................................................................................................................................9
THE FIRST TEST BLAST ....................................................................................................................................9
THE SECOND TEST BLAST ..............................................................................................................................10
BLAST VIBRATION DATA ANALYSIS ...................................................................................................... 11
DYNAMIC STRAIN CALCULATIONS........................................................................................................ 14
DYNAMIC STRAIN .......................................................................................................................................15
ENGINEERING STRAIN COMPONENTS IN THE MINE COORDINATE SYSTEM ..................................................................15
PRINCIPAL STRAINS .....................................................................................................................................18
OCTAHEDRAL STRAINS .................................................................................................................................22
EXTENSIONAL STRAIN ..................................................................................................................................23
THE MAXIMUM SHEAR STRAIN .......................................................................................................................24
DYNAMIC STRAIN VERSUS THE SCALED DISTANCE................................................................................ 25
MEASUREMENT ERRORS ..................................................................................................................... 27
ERROR FROM THE SURVEY ............................................................................................................................27
ERROR FROM THE ANGLE MEASUREMENT USING THE COMPASS .............................................................................27
ERROR OF BLAST VIBRATION MEASUREMENT .....................................................................................................27
METHOD TO IMPROVE THE ACCURACY .............................................................................................................28
DISCUSSIONS AND CONCLUSIONS........................................................................................................ 28
REFERENCES ........................................................................................................................................ 30
ACKNOWLEDGEMENTS........................................................................................................................ 30
APPENDIX A. DATA SHEET OF TRIMBLE 5800 GPS SYSTEM.................................................................... 31
APPENDIX B. RECORDED PARTICLE ACCELERATION WAVEFORMS ........................................................ 33
APPENDIX C. STANDARD ERROR (STDX)CALCULATION……………………………………………………………………36

3
Introduction

A method to calculate the dynamic strain tensor from blast vibration was reported by
Yang (2010). The dynamic strain was calculated based on the three dimensional gradients
of the displacements among the locations of the blast vibration monitors. In the report, it
was recommended to use a minimum of five monitors (locations) to record blast vibration
to establish simultaneous linear equations. The number of independent equations is larger
than the number of the displacement gradients. Least-square solutions of the
displacement gradients are obtained with improved accuracy using extra independent
equations.

Field tests were conducted recently to examine the proposed method, to test the selected
hardware equipment system, and to establish the field procedure for the measurement.

Accelerometer Placement

The method to calculate the dynamic strain (Yang, 2010) requires the tri-axial
components of the blast vibrations recorded at different locations to be in the same
coordinate systems. Therefore, the orientation of each tri-axial accelerometer must be
measured and then the recorded tri-axial blast vibrations at each location are transferred
to the same coordinate system according to the orientation measured during the set up.

Selection of acceleration mounting area

In order to ensure valid calculations, no more than three monitors should be placed at the
same elevation. Because of this constraint, it is important to find a location near the point
of interest which has a significant elevation difference (such as 3 ft) across a relatively
small distance (such as 15 ft). Steep hills or ditches work well for this.

Leveling of the Accelerometers

Since the ground is soft, each tri-axial accelerometer is attached to a plastic mounting
block. The mounting block is buried in the ground with a firm coupling to the ground by
tapping the soil around the block. The monitoring block is leveled with a sensitive bubble
level (Figure 1). This leveling of the block is the most time consuming task of the field
set-up process. The block locations are set within 3 m (10 ft) radius of the eDAQ location
for the present two tests. The selection of the spacing between the locations of the

4
accelerometers is dependent on the maximum frequency of the blast vibration. This will
be discussed in a later section.

Figure 1. Leveled accelerometer mounting block

Survey of accelerometer locations

Once the blocks are set and leveled, GPS coordinates of the locations of each monitor
must be obtained. The most accurate method is to attach the GPS receiver to a tripod set
up over the center of the accelerometer. A tribrach with an optical plummet is used to
translate GPS measurements to ground level by an offset elevation distance to the center
of the accelerometer. The offset distance was measured using an engineering tape
measure. It is advisable to place the receiver on a tripod as opposed to a monopod as a
more accurate means of measurement. This is due to the fact that the tripod has better
stability and it is easier to align top-dead-center over the accelerometer (Figure 2). The
GPS equipment used for the test is accurate to within 5 mm according to the equipment
supplier (Appendix A).

5
Figure 2. Setting up GPS equipment over the accelerometer mounting points

Measuring the direction of the x-axis of the accelerometers

The accelerometers are attached to the mounting brackets and hooked up to the eDAQ
before measuring their orientation (Figure 3). In order to measure the direction of the x-
axis of an accelerometer, an aluminum block which fits over the accelerometer was
fabricated. This block provides a large surface to which a compass can be lined up
(Figure 4). The compass is not affected by magnetic interference because there is no
magnetic material in the block or the accelerometer. It is important to make sure the
compass is not affected by any magnetic material. For the testing, a Brunton field
compass was used accurate to within 0.5 degrees according to the supplier. The true
direction of the x-axis is obtained by taking into account the magnetic declination. In the
absence of a sufficiently accurate compass, or where magnetic interference is
encountered, the direction of the x-axis should be obtained using at least 2 measurements
with the GPS receiver.

6
Figure 3. Accelerometer mounted to the block

Figure 4. Measuring the direction of the accelerometer with the aid of the aluminum block

7
The eDAQ Data Recorder
The eDAQ data recorder was purchased for this dynamic strain measurement project. The
manufacturer of the eDAQ emerged from the former supplier of the Nicolet Oscilloscope.
The key feature of the equipment is the accurate time synchronization between the
channels. As shown in Figure 5, the maximum error of the time synchronization is less
than 0.45 μs.

Figure 5. Time synchronization versus the number of channels

The eDAQ records time-synchronized data from 6 tri-axial accelerometers. During


testing, sampling rates of 2.5 kHz and 5 kHz were used. The eDAQ must be programmed
prior to use in the field. It is essential to know the sensitivity of each channel of each
accelerometer and to program that information accurately into the eDAQ. This means
that once the eDAQ is programmed, it is important to attach each accelerometer channel
to the correct recording channel.

Programming is performed with the provided TCE (Test Control Center) software. Each
channel is programmed individually. It is possible to write code to include triggers for
each channel, however, due to the vast amount of memory in the eDAQ, it is not
necessary to do so for our purposes. For the time being, continuous recording is
recommended. Figure 6 shows the eDAQ connected to accelerometers in a field test.

8
Figure 6. eDAQ in the case connected to accelerometers

Test Blasts
Two test blasts were conducted at Antelope Coal Mine in the Powder River Basin. The
ground was soft shale with an average compressive strength of 800 psi. The ground sonic
velocity was measured to be 1500 m/s on average. The maximum frequency of the
particle acceleration of the blast vibration was observed to be less than 100 Hz (Yang and
Scovira, 2010). Two blasts with significantly different delay timing were selected for the
tests thereby providing greater opportunity to examine the trends of the results for
comparison with our theory.

The first test blast

The first test blast was a shovel-truck conventional pre-strip blast. The blast hole depth
varied from 44 ft to 52 ft (13 – 16 m). The blast hole diameter was 9.88 inches (251 mm).
The charge weight per hole varied from 1100 lb to 1350 lb (500 – 615 kg). The nearest
blast hole to the monitors was 270 ft (82 m). The accelerometers were arranged with
spacing varying from 4.5 ft (1.37 m) to 9 ft (2.74 m) and some of them were placed in
different elevations (by 0.5 – 1.0 m) ensuring no more than three accelerometers to be in
the same plane. Figure 7 shows a plan view of the first test blast setup (with a zoomed-in
sensor layout plan view).

9
917
21 1M13
8
2019
Amplified view
of the monitors
Six tri-axial accelerometer
9 1 Blast vibration monitors
7 18
M13

21
20 19
484
526
568
610 384
652 426
694 468 242
736 510 284
778 552 326
820 594 368 142
862 636 410 184
904 678 452 226
946 720 494 0
988 762 268 42
536 310 84
1030 804 578 352 126
846 620 394 168
888 662 436 210
930 704 478 252
746 520 294
788 562 336
604 378
646 420
688 462
504
546

Figure 7. A plan view of the first test blast and monitor locations

The second test blast

The second test blast was a cast blast. The blast hole depth varied from 90 ft to 130 ft.
The blast hole diameter was 11.25 inches (286 mm). The charge weight per hole varied
from 2500 lb to 5000 lb (1136 – 2273 kg). The nearest blast hole to the monitors was 575
ft (175 m). The accelerometers were placed with spacing varying from 5.7 ft to 12 ft
(1.74 - 3.66 m). Figure 8 shows a plan view of the second test blast and monitor locations
(with a zoomed-in sensor layout plan view).

ip
Free face 12 ms

Amplified view of the 6


tri-axial accelerometer
monitors
22
20
21

18
2022
21
18
1719

17

19

Figure 8. Plan view of the second test blast and monitor locations

10
Blast Vibration Data Analysis
Seven tri-axial accelerometers were used in each test blast. Out of the seven sensors, six
were connected to the eDAQ with the time synchronization for dynamic strain
measurements. The remaining one was connected to an Instantel monitor to record blast
vibration near the six sensors and was used as a reference check of the vibration
measurements. The recorded particle acceleration waveforms are shown in Appendix B.
Figure 9 shows the peak particle acceleration (PPA) versus the charge weight scaled
distance. Although the designs of the two blasts were very different in terms of delay
time and charge weight per delay, the PPA from the two blasts decreases with the
increase in the scaled distance as expected.

Figure 10 shows that the peak particle velocity (PPV) from the two blasts, does not
decrease with an increase in scaled distance. This is somewhat controversial as it defies
the commonly observed trend. However, by examining the dominant frequency of the
particle acceleration waveforms from the two test blasts, the abnormal trends in Figure 10
can be easily explained.

Figure 11 shows a typical power spectrum of the particle acceleration waves (the average
of the three components) from the first test blast. As can be seen, the dominant frequency
is 23.6 Hz, whereas Figure 12 shows the dominant frequency from the second test blast to
be 3.4 Hz. Figure 13 compares the dominant frequencies from the seven monitors in the
two blasts. In the figure, the dominant vibration frequencies from the first blast are over
six times higher than those from the second blast. The first blast was fired with 42 ms
delay between blast holes (refer to Figure 7) and shifted the dominant frequency of the
blast vibration to 23.5 Hz, whereas, the 12 ms hole-to-hole delay in the second blast
(refer to Figure 8) did not shift the frequency. For blast vibrations with the same PPA, the
vibration with higher dominant frequency yields lower PPV. A rough estimate (Yang,
2012) shows that the PPV is inversely proportional to the frequency of the particle
acceleration wave. In other words, for the same PPA, test blast#1 could yield only one
sixth the PPV compared to the blast vibration from test blast #2. The large frequency
difference between the first and the second test blasts explains the reversal trend in Figure
10.

11
PPA vs. MSD for the two test blasts
3.6

3.4
1st test
3.2 2nd test
3.0

2.8

2.6
PPA (g)
2.4

2.2

2.0

1.8

1.6

1.4

1.2

1.0
7.6 7.8 8.0 8.2 8.4 8.6 8.8 9.0 9.2 9.4 9.6 9.8 10.0 10.2 10.4 10.6

MSD (ft/lb0.5)
Figure 9. Peak particle acceleration (PPA) versus the charge weight scaled distance

PPV from two different blasts vs. scaled distance


12

1st test
11 2nd test

10

9
PPV (in/s)

5
7.6 7.8 8.0 8.2 8.4 8.6 8.8 9.0 9.2 9.4 9.6 9.8 10.0 10.2 10.4 10.6
MSD (ft/lb0.5)

Figure 10. Peak particle velocity (PPV) from the two blasts versus the charge weight scaled distance

12
Figure 11. A typical power spectrum of the particle acceleration wave (the average of the three components)
from the first test blast

Figure 12. A typical power spectrum of the particle acceleration wave (the average of the three components)
from the second test blast

13
Acceleration dominant frequency of the two test blasts
26

24 1st test blast


2nd test blast
22

Acceleration wave dominant frequency (Hz)


20

18

16

14

12

10

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Figure 13. Comparison between the dominant frequencies from the seven monitors in the two blasts

Dynamic Strain Calculations


The program is based on the single valued decomposition (SVD) method to solve linear
simultaneous equations with the number of equations larger than the number of variables
(over-determined system). The program is written with Matlab and the C++
programming languages. The SVD method produces a solution that is the best
approximation in the least squares sense. This is equivalent to the following:

that minimizes χ 2 = A ⋅ a − b
2
Find a (1)

where, A and b are the known coefficient matrix and the displacement vector,
respectively. a is the unknown vector. A is a M X N matrix whose number of rows M is
greater than or equal to its number of columns N, and can be written as the product of an
M X N column-orthogonal matrix U, an N x N diagonal matrix W with positive or zero
elements (the singular values), and the transpose of the N x N orthogonal matrix V, that is,

A = U W VT (2)

Then the solution of the least-squares problem (1) can be written as (Press et al, 2005):

⎛ (i ) ⋅ b ⎞
N −1 U
a = ∑ ⎜⎜ ⎟ ⋅ V(i ) (3)
i =0 ⎝ wi ⎟⎠

The above equation shows that the fitted parameters a (vector) are a linear combination
of the columns of V, with coefficients obtained by forming dot products of the columns
of U with the displacement vector b.

14
Dynamic Strain

The blast vibration signals from the six time-synchronized tri-axial accelerometers are
input to the above mentioned program. The program integrates the particle acceleration
waveforms to obtain displacement waveforms and calculates the six strain components as
functions of time, corresponding to the mine coordinates: east (x), north (y) and elevation
(z). The program can also output the three principal strains and their corresponding
directional cosines as functions of time. From these it is then easy to obtain the maximum
shear strain and its acting plane.

Engineering strain components in the mine coordinate system

Yang (2010) proposed a method for estimating the dynamic infinitesimal engineering
strain by measuring the time-synchronized displacement gradients from blast vibration
monitoring. The engineering strain is:

⎛ ∂u ∂u ∂v ∂u ∂w⎞
⎜ + + ⎟
⎜ ∂x ∂y ∂x ∂z ∂x ⎟
(Eij ) = ⎜⎜ ∂v + ∂u ∂v ∂v ∂w⎟
+ ⎟ (4)
⎜ ∂x ∂y ∂y ∂z ∂y ⎟
⎜ ∂w + ∂u ∂w ∂v
+
∂w ⎟
⎜ ∂x ∂z ∂y ∂z ∂z ⎟⎠

where, u, v, and w are tri-axial displacement components at the point of interest (x, y, z).
u, v, and w are functions of x, y, z and time t, i.e.

u = u (x, y, z, t)
v = v (x, y, z, t) (5)
w =w(x, y, z, t)

Consequently, the strain tensor Eij is a function of x, y, z, and t. Figures 14 and 15 show
the dynamic strain calculated from the measured blast vibration in the first and second
test blasts as functions of time. It should be noted that the strain components are
dependent on the selection of the coordinate system. With this in mind, from Figures 14
and 15, it can be seen that the normal x-strains from both tests are smaller than the
normal y- and z- This could be due to the fact that the x-axis is transverse to the blast in
both tests. The normal z-strain is large in both tests. This could be due to the effect of the
free surface causing large displacement gradients along the z-axis.

15
Figure 14. Dynamic strain calculated from the blast vibration in the first test blast

Figure 15. Dynamic strain calculated from the blast vibration in the second test blast

16
Figures 16 and 17 show that the corresponding standard errors for each strain component
are of the same order of magnitude as the corresponding strain components (Figures 14
and 15). The standard error (σ) is a measure of the scatter of each strain component
calculated from the different sets of all the combinations of simultaneous equations
forming an over determined system. We would expect the solution of the strain
components obtained from the least-square fit to be similar to the mean or median of all
individual solutions from the different combinations of the equations. The standard error
σ ⎞
of the mean or median would be much smaller ⎛⎜ ⎟ than the standard errors shown in
⎝ n⎠
Figures 16 and 17.

Figure 16. Standard error of the strain components in Figure 14, Test #1

Figure 17. Standard error of the strain components in Figure 15, Test #2

17
Principal Strains

The infinitesimal strain tensor is defined as:

⎡ ∂u 1 ⎛ ∂u ∂v ⎞ 1 ⎛ ∂u ∂w ⎞⎤
⎢ ⎜ + ⎟ ⎜ + ⎟⎥
∂x 2 ⎜⎝ ∂y ∂x ⎟⎠ 2 ⎝ ∂z ∂x ⎠⎥
⎡ε 11 ε12 ε 13 ⎤ ⎢
⎢ 1 ⎛ ∂u ∂v ⎞ ∂v 1 ⎛ ∂v ∂w ⎞ ⎥ (6)
εˆ = ⎢ε 21 ε 22 ⎥
ε 23 = ⎢ ⎜⎜ + ⎟⎟ ⎜ + ⎟
⎢ ⎥ 2 ∂y ∂x ⎠ ∂y 2 ⎜⎝ ∂z ∂y ⎟⎠ ⎥
⎢⎣ε 31 ε 32 ε 33 ⎥⎦ ⎢ ⎝ ⎥
⎢ 1 ⎛ ∂u + ∂w ⎞ 1 ⎛ ∂v ∂w ⎞ ∂w ⎥
⎢ 2 ⎜⎝ ∂z ∂x ⎟⎠ ⎜ + ⎟
⎣ 2 ⎜⎝ ∂z ∂y ⎟⎠ ∂z ⎥

It can be shown that a coordinate system (n1, n2, n3) exists in which the strain tensor (6)
above can be transformed into:

⎡ε1 0 0 ⎤
εˆ = ⎢ 0 ε 2 0 ⎥ (7)
⎢ ⎥
⎢⎣ 0 0 ε 3 ⎥⎦

The components of the strain tensor in the (n1, n2, n3) coordinate system are called the
principal strains and the directions ni are called the directions (direction cosines) of the
principal strain. Since there are no shear strain components in this coordinate system, the
principal strains represent the maximum and minimum stretches of an elemental volume.
By convention, the tensile strain is positive and the compressive strain is negative.

Figures 18, 19, and 20 show the principal strains ε1, ε2, and ε3 from test blast#1 and their
corresponding directional cosines as a function of time. The compressive strain is ε1
using the rock mechanics convention. As noted in the figures, there are three direction
cosines for each principal strain. The orientation of the plane on which the peak principal
strain acts can be determined from the direction cosines. Such information could be
useful for examining the effect of the dynamic strain on a joint plane or faults.

The maximum compressive strain is 4000 με (ε x 10-6). This means that a 4 mm length
reduction for every meter dimension of the rock. The dynamic Young’s modulus was
estimated to be 0.4 x 106 psi (2.8 GPa) (Yang, 2009). Therefore, the maximum
compressive stress from the first test is estimated to be 1600 psi (11 MPa). The
measurement location was 82 m from the blast and no cracks were observed. Therefore,
the rock dynamic uni-axial compressive strength is estimated to be greater than 1600 psi
(11 MPa). Such an estimate seems to be within the expectation according to the site
geology (Pratt and Pryor, 2003).

Figures 21, 22, and 23 shows the principal strains ε1, ε2, and ε3 from test blast#2 and their
corresponding direction cosines as a function of time. Again, the orientation of the plane
on which the peak principal strain acts can be determined from the direction cosines.

18
Such information could be useful for examining the effect of the dynamic strain on a joint
plane or faults.

Figure 18. Principal strain#1 (compressive) and its direction cosines from Test #1

Figure 19. Principal strain#2 and its direction cosines from Test #1

19
Figure 20. Principal strain#3 (Tensile) and its direction cosines from Test #1

Figure 21. Principal strain#1 (compressive) and its direction cosines from Test #2

20
Figure 22. Principal strain#2 and its direction cosines from Test #2

Figure 23. Principal strain#3 (tensile) and its direction cosines from Test #2

21
Octahedral Strains

It is instructive to compare the strains between the two test blasts. Octahedral normal and
shear strains may be used for the comparison since they are scalar terms derived from all
strains and each correlates with a clear physical deformation. Normal octahedral strain
relates to the volumetric deformation, whereas octahedral shear strain represents the total
shear deformation in three orthogonal directions. Octahedral strains are strain invariants
and are therefore independent of the coordinate systems that are chosen.

The engineering shear strain on an octahedral plane is known as the octahedral shear
strain and is given by

(8)

where, are the principal strains.

The normal strain on an octahedral plane is given by

(9)

Figures 24 and 25 show the octahedral normal and shear strains from Test#1 and #2,
respectively.

Figure 24. Octahedral normal and shear strains from Test#1

22
Figure 25. Octahedral normal and shear strains from Test#2

Extensional strain

The extensional strain concept was proposed by Stacy (1981) to model crack initiation
and propagation for static rock failures and was used by Yang et al (1996) to model
damage from blasting. The extensional strain is defined as the sum of all the tensile
strains at a point

εe= ε3+ ε2 (if ε2>0) (10)

For the present tests, the intermediate principal strain ε2 is small; therefore, the
extensional strain is close to the tensile strain ε3. In the general, the extensional strain
may be used to quantify the damage. Figures 26 and 27 show the extensional strain from
the first test and second test blasts, respectively.

23
Figure 26. Extensional strain from the first test

Figure 27. Extensional strain from the second test

The maximum shear strain

The maximum shear strain can be calculated as:

ε1 − ε 3
τ max = (11)
2
The normal of a maximum shear plane subtends equal angles (45o) with the principal
directions of ε 1 and ε 3 .

Figures 28 and 29 show the maximum shear strains from test blast #1 and #2.

24
Deleted: Text
Figure 28. Maximum shear strain from Test #1

Figure 29. The maximum shear strain from Test #2

Dynamic strain versus the scaled distance

In order to understand the trends of the dynamic strain against the charge weight scaled
distances, the maximum strain values are plotted against the charge weight scaled
distances. Figure 30 shows the maximum compressive strain, the maximum octahedral
shear strain, and the maximum shear strain from the two test blasts against the scaled
distance. It can be seen clearly that the strains decrease with the increase in the minimum
scaled distance. This trend is expected and is consistent with the trend of the peak particle
acceleration shown in Figure 9 above. However, the trends for these strains are not
consistent with that for the peak particle velocity in Figure 10. The inconsistency implies
that high PPV blast vibration may not produce the high compressive or octahedral shear

25
strains. This finding may reveal that the commonly used assumption by the blasting
community that the PPV is proportional to the dynamic strain may not always be true. In
fact the dynamic strain may be related to the frequency as well. In the present tests, the
blast vibration from the first blast has a higher frequency than that from the second blast.
The higher the frequency of the blast vibration, the higher the probability of increased
gradient of displacement, and therefore, the higher the strains. This finding is significant
for understanding damage due to blast vibration.

Scatterplot (dynamic_strain_msd.sta 10v*6c)


4500

Maximum compressive strain (με)


4000 Maximum Octahedral shear strain (με)
maximum shear strain (με)

3500
με

3000

2500

2000

1500
7.6 7.8 8.0 8.2 8.4 8.6 8.8 9.0 9.2 9.4 9.6 9.8

MSD (ft/kg0.5)

Figure 30. Maximum compressive, octahedral shear, shear strains versus the minimum scaled distance

2800

2600

2400

2200

2000 Maximum Tensile strain


Maximum Octahedral Normal strain
1800
extentional strain
με

1600

1400

1200

1000

800

600

400
7.6 7.8 8.0 8.2 8.4 8.6 8.8 9.0 9.2 9.4 9.6 9.8

MSD (ft/kg0.5)
Figure 31. Maximum tensile, octahedral normal, extensional strains versus the minimum scaled distance

26
Figure 31 shows the maximum tensile, octahedral normal, extensional strains versus the
minimum scaled distance. These values do not change with MSD which does not agree
with the trends of the PPA and the PPV in Figures 9 and 10 respectively, where PPA
decreases with MSD and PPV increases with PPV. Again, this shows that PPV alone
may not be enough to represent the effect of blast vibration on rock damage or
deformation. The blast damage may be better controlled by managing the dynamic strains,
rather than focusing purely on PPV.

Measurement errors
Error from the survey

Errors from the survey of the sensor locations exist due to the operation, the accuracy
limitation of the survey equipment. A Trimble 5800 GPS System was used to survey the
sensor locations. As shown in Appendix A, the accuracy of the GPS system is ± 5 mm.
These potential errors are controlled firstly by maximizing the spacing of the sensors with
respect to the upper bound of the vibration frequency (Yang, 2011) as well as by
increasing the number of vibration monitoring points to obtain the number of equations
greater than the number of variables.

Error from the angle measurement using the compass

The Brunton Pocket Transit is used to measure the angle of the sensor mounting. This
combines a surveyor’s compass, prismatic compass, inclinometer, hand level and plumb
into a single instrument. The Brunton Pocket Transit is used to measure azimuth
(compass bearing), vertical angles, inclination of objects, percent grade, slopes, height of
objects and for leveling. The instrument is accurate to ± 0.5o for bearing measurements
and ± 1o for inclination (30 minute readable).

Error of blast vibration measurement

The blast measurement at each monitoring point has an error. This could be due to the
sensor mounting variations, such as coupling between the mounting block and the ground.
10% error could be expected even with a careful setup in the field.

27
Method to improve the accuracy

For the current tests, fifteen equations were used for 9 variables. This means that the
least-square solution is obtained from 5005 sets of solutions because:

⎛15 ⎞ 15!
⎜⎜ 9 ⎟⎟ = 9!6! = 5005 (12)
⎝ ⎠

According to statistics, the error may be reduced by a factor of:

1 1
= (13)
5005 − 1 70.7

Apparently, the more monitoring points, the better the accuracy of the strain
measurement. If eight monitors are used, the error could be reduced potentially by a
factor of over 500.

Discussions and Conclusions

The test results are consistent with some theoretical expectations, i.e. the maximum
compressive and octahedral strains correlate with the charge weight scaled distances. The
tests also show that the commonly used assumption by the blasting community that PPV
is proportional to the dynamic strain may not be always true. The blast damage may be
better controlled by managing the dynamic strains, rather than focusing purely on PPV.

The derived strain quantities from the dynamic strain measurement may be used to
describe potential blast damage to the high walls or rock slopes. Such quantities are more
meaningful to a rock mechanics engineer than PPV only. The dynamic strain
measurement is useful to further our understanding of vibration control for blast damage.

The present two tests were conducted in a soft ground coal mine. The mounting blocks
for the accelerometers were buried in the ground by placing them horizontally with a
bubble level indicator. This technique seems to work for these test conditions. However,
if the mine site is hard rock, the accelerometers have to be plastic bonded (epoxy) to the
rock surface, which can be oriented in any direction. Using a compass to measure the
orientation angles – dip angle and the striking angle is neither easy nor accurate in the
field. For the present test, with level mounting blocks, only the direction (angle) of the x-
axis of the accelerometers needs to be measured. For hard rock surface applications, three
angles need to be measured – surface dip, surface striking, and the direction of the x-axis
of the accelerometers. Although it is difficult, a method for measuring and adjusting the
readings to compensate can be derived.

28
The measurement techniques appear applicable at least to surface coal mines. The
technique could be offered to other open pit mines. At present, the method may not be
applicable to underground applications since the surveying locations and measuring the
orientation angles of the sensor mounting may be too difficult to accurately obtain.

It is recommended that more sensors (monitoring locations) are used to improve the
measurement accuracy. Even with just two more monitoring points, it is estimated that
1 1
the error could be reduced by a factor of 7.66 ( ≈ , refer to the above
⎛ 21⎞ 542
⎜⎜ ⎟⎟ − 1
⎝9⎠
section: Method to improve the accuracy).

The orientation of the plane on which a peak strain quantity (such as, principal, the
maximum shear strain, the maximum octahedral strains, etc.) acts can be determined
from the direction cosines. Such information could be useful for examining the effect of
the dynamic strain on a joint plane or faults.

The setup of the measurement for dynamic strain requires much more effort and diligence
than normal blast vibration monitoring. At least two well trained people are required in
the field for careful setup of the equipment and monitoring in order to obtain accurate
results.

29
References
Pratt, R. W. and Pryor, P. R., 2003, Cordero-Rojo Middle Pit Geotechnical Study,
prepared for Kennecott Energy.

Press, W. H., Teukolsky, S. A., Vetterling, W. T., and Flannery, B. P., 2005, Numerical
Recipes in C++, The Art of Scientific Computing, 2nd Edition, Cambridge University
press.

Stacy, T. R., 1981, A simple extension strain criterion for fracture of brittle rock, Int. J.
Rock mech. Min. Sci. & Geomech. Abstr. 18, pp.469-474.

Yang, R., 2010, Method to calculate 3-D dynamic strain tensor from blast vibration
monitoring. Orica Internal Technical Report.

Yang, Ruilin, 2011, Analysis of the error due to time synchronization between monitors
for dynamic strain measurement, Orica Internal report, August, 2011.

Yang, Ruilin and Scovira, D. Scott, 2010, A model for near and far field blast vibration
based on multiple seed waveforms and transfer functions, Blasting and Fragmentation,
Vol. 4, No. 2, 2010, pp. 91-116 Trimble 5800 GPS System.

Yang, Ruilin, 2009, Overburden Signature Hole Vibration Monitoring and Modeling at
Antelope Coal Mine, Presentation to Antelope Coal Mine of Cloud Peak Energy, 2009
April.

Yang, Ruilin, Bawden, W.F., and Katsbanis, P.D., 1996, A New Constitutive Model of
Blast Damage, International Journal or Rock Mechanics and Mining Sciences &
Geomechanics Abstracts, Volume 33, Number 3, pp.245-254.

Yang, Ruilin, 2012, Relationship between peak particle acceleration, velocity and
displacement of blast vibration, The 38th Conf. Explosives and Blasting Technique,
Nashville, TN, USA.

Acknowledgements: Thanks to Antelope Mine of the Cloud Peak Energy for their
permission to use their site to conduct the tests. Dr. Dale Preece, Dr. Alan Minchinton,
and Dr. Mick Lownds reviewed the report and made useful comments.

30
Appendix A. Data sheet of Trimble 5800 GPS system

31
32
Appendix B. Recorded particle acceleration waveforms
Test#1, Monitor#9 and 17

Test#1, Monitor#18 and 19

33
Test#1, Monitor#20 and 21

34
Test#2, Monitor# 17 and 18

Test#2, Monitor# 19 and 20

35
Test#2, Monitor# 21 and 22

Appendix C. Standard Error (stdx)Calculation

Aa=b

where, A and b are the known coefficient matrix and the displacement vector,
respectively. a is the unknown vector. A is a M X N matrix whose number of rows M is
greater than or equal to its number of columns N.

1
S= {( A' ⋅ A) −1 ⋅ [b ' ⋅ (V −1 − V −1 ⋅ A ⋅ ( A' ⋅ A) −1 ⋅ A' ⋅ V −1 ) ⋅ b]}
M −N

stdxi = Sii

Where,

V is taken as an identity matrix of size N.

36

You might also like