You are on page 1of 16

25

Seismic Investigation for


Rock Engineering
PETER W. McDOWELL
University of Portsmouth, UK

25.1 INTRODUCTION 619


25.2 BASIC PRINCIPLES OF SEISMIC SURVEYING 621
25.2.1 Seismic Velocity 621
25.2.2 Seismic Surveying 622
25.3 EVALUATION OF FRACTURE STATE 623
25.3.1 Fracture Porosity 623
25.3.2 Fracture Index 624
25.3.3 Attenuation 624
25.4 OTHER FACTORS AFFECTING WAVE VELOCITY 625
25.4.1 Moisture Content 625
25.4.2 In Situ Stress 626
25.4.3 Frequency 626
25.5 ASSESSMENT OF ELASTIC MODULI 627
25.5.1 Dynamic Elastic Moduli 627
25.5.2 Static Elastic Moduli 627
25.6 CASE STUDY 628
25.6.1 Site Conditions and Procedures 628
25.6.2 Assessment of Results 631
25.7 CONCLUSIONS 632
25.8 REFERENCES 633

25.1 INTRODUCTION
For most people, seismic surveys conjure up the picture of exploration for petroleum rather than
the investigation of the ground for rock engineering. This is not surprising, as a vast amount of effort
and money has been expended on seismic surveys in the search for oil and gas during the last three
decades; present estimates of expenditure being in excess of £3000 million per annum. A high degree
of sophistication has been achieved for all stages of the investigation, from initial data acquisition to
the well-known impressive graphical presentation of results; especially for seismic reflection
profiling. The expenditure is, of course, justified by the provision of a general, and sometimes
detailed, picture of the variations in geological lithology and structure within an area of interest.
This allows for a corresponding decrease in the requirements for boreholes, which are far more
expensive. Except in water-covered areas, seismic reflection methods are less commonly used in
ground investigations for engineering purposes than boreholes because the sites are usually more
accessible and the depth of interest is considerably less than for oil and gas. Seismic refraction
methods have, however, been used since the 1930s for 'depth to bedrock' determinations and the
location and delineation of faults, buried channels, igneous intrusions and other potential geological
619
620 Geophysics
hazards to engineering works. Also, in recent years, there has been a considerable improvement in
field instrumentation and development in computer-assisted data processing, largely as a spin-off
from oil and gas exploration technology, and engineering geophysicists have tailored and refined the
seismic investigation procedures for detailed rock mass evaluation. Seismic refraction is, at present,
much more widely used than seismic reflection for on-shore site investigation, but high resolution
seismic reflection profiling techniques have now been developed for use on land. A wide variety of
borehole seismic techniques have also been adapted for shallow depth investigations.
Seismic measurements can be more informative than any other geophysical method used for rock
mass characterization and testing. This is because the geophysical parameters measured, such as the
velocity of compressional and shear-wave events, and occasionally also the attenuation coefficients
of these events, are all directly related to strength and elasticity and indirectly related to bulk density
and porosity. Rock mass 'fracture state' can be evaluated by comparing the velocity values measured
on site to corresponding measurements made upon rock samples in the laboratory. The grade of
weathering of rock types in which weathering profiles characteristically develop, such as granite and
chalk, can be related to typical ranges in value of compressional wave velocity or other seismic
parameters. In recent years, it has become common practice to measure shear-wave velocities as well
as compressional-wave velocities, particularly when dynamic modulus values are required for the
ground at low rates of strain and consideration has to be given to the possibility of resonance. An
example of this application is provided for the site of a power station in Mauritania, West Africa,
where design criteria were required for the foundation blocks of the electricity generators. In this
case, the velocity values were also used to assess ground deformation under static loading.
There are many other cases where seismic data can assist the evaluation of the response of the
ground to dynamic loading. Obvious examples are sites for engineering structures which are
subjected to wave and wind action; such as dams, off-shore platforms and wind generators. In areas
where substantial ground vibrations can be produced by earthquakes, the investigation of fault
zones, the measurement of dynamic values of Young's modulus, rigidity modulus and Poisson's
ratio, and the long-term monitoring of ground vibration all necessitate the use of seismic methods.
Seismic methods are very versatile and the field procedures can be designed to fully evaluate a site,
which is not usually economically or practically viable by direct investigations alone. The seismic
refraction spreads at the ground surface can be arranged in any azimuth to investigate lateral
anisotropy in ground conditions and rock mass properties. As the ray paths are essentially parallel
to the ground surface, i.e. subhorizontal, narrow near-vertical features, such as fracture zones, can
usually be located and defined. This process can be carried to any depth by 'shooting' between pairs
of boreholes and using velocity tomograms produced by computer processing of the data. Where
boreholes are available, seismic measurements can also be made between any point on the ground
surface and any point within the borehole. Also wire-line logging equipment, such as the continuous
velocity logger, can be used to measure the variation of the velocity and attenuation of com-
pressional and shear waves of small volumes of rock around the borehole with depth. These in-hole
measurements are usually made of ultrasonic frequencies and the ray paths are parallel to the
borehole. Similar high frequency measurements could be made, if required, at the ground surface or
between boreholes, adding yet another dimension to the seismic investigation procedures. All of
these variables need to be taken into account in the design of a seismic survey.
The energy sources available for seismic measurements can also be varied to suit a specific
application. Directional reversible impact sources are utilized to assist in the identification of shear-
wave events and to provide accurate values of shear-wave velocities. Gas guns, air guns and sparker-
sources are often utilized for borehole measurements where a rapid repeatable source is required.
Continuous sources, operating over a range of frequencies, can be used for shallow depth reflection
surveys and for ground resonance studies.
In principle, these seismic investigation procedures can also be utilized for rock mass monitoring.
An obvious example is the repeat of measurements before and after grouting with cement to assess
the effectiveness of the process. Similarly, damage or stress relief resulting from excavation can be
assessed, although in this case a decrease in seismic wave velocity would be expected. As velocity
varies with stress, continuous measurement can act as an independent assessment of stress level, or
readjustment of the rock mass to excess stress, such as rock bursts in deep mines.
This introduction to seismic investigations for rock engineering indicates the sophistication and
complexity of many of the procedures as well as the many applications. The following sections will
highlight the main principles that should be appreciated by any person involved in any way in these
types of investigations, with emphasis being placed upon the effect of frequency, stress level and
fracture state upon seismic properties; factors not adequately covered in existing texts. In practical
Seismic Investigation for Rock Engineering 621

terms, one needs to consider the type of equipment used, the number and spacing of boreholes and
the integration of the seismic investigations within the framework of the complete ground investiga-
tion. The information provided from seismic investigations in Mauritania should illustrate some of
these principles, applications and practical difficulties.

25.2 BASIC PRINCIPLES OF SEISMIC SURVEYING


25.2.1 Seismic Velocity
Seismic methods depend primarily upon the measurement of the velocity of propagation of sound
waves through rocks and associated materials. When the energy source is at or near the ground
surface, both surface and body waves will be produced which travel at different velocities. For most
applications the following types of body waves are of particular interest.
(i) Compressional or longitudinal waves (P-waves). These have a particle motion in the direction
of wave propagation and travel faster than all other wave forms.
(ii) Shear or transverse waves (S-waves). Ground motion is transverse to the direction of wave
motion, i.e. polarized in one plane. In practice, the S-wave motion is normally polarized into
components which are parallel and perpendicular to the surface of the ground, i.e. SH- and Sv-waves.
The energy source and recording system has to be carefully considered to ensure correct
identification and accurate recording of the travel times of these events. This is particularly
important for S-wave measurements as these are late arrivals on a seismic trace and may be confused
with refracted or reflected P-wave events or even surface waves.
The velocity of seismic waves depends upon a large number of factors, the most important being
the elastic moduli. For homogeneous isotropic elastic media, the P- and S-wave velocities can be
related to dynamic elastic moduli and bulk density using the following formulae

__ Eo (1 - flo) -1
V - ms
P p (1 - 2flo)(1 + flo)

and
Eo 1 -1
V - ms
S p 2(1 + flo)

where ED = dynamic Young's modulus of elasticity (N m -2); liD = dynamic Poisson's ratio and
p = bulk density (kg m - 3).
Although most soil and rock masses are not homogeneous and isotropic, they can usually be
considered to behave elastically where the stress levels are low, and the wavelength is significantly
greater than the dimensions of pore spaces, voids and fractures. It is clearly, therefore, very
important when deriving dynamic elastic parameters from seismic velocity values to consider the
relationship of wavelength to the size, spacing, infill, etc., of fractures. Although these formulae
suggest that velocity will decrease with increasing density, the opposite is normally the case because

Table 1 Some Typical Bulk Density and Velocity Values

Medium Bulk density Vp Vs


(kg m -3) (m S-l) (m S-I)

Air 1 330
Water 1000 1450
Sands and clays 1500 300-1900 100-500
Glacial till 1800-2000 1500-2700 600-1300
Chalk 1800-2200 1700-3000 600-1500
Strong limestones 2500-2700 3000-6500 1500-3500
and dolomites
Fractured and/or 2000-2500 1000-3000 500-1500
weathered granite
Fresh granite 2600-2800 3000-6000 1500-3000
Slate 2700 5000-7000 2500-3800
622 Geophysics

an increase in density is usually related to an increase in compaction and/or cementation, resulting


in an even more significant increase in elastic moduli.
Corresponding bulk density and velocity values for a few selected rock types and other materials
of interest are provided in Table 1. There is an overlap in the ranges of velocity values for typical
groups of rocks because of the wide variety of rock types in each group. Values for a specific rock
type can be obtained from published literature but these should be used with caution as they depend
upon the specific site conditions. For example, most of the velocity values obtained from oil
exploration and quoted in many of the more general geophysical texts are relatively high because
they relate to fresh saturated rocks beneath a large depth of cover. Table 1 indicates that the velocity
ranges can be even greater within weathered zones, or when the rock mass is fractured. These factors,
which are of paramount importance in rock engineering, will be considered more fully in subsequent
sections.

25.2.2 Seismic Surveying


The basic principles of the seismic refraction and reflection surveying methods are fully covered in
textbooks of applied geophysics. However, a text written primarily for civil engineers, such as that by
Griffiths and King [1] would probably provide a more appropriate introduction to these methods.
Particular attention should be paid to the seismic refraction method, which can provide P- and S-
wave velocity values as well as the variation in thickness of seismic layers, and to borehole seismic
methods.
A particularly important procedure for site investigation and rock mass evaluation is the
reciprocal method of continuous seismic refraction surveying. This was described by Hawkins [2]
and is similar to the plus-minus method of interpreting seismic sections [3]. The technique involves
the setting out of a line of 12 or 24 geophones, each connected through a multicore cable to a
seismograph. Energy is generated by an appropriate source, which can be a sledgehammer blow if a
signal-enhancement seismograph is available, but may need to be an explosive charge when the
spread lengths are large, the surface layer is loose or broken, and the background noise level is high.
The length of the seismic spread, i.e. geophone array, should be at least five times the maximum
required depth of investigation. Several positions of the energy source are required for each spread,
including those offset from each end of the spread to provide the reciprocal time for time-depth
analysis. Figure 1 shows the layout described and the procedure for calculating the time depth at
each geophone station. This is essentially the same as the intercept-time method for finding the
depth to a seismic refractor at the 'end-shot' and 'middle-shot' locations and enables seismic profiles
to be constructed. The procedure also enables a full analysis of the variation in seismic velocity
laterally within each seismic layer. An example is provided (Figure 2) from Markham, South Wales,
where a sewer pipe had to be constructed at a depth of 5 m; the P-wave velocity value of 3250 m s - 1
represented sound sandstones and mudstones and the intermediate velocity layer of
1300-2400 ms - 1 represented glacial till [4].
The seismic refraction method has some important limitations which need to be fully understood
if the profiles are to be translated in terms of rock mass properties. Firstly, the seismic velocity values
are for essentially horizontal ray paths when the ground surface is level, and represent only the upper

----·1
End shot End shot

I. Seismic spread
* Geaphone station Middle shot
Long shot
Reciprocal geophone
* *x

Time depth = 1/2 (TTAeCG + TTXYZG - TTABYX )=1/2 (TTcG + TTZG - TTczl (wherB TT is travBI timB in msl

TTAByX ' TT XYBA ' reciprocal time; 0G= f (Time depth; V, V2 V3 l

Figure 1 The 'reciprocal method' of seismic refraction surveying


Seismic Investigation for Rock Engineering 623

o 1.5 3m
Horizontal scale L.-L.-J
S N
SPI SP2 SP3

OJ ~I
1 I 2
I
3
I
4
r
5!
I
6
I
7
I
8
r
9
1
10
I
II
1
L
400
1.5 068~~7'~ - - - ------o.99 ~.;~--------------0.95-;~~
3m 270-' .___ __.- - - ._ _
3250 ._. 3250 ~

600 Compressional wave velocity m S-I .~._.,;{~~


_.-....../ Main refractor profile

- - - - - Interpolated refractor profile

Figure 2 Seismic profile - Markham, South Wales

part of each seismic layer. Thin intermediate layers may remain unidentified when only first arrival
P-wave events are recorded; the hidden layer or blind zone problem. The maximum possible
thickness of such layers can, however, be calculated if the problem is identified and velocity values
for the intermediate layer are provided from the study of second events, or from adjacent spreads
[5]. A thin 'high speed' layer can mask the underlying material and also produce errors in depth
determinations to deeper refractors, but this problem can usually be identified by the severe
attenuation of seismic wave amplitude within such layers [6]. The problem of a continuous velocity
variation with depth, as might occur within the weathering profile of igneous rocks, is usually best
resolved by considering this as a large number of thin discrete velocity layers.
By far the most serious limitation of the seismic refraction method is velocity inversion and the
associated problem oflow velocity layers. An intermediate layer oflow velocity, such as a shale layer
in a limestone sequence, will not be recognized and depth calculations to underlying refractors will
be erroneously large. It is possible, however, to obtain useful information from seismic surveys if the
problem is recognized and the velocity of the low-velocity layer is obtained independently, e.g. from
up-hole shooting from a borehole to the seismic spread. An example of the mapping of a buried
channel by seismic refraction in an urban area with a velocity inversion problem is provided by
McDowell [6].
A wider consideration of the advantages and limitations of the seismic refraction method is
included in the Geological Society Working Party Report on Engineering Geophysics [7].

25.3 EVALUATION OF FRACTURE STATE


25.3.1 Fracture Porosity
The effect of fractures upon the velocity of propagation of seismic waves is best illustrated by
considering a particular example. If a compressional seismic wave travels directly through 20 m of
fresh limestone without joints in 5 ms, the velocity is 4000 m s -1 • The corresponding velocity when
10 water-filled fractures of width 0.05 m have to be traversed can be calculated using the following
time-average formula
L nw (L - nw)
------ + -------
V p (rock mass) V p (fracture filler) V p (rock material)

where L is the direct path length in m, n is the number of fractures, and w is the average width of the
fractures.
For this particular example, the value of Vp is reduced to 3840 m s -1. If the rock mass is dry, i.e.
the joints are air filled, the calculated value of Vp is 3130 m s - 1 . Clearly, the larger the number and
width of the fractures and the lower the velocity of the infill, then the lower will be the measured
velocity.
In practice, the velocity of both P- and S-waves through dry fractured rock is usually much lower
than the values calculated using the time-average formula. This is because air filled fractures will
tend to act as acoustic barriers, reflecting back most of the incident energy, and the signal recorded
by the seismograph is for an event diffracted around the end of the fracture or crossing the fra.cture at
points of contact. As the assumed distance is unaltered, the increased travel time from the actual
distance of travel will result in a lower P-wave velocity value being recorded for the dry rock mass.
624 Geophysics

Water is a good couplant and the time-average formula will usually be valid for saturated rock
masses, as it will for fractures infilled with soil or secondary mineralization. In such cases, the use of
shear waves is advantageous as the velocity of shear waves through soils is very low and shear waves
are not transmitted through water at seismic frequencies.
A more comprehensive account of the effect of fracture state upon compressional wave velocity
values is given by McDowell [8]. Laboratory experiments were carried out upon blocks and cores of
limestone which were cut to simulate fractures. The results confirmed the significant effect of
fractures upon P-wave velocity and evaluated the importance of different fracture parameters, such
as length, aperture width, spacing and infill.

25.3.2 Fracture Index


The relationship between seismic velocities measured in the field and corresponding values
measured upon intact rock samples in the laboratory can be used as an index of the rock mass
fracture state. Onodera [9] suggested the ratio of the compressional wave velocity measured in situ
(VF ) to that measured in the laboratory (Vd as a fracture index. Subsequently, Deere et ai. [10]
showed that the square of this ratio (velocity index) is numerically equal to Rock Quality
Designation (RQD), which is dependent upon fracture state. For boreholes, RQD is expressed as the
percentage of solid core recovered greater than 0.1 m in length.
This correlation between velocity index and RQD now appears to be tenuous. Cratchley et ai.
[11] found little correlation between fracture spacing and P-wave velocity in tightly jointed rock at
the site of the Foyer's pumped storage scheme in Scotland. Sj0gren et ai. [12] obtained a good
correlation of average relationships between compressional wave velocity, cracks per meter and
RQD for unweathered igneous and metamorphic rocks in Scandinavia but at specific sites found
that both Vp and Vs were required to assess rock mass quality. More recent theoretical considera-
tions of wave propagation in anisotropic and fractured elastic media have confirmed the complexity
of the relationships [13].
Results relating seismic velocities to rock mass quality obtained from various authors [14, 15] are
shown in Table 2.

25.3.3 Attenuation
The attenuation of wave amplitudes is also influenced by rock fractures. Compressional wave
attenuation is considered to have considerable potential for dry rock fracture evaluation by various
workers [15, 16]. Technical difficulties include maintaining a constant energy source and good
seismometer coupling to the ground, but with a multigeophone array this problem is minimized.
Seismic attenuation data are usually presented in terms of either the dimensionless seismic
attenuation factor Q or the attenuation coefficient Ct.. The following formulae relate these parameters
to seismic velocity V and the dominant frequency f
Q = nf/rxV

where Ct. is in units of inverse length and

Q = 8.868n/rx

where Ct. is in units of decibels per wavelength (dB A. -1).

Table 2 Seismic Evaluation of Rock Mass Quality

VF
Rock quality
classification
RQD
(%j
Fracture frequency
(m-1j VL (~:r
Very poor 0-25 15 0-0.4 0-0.2
Poor 25-50 15-18 0.4-0.6 0.2-0.4
Fair 50-75 8-5 0.6-0.8 0.4-0.6
Good 75-90 5-1 0.8-0.9 0.6-0.8
Excellent 90-100 1 0.9-1.0 0.8-1.0
Seismic Investigation for Rock Engineering 625
Table 3 Relation Between Q and IX Values and Rock Mass Type (after
Hatherly [17])

Q IX (db A. -1) Rock mass description

20-100 1.36-0.27 Clastic sedimentary rocks


e.g. sandstones and shales
150-600 0.18-0.05 Metamorphic rocks
e.g. slates and phyllites
200-600 0.14-0.05 Igneous rocks
e.g. granites and basalts

Table 4 Some Results of Seismic Data Analysis with Rock Mass Ratings

Dominant lithology Q value Standard error Vp (ms -1) RMR


ofQ

Shale 8.69 2.70 2300 45


Shale and fractured,
weathered sandstone 6.20 1.41 2330 47
Sandstone 17.42 5.84 3100 59

Table 3 shows typical values relating these attenuation parameters to rock mass type obtained by
Hatherly [17]. Table 4 shows a comparison between Q values, and rock mass quality for some
Devonian shales and sandstones in Ireland [18]. These are limited investigations but they do
provide a relationship between seismic attenuation and the fracture intensity of a rock mass. More
intense fracturing, such as in faulted zones, produces more severe attenuation of P-waves in dry rock
and of S-waves in saturated rock.

25.4 OTHER FACTORS AFFECTING WAVE VELOCITY


The sensitivity of seismic velocity variation to fracture state is reduced to some extent by other
variables which affect the seismic velocity such as lithology, moisture content, in situ stress and the
wavelength of the seismic events. A geological investigation, incorporating boreholes, carried out in
conjunction with the seismic surveys, will enable these other factors to be assessed. Laboratory
measurements of P- and S-wave velocities through rock samples in the laboratory at different
moisture content and axial stress values will also assist this evaluation. A full account of sonic
laboratory methods is provided in Chapter 24 of this volume.

25.4.1 Moisture Content


In general the compressional wave velocity of the ground will increase with increasing moisture
content, whilst shear-wave velocities remain relatively unaffected, but for many rock types the
relationships can be quite complicated. Laboratory experiments, using ultrasonic pulse techniques
on core samples of sandstones and limestones often show a reduction in Vp for a small increase in
moisture content from zero, followed by an increase in velocity with increasing moisture content
until saturation. Clearly, the increase will be greater for rocks with high porosity and relatively low
velocity, such as chalk, than for sound igneous rocks; perhaps 20% compared to 1%. In mudstones
and other argillaceous rock types, it is possible that there will be an inverse relationship between Vp
and moisture content due to chemical or volume changes.
When considering in situ measurements on fractured rock masses the effect of variations in
moisture content upon P-wave velocity can be even more significant. This is the result of improved
transmission across fractures, hence the increase in P-wave velocity with saturation will be much
greater for fracture zones than for intact sound rock.
626 Geophysics

25.4.2 In Situ Stress


The variation of velocity with stress is an important consideration for subsurface excavations such
as tunnels and underground storage chambers. Generally, the velocity will increase with stress
primarily through the acoustic closure of fractures. Laboratory experiments carried out on rock
samples at the UK Transport and Road Research Laboratory [15] showed an initial rapid increase
in P-wave velocity with increasing uniaxial stress as the fractures closed, followed by a slight increase
in velocity with deformation of the intact material. Low normal stresses in the range 0.1 to
0.75 MN m -2 were sufficient to acoustically close a single simulated fracture in rock types varying
from weak sandstone to strong basalt. This suggests that single subhorizontal fractures could be
closed by overburden pressures at shallow depths; between approximately 4 m and 30 m, respect-
ively. It should be noted, however, that the simulated fractures were flat and sometimes polished;
which, in nature, would probably only occur in shear joints and cleavage planes. Tension joints
produced by the cooling of igneous rocks and by tectonic stresses can remain open in stronger rock
types to depths of hundreds of meters, especially with high porewater pressures. In most cases,
subvertical fractures will remain open to greater depths than subhorizontal fractures and considera-
tion should be given to the direction of the ray paths relative to the fracture orientation.
New and West [19] conclude, from the results of field and laboratory experiments, that
measurements of P-wave velocity may not always be sensitive to the frequency of joints in a rock
mass. A low velocity ratio probably indicates poor rock quality, irrespective of depth, and zones of
intense fracturing or disruption by faulting can be readily identified seismically. For deep sites, where
fractures are acoustically closed, there may be no correlation between P-wave velocity and fracture
spacing. This type of problem was noted by Cratchley et ai. [11] in their report of seismic
measurements in a tunnel in metamorphic rocks at approximately 100 m depth for the Foyers
Hydroelectric Scheme in Scotland. New and West [19] obtained similar results for the Kielder
Aqueduct Tunnel, northern England, but they also recorded significant reductions in P-wave
velocity, related to closely spaced joints and faults in mudstones, where the depth of cover exceeded
100 m.

25.4.3 Frequency
The dominant frequency of energy sources and the corresponding wavelength of the seismic waves
affect the resolution of seismic methods as well as the recorded wave velocity and attenuation.
Through intact material there appears to be only a small variation in P-wave velocity with
frequency, and a different relationship for dry and saturated rocks [8]. However, the ability of a
fracture to transmit either P- or S-waves depends upon the relationship of wavelength to gap width,
as well as the nature of the material within the fracture. A 'closed' joint in a dry rock mass may be an
acoustic barrier at wavelengths of a few millimeters when ultrasonic methods are used with sources
of natural frequency in the range 100 kHz to 1 MHz. The same joint would transmit most of the
incident energy at wavelengths of a few tens of meters, as produced by seismic sources utilizing
frequencies in the range 100 Hz to 1 kHz.
Whether or not a significant amount of energy is transmitted across an air gap can be calculated
from the ratio of wavelength to gap width [20].
Table 5 illustrates the relationship between gap width, frequency and attenuation for a dry strong
limestone for which the compressional wave velocity without visible fractures is 4000 m s - 1. This

Table 5 Relationship Between Gap Width, Wavelength and Transmission Coefficient for Dry Bem-
bridge Limestone

f }. la for t b = 0.07% Ifor t = 7% Ifor t = 88%


(m) (m) (m) (m)

I Hz 330 10 1 10- 1
10Hz 33 1 10- 1 10 -2
100Hz 3.3 10- 1 10- 2 10- 3
1 kHz 3.3 X 10- 1 10- 2 10- 3 10- 4
10kHz 3.3 X 10- 2 10- 3 10- 4 10- 5
100kHz 3.3xl0- 3 10- 4 10- 5 10- 6
1 MHz 3.3 X 10- 4 10- 5 10- 6 10- 7

al = gap width. b t = transmission coefficient.


Seismic Investigation for Rock Engineering 627
table shows that for 7% of the normally incident energy to be transmitted, the wavelength must be
approximately 300 times as large as the air gap. This explains the dramatic attenuation that is
usually observed through dry fractured rocks at ultrasonic and seismic frequencies across open
fractures and the consequent severe reduction in the recorded P-wave velocity. A similar relation-
ship is anticipated for transmission of shear waves across water filled fractures, whereas the
transmission of compressional waves will be dramatically improved by the presence of water or solid
material within the fractures.

25.5 ASSESSMENT OF ELASTIC MODULI


25.5.1 Dynamic Elastic Moduli
Geophysical measurements, particularly by seismic pulse techniques, are being increasingly used
to determine the elastic properties of rock masses, largely because the tests are carried out in situ at
an appropriate strain rate. P- and S-wave velocities can be used to calculate the dynamic values of
Young's modulus (ED), shear modulus (Go) and Poisson's ratio (J.lo) using the following formulae

Eo = 2p(1 + JlO)(VS)2
Go = p(VS)2
and

where p is the bulk density of the rock mass.


As Vp can usually be measured more readily and accurately than Vs , it is tempting to use the
following formula

with an assumed value of Poisson's ratio. This may provide reasonably accurate values of ED for
strong, dry, massive, unweathered rock masses, where Vp is in excess of 3000 m S-l and Poisson's
ratio is expected to be between 0.1 and 0.2. However, where the rock mass is composed of weaker
rock types, or has relatively low Vp values because of weathering, alteration or fracturing, Poisson's
ratio will probably be within the range 0.2 to 0.4 and should be measured. To assume a value of 0.25
for J.lo, when the actual value is 0.4, would produce an error of + 80% in ED. The measurement of
both Vp and Vs is, therefore, to be recommended in most cases where accurate values of Young's
modulus are required. Below the water table, Vp should never be used without corresponding values
of Vs to determine dynamic elastic moduli.
When the ground is expected to be subjected to dynamic loading, the shear modulus (Go) is also
often required. This parameter can be calculated directly from the shear-wave velocity, as shown. An
example is the design of foundations for reciprocating machines [21], where consideration also has
to be given to the possibility of resonance. The dynamic shear modulus is also required to assess the
response of the ground to earthquake loading, particularly where geological faults are present.

25.5.2 Static Elastic Moduli


Seismic methods can also be used to assess the deformability of the rock mass under static loading
conditions produced by large engineering structures, such as concrete dams. Bieniawski [22]
recommends the use of the following formula for initial assessments
Es = 2(RMR) - 100 GPa

where RMR is the Rock Mass Rating, and E s is the static elastic modulus.
All of the parameters used in the RMR classification system - the strength of the intact rock
material, the spacing and condition of fractures, and groundwater conditions - all affect the seismic
wave velocities and attenuation. Seismic characterization of a site can, therefore, be used as a basis
628 Geophysics
for the location of the more expensive methods utilizing static loading conditions; such as plate-
bearing tests. Once a direct relationship has been established between ED and E s , or between ED and
the modulus of deformation, the seismic velocities can be used to assess Es and the modulus of
deformation for the remainder of the site.
In an extensive geological and geotechnical survey carried out in the Middle Chalk at Mundford,
Norfolk, UK, Ward et ai. [23] were able to classify the rock mass by observation into five
weathering grades and to relate this classification to deformability, based upon the results of plate-
bearing tests. Subsequent seismic refraction surveying by Grainger et ai. [24] established typical
ranges of P-wave velocity for these grades. Although of considerable value to this particular site,
where the construction of a sensitive engineering structure was proposed, extrapolation of the results
outside the immediate calibrated area was not recommended.
At less accessible sites, or at an early stage in an investigation programme, there may be very little
opportunity for direct observation of rock quality or for static loading tests. In such cases, the
seismic results can provide an assessment of rock deformation under static loading, especially if
sonic tests are also carried out on representative rock samples in the laboratory. There have been
many attempts to establish empirical relationships between ED and E s [25-29] based upon in situ
velocity measurements. All of them are different and relate only to the specific site geology. A more
general empirical relationship between ED and E s has been provided by Coon and Merritt [30],
based upon plate jack or pressure chamber tests and P-wave velocity measurements. As a rule of
thumb, for rock masses with velocity index values less than 0.6, Es can be taken to be approximately
15% of ED.
Initial assessments could also be made by use of the Petite Seismique method. This was introduced
by Schneider [31] and is discussed in the context of rock mass deformation assessment by
Bieniawski [22]. Essentially, this method involves the evaluation of the frequency component of
shear waves and comparison with similar results obtained from other sites where the static modulus
of deformation has been established. A linear relationship of the form
Es = 0.054J - 92

wherefis the shear wave frequency in Hz, has been established for several large construction sites in
South Africa. Unfortunately, the results apply only to siltstones, sandstones, greywackes and
phyllites with E s values in the range 2 to 42 GPa and the relationship may not be applicable to other
rock types, particularly very weak ones.

25.6 CASE STUDY


25.6.1 Site Conditions and Procedures
The results of a ground investigation carried out at the proposed locations of generators for the
Akjoujt power station in Mauritania, West Africa, are presented to illustrate the advantages and
limitations of seismic techniques for assessing rock mass deformation moduli. The site is located at
the Akjoujt copper mine, approximately 4 km to the west of Akjoujt in Mauritania, West Africa,
(Figure 3). An area of approximately 900 m 2 at the western end of the existing power station was the
proposed location of three new generator sets and associated plant (Figure 4). The surface was
almost level and covered with tarmac at the time of the site investigation. Subsurface conditions
could be assessed from inspection of the adjacent mine pit and road cuttings, where a surface layer of
ferruginous conglomerate of thickness 1.0-2.0 m partially covered very weathered steeply dipping
beds of schist and tuff. A hole previously bored to investigate the ground beneath the existing power
station recorded nil core recovery from 0-6.6 m and a small recovery of 53 % and 30% of weathered
schist from depth ranges 6.6-8.0 m and 8.0-10.0 m respectively. A large truck mounted rotary drill
was set up at BHl (Figure 4) but it proved to be unsuitable for obtaining core samples from this type
of ground because of the inclination of the beds and the weathered and fractured nature of the rocks;
eventually only providing rock chippings to a depth of 15 m. Although the water table was expected
to be at a depth of approximately 50 m, water was encountered at a depth of 10.5 m in this borehole
and rose to a depth of 9.0 m after 5 days. A further surprise was the significant amount of moisture
within the near-surface residual soils (7-10%) and rocks (1-5%), presumably protected from
evaporation by the tarmacadam surface.
Seismic refraction surveying was carried out for two spreads, each of 30 m length, to investigate
the possibility of lateral changes in seismic velocity values over the site. The localities of these
spreads relative to the boreholes and proposed generator locations are shown in Figure 4. A
sledgehammer blow to the ground, via a metal impact plate, was used as the energy source in each
Seismic Investigation for Rock Engineering 629

Sahara
Desert

Scale

oLI_ _.J...1_
400 800 1200 km
_....L1_ _--'1

Figure 3 Location of Akjoujt power station

TP~(b)
== ==
TPI(a)

IeH051
I I
~SU"tPlt

I
BHI I BH7 BH3
I Seismic I 30m
L_..J
r-,I

!
1 Proposed
I I generator sites
L_J
TP =Trial pit
Approx N BH =Borehole

E o 5m
o Scale L..I_ _~I

Figure 4 Layout of seismic spreads and boreholes - Akjoujt

case and vertical velocity geophones of 10 Hz natural frequency were used as the sensors. First
arrival P-wave events were confidently recorded and accurately timed but the time-distance graphs
were slightly complicated for Spread 1 by the presence of a thin surface layer of cemented
conglomerate (Vp = 980 m s -1), over decomposed schists (Vp = 570-750 m s -1), i.e. a velocity
inversion. The lower velocity values of the schists were established from the results of Spread 2, along
which the conglomerate was either absent or too thin or weak to affect the results. The P-wave
630 Geophysics

velocity values for both seismic refractors within the schists could be accurately established along the
length of each of the spreads because of the large amount of data available from forward and reverse
shooting to multigeophone arrays. Also, depths to each refractor at each geophone station could be
calculated using time-depth analysis. A further advantage of this procedure was the recognition of
velocity anisotropy resulting from the relationship of the spread orientation to the strike of the beds,
the variation being from 1200 m s -1 to 1630 m s -1 at a depth of approximately 5 m.
Shear-wave analysis could also be carried out with some confidence from the results ofthe seismic
refraction surveying. This was unexpected as a reversible horizontal shear-wave source and
horizontal geophones are usually required to produce records from which shear-wave events can be
recognized with confidence. The Vs values obtained were used with corresponding Vp values to
calculate the elastic parameters recorded in Table 6.
Cross-hole shooting was carried out between the boreholes along seismic spread 1 (Figures 4 and
5). Borehole 1 was lined with 75 mm diameter steel casing, grouted in place with cement, but
borehole 2 was an open hole of approximately 100 mm diameter. A hydraulic system was used to
clamp the borehole three-component geophones against the sides of these boreholes, at depth
intervals of 1.0 m to a depth of 15 m. The energy source used was a mechanical impact to the base of
boreholes 3 and 7, which was advanced in each case at 1.0 m intervals as the geophones were
lowered in the adjacent boreholes. The travel times recorded for compressional and vertically
polarized shear-wave events were converted into corresponding velocity values, assuming that the
ray-path at each 1.0 m depth level was a direct path between the impact point and the borehole
geophones. As the depth increased, the travel times between BH3 to BH2 eventually became larger
than the corresponding travel times between BH2 and BH1, at the same depth level. It is likely that
refracted events were being recorded as first arrivals between the borehole geophones, and that the
velocity values were slightly too high for each depth level. Below the water table, i.e. at depth ranges
10-15 m, the P-wave velocity values were significantly increased because of water saturation of the
rocks and the decrease in weathering.
The velocity values from the cross-hole shooting were also used to calculate the elastic parameters
shown in Table 6. Bulk density values were assessed from laboratory measurements on borehole
samples and bulk in situ density tests in trial pits. At depth intervals 1.0 to 4.0 m, the modulus values
obtained from cross-hole shooting data were much higher than from the refraction survey results.
This is possibly because of refraction at the base of the surface layer of cemented conglomerate
leading to erroneously large velocities from the cross-hole shooting. However, the identification of
second arrival events as shear waves from the refraction results may have been incorrect.
Down-hole shooting was carried out at only one of the boreholes, i.e. BH2, with two three-
component geophones being placed in the hole at 2 m spacing. Clamping of the geophones was only
possible to a depth of 5.0 m because of damage to the hydraulic system resulting from the irregular
nature of the unlined borehole. From depths 5 to 13 m, coupling of the geophone to the host rock
was achieved by filling the boreholes with water. A rectangular wooden block was bedded into the

Table 6 Akjoujt - Dynamic Elastic Moduli and Velocity Ratios

Depth Eo (GPa) Ii Go (GPa)


(m) R" CR b DR c R CR DR R CR DR
(~:) (~:r
1.0 0.34 0.72 0.40 0.32 0.13 0.40 0.15 0.02
2.0 0.51 1.34 1.66 0.40 0.30 0.28 0.20 0.81 0.76 0.15 0.02
3.0 1.33 2.87 0.32 0.33 0.32 0.71 1.06 0.20 0.04
4.0 1.73 3.51 1.58 0.32 0.29 0.32 0.81 1.48 0.59 0.23 0.05
5.0 2.1-4.2 3.41 0.34 0.33 0.82-1.56 1.29 0.32 0.10
6.0 4.23 2.43 0.33 0.31 1.58 0.60 0.35 0.12
7.0 4.90 0.28 2.12 0.37 0.13
8.0 5.12-5.55 0.26-0.35 2.01 0.40 0.16
9.0 6.7 d 0.32 d 2.51 0.42 0.17
to.O 6.2 0.35 2.35 0.43 0.18
11.0 17.2 0.33 6.56 0.69 0.47
12.0 17.4 d 0.33 d 6.56 0.69 0.47
13.0 17.6 0.33 6.56 0.69 0.47
14.0 13.3-14.3 0.2-0.3 5.55 0.60 0.36
15.0 17.2 d 0.33 d 6.56 0.69 0.47

"Refraction. bCross-hole. cDown-hole. dlnterpolated values.


Seismic Investigation for Rock Engineering 631

Seismic spread I

BHI BH2 BH7 BH3

o
0--- - -
-----+lr>--------.H-"~ ........- - - - L . . - - - - - R e f r a c t o r

+J Energy sources
Scale 0L..!_ _ -,1 m ... Vertical geophone
o Three-component geophones

Figure 5 Seismic ray path diagram

ground at the ground surface close to the top of the borehole, to enable a sledgehammer to be used
as a reversible horizontal shear-wave energy source. A seismograph capable of reading to ± 10 J.1S
was used to accurately establish the difference in arrival times at the two three-component
geophones. The elastic parameters obtained from these results are shown in Table 6 for the depth
range 0-9 m. Between 9 and 13 m the results were considered to be suspect, despite the fact that
potential coupling errors should have been minimized for investigations below the water table.
Laboratory measurements were subsequently made on prepared cylinders of the few suitable
borehole samples available. An ultrasonic materials tester with 200 kHz transducers was used to
determine the bulk compressional wave velocity (Vd. These values were then used with the
corresponding field compressional wave velocity values (VF) to calculate the fracture index (VF/Vd
and velocity index (VF/Vd2.
The results were very useful in establishing a scaling factor between the laboratory values of
Young's modulus determined statically and the in situ static Young's modulus values.

25.6.2 Assessment of Results


The example provided illustrates many of the difficulties that may be experienced with in situ
seismic pulse techniques for assessment of elastic properties. Problems can arise from a misconcep-
tion of the basic principles of wave propagation, particularly in layered media, or from an
underestimation of the variability of ground conditions. There are probably very few sites where a
straight line is the only ray path between the source and receiver and the advantages and limitations
of all available equipment, field procedures and data processing techniques should be fully
considered before field tests are carried out. Sometimes at remote overseas locations, such as the
Akjoujt site, the most suitable drilling rigs and test equipment may not be available.
The seismic wave velocity is by definition the speed of advance of the wave front in one direction,
i.e. along part or all of a particular ray path. Figure 5 indicates the approximate ray paths for seismic
refraction surveying, cross-hole 'shooting' and down-hole 'shooting' for a multilayered ground, the
layering being in terms of seismic velocity and not necessarily lithology. For seismic refraction and
interborehole investigations, the recorded velocity values are for direct rays, or ray paths which are
essentially parallel to the seismic refractors, i.e. usually subhorizontal, whereas for down-hole
measurements the ray paths are usually subvertical. Allowance has to be made for refraction of the
ray paths when down-hole 'shooting' in layered ground. The seismic velocity values and derived
elastic properties may vary with direction both in section and in plan at one particular part of the
rock mass and the technique used must, therefore, provide the appropriate information for design
purposes. Care should also be exercised in relating the dynamically determined moduli to those
obtained statically, such as from plate bearing and borehole pressuremeter tests, which are also
directional.
A further problem, indicated by Figure 5, is the critical refraction of seismic events from acoustic
interfaces, which can result in the acceptance of erroneously large velocity values and unconservative
632 Geophysics

modulus values for design purposes. This problem is discussed fully by McCann et al. [32], and
Butler and Curro [33] but the main points warrant further comment. When recording either P- or
S-waves, the critically refracted events will overtake those traveling on a direct path when the source
and receivers are close to the refractor. The velocity values for the refractor will then be assigned to
the measurement depth. This problem can be minimized by using the results of the seismic refraction
survey to assess the velocity values at and above the refractor of interest and to calculate the
optimum borehole spacing.
Most currently available borehole shear-wave energy sources impart a vertical impact to the
ground and generate vertically polarized shear waves. Also, strong P-wave energy is transmitted
downwards and is reflected or refracted back to the detectors. Strong motion of a vertical geophone
will be produced by all of these events but when the source and receivers are close to reflectors or
refractors the P-wave events will probably mask the direct S-wave event. This can also occur at
shallow depths when the shear-wave source is reversed, because of reflection from the ground surface
or refraction from a high-velocity surface layer, such as concrete, above the measurement level.
Combined use of the records from downward and upward impacts can assist the identification of
shear-wave events but borehole spacing is now even more critical as the P-wave velocities are greater
than the corresponding S-wave velocities. Several reversible borehole shear-wave sources have been
developed in recent years, such as those described by Cosma [34] and Bertrand et al. [35, 36] or the
in-hole hammer commercially provided by Bison Instruments [37]. Where such a source is
available, a drilling rig does not have to be provided for the duration of the seismic tests, as was the
case with the Akjoujt survey.
A reversible horizontal impact, as used at the ground surface for refraction or down-hole surveys,
would be advantageous for interborehole surveys. The problem is basically one of providing
sufficient energy to the borehole wall but this limitation can be overcome to some extent by using
signal enhancement recorders and cross-correlation techniques [38]. Horizontally polarized shear-
wave events for direct or critically refracted ray paths are usually readily identified on horizontal
geophones, especially if a reversible source is utilized. Orientation of the horizontal geophones of
three-component systems in-line and transverse to the line of the boreholes is also desirable.
In all seismic tests, the geophones must be firmly coupled to the ground, otherwise there will be
signal distortion and positive timing errors. With multigeophone arrays, as used in seismic refraction
surveying, the effects of coupling errors are minimized by the statistical analysis of a large amount of
data, but this luxury is rarely available for cross-hole surveys because of the expense of drilling a
large number of holes. Down-hole shooting lends itselfto the multisensor approach as evidenced by
the recent developments in vertical seismic profiling for oil exploration. Ideally the geophones
should be clamped firmly to the walls of the boreholes, particularly for dry holes, and there are a
variety of clamping systems which achieve this objective. An interesting recent development is the
use of a pressuremeter to clamp a three-component sensor system into a borehole [35].

25.7 CONCLUSIONS
The main advantage of seismic investigation for rock engineering is in rock mass classification.
Seismic refraction surveys are particularly useful in this respect and can be carried out relatively
quickly and cheaply, compared to most direct investigation methods. Boreholes and trial ex-
cavations can be located to confirm the variability indicated by the seismic results and to investigate
potentially hazardous conditions, such as fault zones. Where a geomechanical classification is
required, it is usually preferable to measure both compressional and shear-wave velocities on site
and on rock samples in the laboratory. As the seismic refraction survey utilizes essentially the same
recording equipment as the more expensive borehole sonic measurements, a combined approach is
usually cost-effective.
A variety of seismic techniques is also available, utilizing boreholes or subsurface excavations,
which are particularly advantageous when rock mass properties are required. A seismic investiga-
tion carried out at the Carwynnen Test Mine in the Carmenellis Granite of Cornwall by New [39] is
an interesting example of measurement from within mine workings and included recently developed
techniques, such as spectral analysis and acoustic tomography. More commonly, measurements are
made between two or more boreholes or between a single borehole and the ground surface. A
combination of surface and borehole measurements is desirable for geological and geomechanical
classification and the derivation of elastic moduli values.
The tomographic reconstruction techniques, originally developed for medical purposes, are now
widely used to map variations of P- and S-wave velocities between a series of sources and receivers.
Seismic Investigation for Rock Engineering 633
Usually travel times are measured between a borehole source and an array of sensors in a second
borehole or at the ground surface. The assumption is often made that the P- and S-wave events
recorded have traveled directly from source to receiver in each case and a straight-line tomography
program is used. However, curved ray path tomography is preferred to allow for velocity anisotropy
and refraction [36, 40]. Inadequate coupling will result in inaccurate travel times being recorded,
which might produce spurious velocity and modulus values. The impressive computer processing
and presentation procedures now available do not diminish the need for good geological control and
data acquisition.

25.8 REFERENCES
1. Griffiths D. H. and King R. F. Applied Geophysics for Geologists and Engineers, 2nd edn. Pergamon Press, Oxford
(1981).
2. Hawkins 1. V. The reciprocal method of routine shallow seismic refraction investigations. Geophysics 26, 806-819
(1961).
3. Hagedoorn 1. G. The plus-minus method of interpreting seismic refraction sections. Geophys. Prospect. 7, 158-182
(1959).
4. Prentice 1. and McDowell P. W. Geological and geophysical methods in the location and design of services. Under-
ground Services 4, 15-18 (1976).
5. Green R. The hidden layer problem. Geophys. Prospect. 10, 166-170 (1962).
6. McDowell P. W. Seismic refraction surveying in urban areas. In Proc. IMM Conf. Rock Engineering and Excavation in
an Urban Environment, Hong Kong, pp. 241-247. (1986).
7. McDowell P. W., Hooper J. W., Barker R. D., Darracott B. W., Jackson P. D., McCann D. M. and Skipp B. O.
Engineering geophysics - Report of the Geological Society Engineering Group Working Party. Q. J. Eng. Geo/. 21,1-65
(1988).
8. McDowell P. W. Investigation of the factors which affect the velocity of compressional elastic waves through the
Bembridge Marls and Limestone. M.Phil. Thesis (CNAA) (1974).
9. Onodera T. F. Dynamic investigation offoundation rock in situ. In Proc. 5th U.S. Symp. Rock Mech., Minneapolis, MN
(Edited by C. Fairhurst), pp. 517-533. Pergamon Press, Oxford (1963).
10. Deere D. U., Hendron A. 1., Jr., Patton F. D. and Cording E. J. Design of surface and near-surface construction in
rock. In Proc. 8th U.S. Symp. Rock Mech., Minneapolis, MN (Edited by C. Fairhurst), pp. 237-302. Port City Press,
Baltimore, MD (1967).
II. Cratchley C. R., Grainger P., McCann D. M. and Smith D. I. Some applications of geophysical techniques in engineer-
ing geology with special reference to the Foyers Hydroelectric Scheme. In Proc. 24th Int. Geological Congr., Montreal,
Section 13, pp. 163-175. (1972).
12. Sjogren B. A., Ofthus A. and Sandberg 1. Seismic classification of rock mass qualities. Geophys. Prospect. 27, 409-442
(1979).
13. Crampin S. A review of wave motion in anisotropic and cracked elastic media. Wave Motion 3, 343-391 (1981).
14. Krishnamoorthy K., Goldsmith W. and Sackman J. 1. Measurements of wave processes in isotropic and transversely
isotropic elastic rocks. Int. J. Rock. Mech. Min. Sci. & Geomech. Abstr. 11, 367-378 (1974).
15. New B. M. Ultrasonic wave propagation in discontinuous rock. TRRL Lab. Rep. 720 (1976).
16. Da Gama C. D. Studying rock fractures by wave attenuation methods. In Proc. Symp. Soc. Int. Mechanique des Roches,
Nancy, France, Paper 1-2 (1971).
17. Hatherly P. J. The analysis of shallow refraction seismograms, PhD. Thesis, Macquarie University, Australia (1983).
18. Murphy B. 1. and Rosenbaum M. S. Attenuation from seismic refraction surveying as a ground investigation aid. Q. J.
Eng. Geol. 16, 187-195 (1989).
19. New B. M. and West G. The transmission of compressional waves in jointed rock. Eng. Geol. (Amsterdam) 15, 151-161
(1980).
20. Kinsler 1. E. and Frey P. Fundamentals of Acoustics, 2nd edn. Wiley, New York (1962).
21. CP2012 Part I. Foundations for machinery. British Standard Code of Practice (1974).
22. Bieniawski Z. T. Determining rock mass deformability: experience from case-histories. Int. J. Rock Mech. Min. Sci. &
Geomech. Abstr. 15,237-247 (1978).
23. Ward W. H., Burland 1. B. and Gallois R. W. Geotechnical assessment ofa site at Mundford, Norfolk, for a large proton
accelerator. Geotechnique 18, 399-431 (1968).
24. Grainger P., McCann D. M. and Gallois R. W. The application of the seismic refraction technique to the study of
fracturing of the Middle Chalk at Mundford, Norfolk. Geotechnique 23, 219-232 (1973).
25. Kujundzic B. Correlation between static and dynamic investigations of rock mass in-situ. In Proc. lst Congr. Int. Soc.
Rock Mech., Lisbon, Vol. I, pp. 565-570 (1966).
26. Link H. Evaluation of elasticity moduli of dam foundation rock determined seismically in comparison to those arrived
at statically. In Proc. 8th Int. Congr. Large Dams, Vol. 1, R. 45, pp. 833-858. (1964).
27. Dvorak A. Seismic and static modulus of rock masses. In Proc. 2nd Congr. Int. Soc. Rock Mech., Belgrade, Vol. 2,
pp. 313-317. (1970).
28. Lykoshin A. G., Jashchenko Z. G., Savich A. J. and Koptev V. I. Studies of properties and conditions of rock massifs by
seismic acoustic methods. In Proc. 2nd Congr. Int. Soc. Rock Mech., Belgrade, Vol. 1, pp. 81-87. (1970).
29. Drozd K. and Louma B. The correlation of moduli of elasticity determined by microseismic measurement and static
loading test. In Proc 1st Congr. Int. Soc. Rock Mech., Lisbon, Vol. 1, pp. 291-293. (1966).
30. Coon R. F. and Merritt A. H. Predicting in-situ modulus of deformation using rock quality indexes. In Determination
of the in-situ modulus of deformation of rock. ASTM Spec. Tech. Pub/. 477, 154-173 (1970).
634 Geophysics
31. Schneider B. Contribution a I'etude de massifs de la fondations de barrages. Trans. Lab. Geol. Fac. Sci. Univ. Grenoble,
Monsoir no. 7 (1967).
32. McCann D. M., Grainger P. and McCann C. Inter-borehole acoustic measurements and their use in engineering
geology. Geophys. Prospect. 23, 50-69 (1975).
33. Butler D. K. and Curro 1. R., Jr. Cross-hole seismic testing - procedures and pitfalls. Geophysics 46, 23-29 (1981).
34. Cosma C. Determination of Rock Mass Quality by the cross-hole seismic method. Bull. Int. Assoc. Eng. Geol. 26-27,
219-225 (1983).
35. Bertrand Y., Bozetto P., Lakshmanan 1. and Sanchez M. The use of the sheargun and the sismopressiometre in seismic
studies. First Break 5, 335-342 (1987).
36. Bertrand Y., Herring A. T., Lakshmanan 1. and Sanchez M. Curved ray seismic tomography-17 years experience from
Zaire (1970) to Kenya (1987). In Proc. Soc. Prof Well Log. Anal. 2nd Int. Symp. Borehole Geophysics for Minerals.
Geotechnical and Groundwater Applications, Colorado, pp. 42-50. (1987).
37. Bison Instruments. Bison's Cross-hole shear wave hammer. Advertisements Geophys. 45, A-59 (1980).
38. Wong J., Hurley P. and West G. F. Crosshole audiofrequency seismology in granitic rocks using piezo electric
transducers as sources and detectors. Geoexploration 22, 261-279 (1984).
39. New B. M. The seismic investigation of rock properties at the Carwynnen Test Mine. Dept. Environ. Rep. no.
DoE/RW/85/016 (1984).
40. Cottin 1. F., Deletie P., Jaquet Francillon H., Lakshmanan J., Lemoine Y. and Sanchez M. Curved ray seismic
tomography application to the Grand Etang Dam (Reunion Island). First Break 4, 25-30 (1986).

You might also like