You are on page 1of 10

Assessment of Excavatability in

Sedimentary Rocks Using Shallow


Seismic Refraction Method

Sayed SR Moustafa
Geology and Geophysics Dept., Faculty of Science
King Saud University, Saudi Arabia
National Research Institute of Astronomy and Geophysics (NRIAG), Cairo, Egypt
e-mail: smoustafa@ksu.edu.sa

ABSTRACT
The seismic wave velocities are significant factors in the evaluation of excavation characteristics
of different rock materials. Nine shallow seismic profiles are used to delineate the shallow
subsurface structure of an area in Marsa Matrouh in the northerly portion of Egypt. From the
tomographic inversion, two refractors were detected in the surveyed site indicating the presence of
three distinct velocity layers; starting from the topmost low velocity layer (≈ 450-1250 m/s) to the
high velocity layer of the hard massive bedrock (≈ 3500-4150 m/s). Estimated velocity values are
utilized in the assessment of rippability by tractor mounted rippers. Qualitatively, more than 60 %
of the study area can be classified as rippable, 14 % as marginal, and 26 % as non-rippable.
KEYWORDS:shallow seismic refraction, rippability, north Egypt.

INTRODUCTION
Ripping is one of the primary methods used for mechanical breaking or loosing the soil to a
level sufficient enough for digging and loading. Over the years techniques for preparing
construction programmers and estimating the costs of excavation have also amended. It is
significant in preparing these estimates to know the type of equipment required to excavate a
particular site. Ripping performance is known to be influenced by the material strength and its
discontinuity characteristics (Pettifer and Fookes, 1994). In most cases ripping with a bulldozer is
cheaper than boring and blasting; however, as the ripping becomes harder, the wear and tear on
the bulldozer increases and the productivity decreases. More or less of the surface excavation
assessments are grounded purely on P-wave velocity deduced from the shallow seismic refraction
method.
Seismic refraction method is the most popular and useful for the determinations of rock mass
characterization which can contribute to the selection of an excavation system (Atkinson, 1971).
This method can represent lots of intrinsic rock properties like porosity, density, grain size and
shape, anisotropy, mineralogy, degree of cementation and moisturizing effect of the rock material
(Braybrooke, 1988). For the excavation, this method has proved its usefulness as good indicator
for the prober machinery selection since their first introduction (Tonnizam et al., 2011).
The need for a reliable excavation assessment method has been increased with developing
technology, especially in the newly developed governorate like Matrouh Governorate. It uses up a
wide sector in the northwest of Egypt and is stretched from km 61 in the west of Alexandria up to

- 1373 -
Vol. 20 [2015], Bund. 4 1374

the Egyptian-Libyan borders on the northerly coast of Egypt (Figure 1). The area of Matrouh
governorate is about 212,112 km2 (about 21% of Egypt's area) of which 3921.40 km2 has only
been the total inhabited area. This area ranked the second largest Governorate in Egypt.
In the newly developed industrial area, the work contractors were levelling the ground for a
housing project by ripping hard limestone sediments during the construction. The rock mass has
undergone extensive weathering stage and proper excavator to be used needs to be limited.
Hence, the primary aim of the current study is to delineate the thicknesses of the subsurface
materials and corresponding velocity values in order to determine the excavatability of limestone
rocks. Different velocity values can be correlated to identify different materials. Estimated
velocity can then be correlated to the subsurface structure and materials characteristics in terms of
their type of material, degree of weathering, the thickness of the materials and the depth to the
bedrock (Moustafa et al., 2013). The results obtained will then be used to determine the
rippability of the different materials on site (Caterpillar, 2001).

Figure 1: Location map of the study area. Crosses show the locations of seismic profiles.

MATERIALS AND METHODS


Study Area
Marsa Matrouh is a harbor, summer tourist destination and the capital of Matrouh
Governorate, distinguished by its 7 km long beach. Morphologically the area can be divided into
Vol. 20 [2015], Bund. 4 1375

five sectors. The first is a coastal plain area paralleled to the Mediterranean Sea Coast, with 25-60
km width, formed by deposits made by heavy rain. The second sector is the Libyan plateau,
which adjoins the plain in the south. It is regarded a high surface stretches to the Sidra Gulf in the
west of Tripoli. The third is the called the Qattarah depression. It is a huge depression and
extends from the south of Al-Alamein region. The Oases area is the fourth sector and located
about 17 m below the sea level. It takes a great deal of water springs which are enough for local
expenditure and offer thousands of Acres with water. The last sector is called the great sand sea at
the southwest of Matrouh. The soil of this area is very soft and heavy and extends for thousands
of kilometers. The topography of the region is of singular features. The area contains different
elevated hills varying from 5 m to 90 m above the sea level (Hilmy, 1951).
The geology of Matrouh area is covered by sedimentary rocks that vary from limestone to
lime-sandstone and marl. The country rocks belong to the Miocene, Pliocene and Pleistocene age.
All the beaches are composed of white, loose carbonate sands, well-polished and round. The
loose carbonate sand gradually changes to fairly consolidate limestone-forming ridges that skirt to
coast. The ridges are of marine origin and represent bars and depressions, which separate ridges
from lagoons in which alluvial loam deposits are present, mixed with calcareous sand (Holail,
1993). The depressions that are close to the shore are salt affected, i.e. unsuitable for cultivation.
In winter, salt marsh conditions prevail in the low parts of these depressions. Generally; the soil
in the beaches – that is affected by salt –are unsuitable for cultivation; opposite to the solid in the
wadis and around the highways. The coastline of this region is of sandy rocks covered by soft
sand along the shore, with certain solid rocks headland and sandy beach, with smooth degradation
in the field, which makes a full opportunity for tourist growth. The shoreline is characterized by
the presence of successive bays, formed by rocky promontories. This rocky edge decreases in the
south (Shata, 1955).

Rippability Assessments
Rippability assessment needs evaluation of several rock mass parameters from core borings
and/or geophysical work (Weaver, 1975). Six geological factors which are likely to influence the
assessment of rippability have been defined (Bieniawski, 1989; McCormick Jr, 1983). Five
elements are linked to the subsurface rock masses such as type, structure, hardness, weathering
and fabric while the six elements are immediately linked to seismic wave velocity. The speed of a
seismic wave depends on the density and the degree of cementation of materials. Rock masses
having lower wave velocities are more easily ripped. The seismic wave velocity method for
rippability assessment was developed first in the last century by the Caterpillar Tractor Company
(Caterpillar, 2001). The physical principle used for the determination of rippability is that seismic
waves move faster through rock having a higher mass density than through the rock less
consolidated. The wave velocity is influenced by such geological factors as rock hardness,
stratification, degree of fracturing, and amount of decomposition or weathering, all of which
influence rippability. In general, a lower seismic wave velocity indicates material more easily
rippable (Mohamad et al., 2005; Tonnizam et al., 2011). Caterpillar found that a comparison of
the wave velocities recorded with those obtained in a similar material from previous experience
gives a good indication of ripper performance. They have published charts showing ripper
performance as related to seismic wave velocities for their equipment (Caterpillar, 2001).
Rippability may be qualitative: rippable, marginal, and/or non-rippable. Or it may be semi-
quantitative on a scale of rippability ratings from 0-100: 0 being highly rippable and 100 being
unrippable. In either description, rippability is dimensionless. In the current study, we used
Vol. 20 [2015], Bund. 4 1376

Caterpillar (2001) rating charts (Figure 2) to assess the rippability of investigated sedimentary
rocks based of seismic wave velocity values.

Figure 2: Rippability classification of different rock masses according to their P-wave


seismic velocity values (Caterpillar, 2001)

Seismic Data Acquisition and Processing


Nine seismic profiles were acquired in the study area as seen in Figure 1 to get a complete
coverage of the subsurface. For collecting seismic data, we used Geometrics SmartSeis data
logger. It is consists of 24-channel spread, with one geophone per channel, and five meters
geophone spacing. Seismic line locations were determined by handheld GPS. Each seismic line
employed 24 active geophone stations spaced at 5-meter intervals for a total line length of 120
meters each. Accelerated weight drop energy source points were applied between geophone
station and also included energy source points off the ends of each line for a total of five source
points per line. A total of 1080 linear meters of data was acquired. From the dimensions of the
geophone array, we expected the waves to reach a maximum penetration of 30 to 40 meters, to
obtain subsurface seismic velocity information using refraction method. An example of the raw
records from off-end shot for P-wave is shown in Figure 3a. All data were processed in house, on
our data reduction and plotting workstation. Our refraction seismic processing software utilizes
Wavepath Eikonal Traveltime (WET) tomography (Schuster and Quintus-Bosz, 1993) which
models multiple signal propagation paths contributing to one first break (the Fresnel volume
approach), while conventional ray tracing tomography is limited to the modeling of just one ray
path per first break. An Eikonal solver (Lecomte et al., 2000; Schuster and Quintus-Bosz, 1993)
is used for traveltime field computation which models refraction and transmission of acoustic
waves. As a result, the velocity anomaly imaging capability is enhanced with the WET
tomographic inversion method compared to conventional ray tomography (Sheehan et al., 2005).
A color-coded seismic velocity cross-section of the subsurface has been generated for each line.
Color scaling of these seismic velocity sections is based on the range of seismic velocity values
Vol. 20 [2015], Bund. 4 1377

calculated. Scaling has been normalized for all velocity sections. Examples of estimated time-
distance curve along with the initial velocity model used for tomographic inversion are presented
on Figure 3b.

Figure 3: a- P-wave first-arrival picking for the middle shot. (b) Travel time versus
distance plot for acquired P-wave picking and initial velocity model used for tomographic
inversion of the interpreted data for the first profile (L1) conducted in the Marsa Matrouh
area. The initial model consists of a gradient model with increasing velocities with depth,
reflecting the gross structure of the study area.

RESULTS AND DISCUSSIONS


The seismic refraction method measures the velocity at which a seismic wave propagates
through a soil or rock medium. In this case, the compressional (P-wave) seismic wave was
measured. Higher seismic P-wave velocities indicate the higher density material, thus quantifying
the competency, or strength of the soil or rock medium and providing an estimation of the
rippability and/or excavatability of the subsurface materials. The results of the current
investigation are used to determine rippability of the limestone bedrock layers based on velocity
information. The seismic refraction data were interpreted using the travel-time curves. First-
arrival phases (Figure 3) were picked assuming that, they were refracted from the same interface.
No correction for geophones elevations was needed as the studied areas were flat and horizontal
enough. To estimate the velocity values for each refractor, the WET inversion method is used
utilizing a proposed 1D initial model and the picked travel times. This 1D initial model (Figure
3b) is iteratively refined with 2D WET tomographic inversion. Any artifacts of the 1D initial
model are progressively removed, with increasing number of WET iterations.
The results of the WET tomographic inversion are presented in the form of nine seismic
sections (Figure 4). These seismic velocity sections, which were created through the inversion
process, have low error and provide a sufficient degree of lateral definition of the seismic velocity
horizons found beneath each profile. In our study, we considered seismic P-wave velocities less
than 1,000 m/s indicates native soil, fill material or heavily weathered and/or decomposed rock,
while velocities in excess of 3,000 m/s indicate fresh (essentially non-weathered) rock (Weaver,
Vol. 20 [2015], Bund. 4 1378

1975). Seismic velocities between these two values typically indicate rock with varying degrees
of weathering and/or fracturing. Consolidation and cementation, as good as, fracture spacing and
density as well affect the measured seismic velocities. Moderate velocities may indicate
compacted soil, moderately weathered rock or loosely consolidated sediment such as gravel, sand
and silt. Saturated sediment below the water table characteristically displays seismic velocities
near or slightly above 1,500 m/s (Moustafa et al., 2013).
In our study area initial velocity models were utilized as starting models for tomographic
inversion. The tomographic inversion method iteratively computes the travel time of the initial
model and compares it with the actual data. The initial model is then modified to minimize the
misfit (≈ 0.5 ms) between the computed travel time and the actual data. The final model is
obtained when the misfit or the root-mean square (rms) error converges such that further
iterations would not reduce this error, which usually occurs after less than 15 iterations.
From the tomographic inversion, two refractors were detected in the seismic sections in the
surveyed site indicating the presence of three distinct velocity layers; starting from the topmost
low velocity layer (≈ 450-1250 m/s) to the high velocity layer of the hard massive bedrock (≈
3500-4150 m/s). The final seismic sections obtained from the present survey along the conducted
seismic lines are depicted in Figure 4. They are presented in the form of geologic-geophysical
sections defining zones (layers) with different average seismic wave velocities.
Visual and analytical inspection of the tomographic seismic inversion results indicates that
profiles L1, L2, L4, L6 and L7 could be grouped as they have more or less similar geologic-
geophysical features while profiles L3, L5, L8 and L9 form another group with other distinctive
geologic-geophysical characteristics.
The first group (profiles L1, L2, L4, L6 and L7) has each total length of 120 m and running in
a NW-SE direction. The results of tomographic inversion revealed a subsurface structure (Figure
4) consists of three layers; the first layer is man-made landfill, weathered rock fragments and
alluvium with seismic velocity ranges from 200 to 2500 m/s with maximum thickness reaches 36
meters observed at profile L4. The seismic velocity in the second layer ranges from 2600 to 3100
m/s, indicating a saturated fractured limestone rock deposits with thickness reaches 7 meters.
Compacted limestone basement is mapped below this layer with seismic velocities range from
3100 to 4000 m/s. The interpreted subsurface model for these lines shows variation of the
interfaces between the layers with different velocities indicating heterogeneity of the
composition. Their probable classification is given in Table 1.
The second seismic group (profiles L3, L5, L8 and L9) also each has a total length of 120 m
(Figure 1) and running W-E direction. The results of tomographic inversion revealed a subsurface
structure (Figure 4) consists of three layers; the first layer is unsaturated overburden and rock
fragments with seismic velocity ranges from 236 to 2590 m/s with thickness reaches 5.8 meters.
The seismic velocity in the second layer ranges from 2600 to 3200 m/s, indicating a compact
saturated alluvium and/or fractured saturated limestone rocks with thickness reaches 46.9 meters.
Hard and massive bedrock are mapped below this layer with seismic velocities range from 3280
to 4000 m/s. The interpreted subsurface models for these profiles shows low velocity zones
centered at distances of 55 and 60 meters from the first shot point. These low velocity/depression
zones indicates the existence of shear zones cutting the massive bedrock, filled with alluvium
deposits, and fractured rocks that might associate with faulting affecting the site. The basic
geologic units were identified based on the interpreted seismic velocities and the available site-
specific information. Their probable classification is shown in Table 1.
Vol. 20 [2015], Bund. 4 1379

Figure 4: Final ground models form the acquired seismic profiles at Marsa Matrouh resulting
from tomographic inversion.

Table 1: Obtained seismic wave velocities for Marsa Matrouh area.


Profile Layer Velocity Thickness Description
(m/s) (m)
1st 449 – 2597 36 Weathered rock fragments and alluvium
L1 2nd 2600 – 3196 4.5 Saturated fractured limestone rocks
3rd 3204 – 4000 – Compacted limestone basement
1st 347 – 2598 30 Weathered rock fragments and alluvium
L2 2nd 2600 – 3199 3 Saturated fractured limestone rocks
3rd 3200 – 4000 – Compacted limestone basement
1st 311 – 1598 22.5 Weathered rock fragments and alluvium
L3 2nd 2601 – 3194 5.3 Saturated fractured limestone rocks
3rd 3200 – 3874 – Compacted limestone basement
1st 286 – 2598 35 Weathered rock fragments and alluvium
L4 2nd 2601 – 3195 3.4 Saturated fractured limestone rocks
3rd 3200 – 4000 – Compacted limestone basement
1st 236 – 2598 49.5 Weathered rock fragments and alluvium
L5 2nd 2600 – 3198 7.5 Saturated fractured limestone rocks
3rd 3203 – 4000 – Compacted limestone basement
1st 315 – 2596 20 Weathered rock fragments and alluvium
L6 2nd 2600 – 3198 5.7 Saturated fractured limestone rocks
3rd 3200 – 4000 – Compacted limestone basement
1st 200 – 2596 31 Weathered rock fragments and alluvium
L7 2nd 2601 – 3198 6 Saturated fractured limestone rocks
3rd 3200 – 4000 – Compacted limestone basement

1st 309 – 2590 49 Weathered rock fragments and alluvium


L8
2nd 2602 – 3196 9 Saturated fractured limestone rocks
Vol. 20 [2015], Bund. 4 1380

3rd 3202 – 3814 – Compacted limestone basement


1st 423 – 2585 49 Weathered rock fragments and alluvium
L9 2nd 2600 – 3197 6.5 Saturated fractured limestone rocks
3rd 3202 – 3870 – Compacted limestone basement

Based on Caterpillar recommendations (Caterpillar, 2001), it was estimated that about


60.03% of the total surveyed volume is rippable, 14.05% is marginal, and 25.92% is non rippable
rock mass. Individual percentages of each class for all the surveyed profiles are given in Table 2.
The final ground rippability classification models interpreted form acquired seismic profiles at
Marsa Matrouh resulting from tomographic inversion is depicted in Figure 5.

Table 2: Percentage distribution of rippable volume


Profile Rippable Marginal Non-Rippable
L1 57.51 % 20.55 % 21.95 %
L2 47.55 % 16.04 % 36.41 %
L3 76.50 % 13.91 % 9.59 %
L4 57.24 % 10.28 % 32.48 %
L5 70.48 % 13.33 % 16.19 %
L6 26.58 % 20.38 % 53.05 %
L7 35.12 % 11.94 % 52.94 %
L8 83.75 % 10.71 % 5.53 %
L9 85.57 % 9.28 % 5.15 %

ACKNOWLEDGEMENTS
I highly appreciate the help and support I get from the Seismology Department at the
National Research Institute of Astronomy and Geophysics. I am very grateful to Professor A.
Khairy and all the staff of the Egyptian National Seismological Network for providing the data.
Vol. 20 [2015], Bund. 4 1381

Figure 5: Final ground rippability classification models interpreted form acquired


seismic profiles at Marsa Matrouh resulting from tomographic inversion.

REFERENCES
Atkinson, T. (1971). Selection of open pit excavating and loading equipment. Trans Inst
Min Metall, 80, A101-A129.
Bieniawski, Z. T. (1989). Engineering rock mass classifications: a complete manual for
engineers and geologists in mining, civil, and petroleum engineering: John Wiley &
Sons.
Caterpillar. (2001). Performance Handbook Edition 32. Caterpiller Inc., Peoria, Illinois.
Hilmy, M. E. (1951). Beach sands of the Mediterranean coast of Egypt. Journal of
Sedimentary Research, 21(2).
Holail, H. (1993). Diagenetic trends of the Pleistocene calcareous ridges, Marsa Matruh
area, Egypt. Chemical geology, 106(3), 375-388.
Lecomte, I., Gjøystdal, H., Dahle, A., & Pedersen, O. C. (2000). Improving modelling
and inversion in refraction seismics with a first‐order Eikonal solver. Geophysical
Prospecting, 48(3), 437-454.
McCormick Jr, W. N. (1983). Engineering and Design: Rock Mass Classification Data
Requirements for Rippability: DTIC Document.
Mohamad, E. T., Kassim, K. A., & Komoo, I. (2005). An overview of existing rock
excavatability assessment techniques. Jurnal Kejuruteraan Awam, 17(2), 46-59.
Moustafa, S. S., Ibrahim, E. H., Elawadi, E., Metwaly, M., & Al Agami, N. (2013).
Seismic refraction and resistivity imaging for assessment of groundwater seepage
under a Dam site, Southwest of Saudi Arabia. International Journal of Physical
Sciences, 7(48), 6230-6239.
Vol. 20 [2015], Bund. 4 1382

Pettifer, G., & Fookes, P. (1994). A revision of the graphical method for assessing the
excavatability of rock. Quarterly Journal of Engineering Geology and Hydrogeology,
27(2), 145-164.
Schuster, G. T., & Quintus-Bosz, A. (1993). Wavepath eikonal traveltime inversion:
Theory. Geophysics, 58(9), 1314-1323.
Shata, A. (1955). An introductory note on the geology of the northern portion of the
Western Desert of Egypt. The Desert Institute Bulletin, Egypt, 5, 96-106.
Sheehan, J. R., Doll, W. E., & Mandell, W. A. (2005). An evaluation of methods and
available software for seismic refraction tomography analysis. Journal of
Environmental & Engineering Geophysics, 10(1), 21-34.
Tonnizam, M. E., Rosli, S., Muhazian, M., & Fauzi, M. (2011). Assessment on
Excavatability in Weathered Sedimentary Rock Mass Using Seismic Velocity
Method. Journal of Materials Science and Engineering A: Structural Materials:
Properties, Microstructure and Processing, 1(2), 258-263.
Weaver, J. (1975). Geological factors significant in the assessment of rippability. The
Civil Engineer in South Africa, 17(12), 313-316.

© 2015 ejge

You might also like