Classifying Ground Rippability and Weathering Grades in a Sedimentary Rock Geological Environment Using Seismic Refraction Survey

Abstract

Introduction

An in-depth understanding of the ground subsurface is crucial for foundation design and excavation works and for avoiding potential hazards during land development.In this regard, the ground rippability and weathering grades are some of the ground information needed. While geotechnical works are preferred, their limited horizontal coverage and high cost are often constraints that limit their use.

Aims

To counter this, a geophysical survey is employed for its wider area coverage and cost-efficiency. Therefore, this study used the seismic refraction method to assess the rippability and weathering grades in a sedimentary rock geological setting (interbedded sandstone, siltstone, and shale) as a preliminary ground assessment.

Methods

A seismic refraction survey was carried out using Aktiebolaget Elektrisk Malmletning (ABEM) Terraloc Pro 2, where the survey line was 115m long. Rippability was obtained by correlating seismic values with the Caterpillar D10R rippability table. Meanwhile, the weathering grades of the ground were determined by correlating the study area with another study area of a similar geological setting.

Results

Within the 39m penetration depth, three layers can be classified from the ground’s P-wave velocity values and D10R Caterpillar rippability chart, which include rippable, marginal, and non-rippable layers. A break in the continuous ground layers could be seen, causing lower velocity values to be sandwiched between high velocities, which signified the presence of fracture. The weathering grades were also successfully classified from the seismic velocity values.

Conclusion

Using seismic refraction method, this study successfully employed seismic velocity values in determining the rippability and weathering grades of interbedded sedimentary rock without borehole record.

Keywords: Rippability, Earthworks, Excavation, Cut and filled, Bedrock mapping.

1. INTRODUCTION

The study of subsurface geology using intrusive techniques, including soil borings, rock coring, and borehole logging, has its limitations where the data obtained is only applicable to a specific point in the survey area, which is not a good representation of a wider space area [1, 2]. It is less reliable when used in interpreting the surrounding conditions with lateral variations [3]. For this reason, geophysical methods, such as seismic refraction, have been commonly employed to obtain information on the ground subsurface of an area with a wide spatial coverage, both laterally and vertically.

Seismic refraction is known for bedrock mapping prior to any earthworks. This method is able to map bedrock and ground rippability at a lower cost [4]. Furthermore, the method is a non-invasive, user-friendly technology that does not damage the environment and any potential utilities nearby [5, 6]. The usage of this method has risen in subsurface mapping, especially in the field of geotechnical and civil engineering, due to its capability in characterizing rock mass, which leads to the effective selection of geotechnical methods [7-9]. Mining industries also use seismic refraction methods for quarrying, underground opening, ripping, and blasting [10, 11]. This geophysical method studies the qualitative measurements and distinguishes between the properties of rocks, structures, stratigraphy, and mineralization [12]. The seismic refraction method is based on the assumption of increasing density of the ground with depth.

Extensive development has caused a significant increase in the construction of sedimentary rocks in Malaysia over the past years. While the study of sedimentary rocks is exciting, it is also typically challenging compared to other rock types due to its geology, which is particularly varied in its physical properties [13]. Its geology can consist of composites and interbedded sandstone, shale, siltstone, and limestone [14].

Ground rippability and weathering grade must be studied and identified prior to any preliminary designs in civil engineering projects to ease excavations. As rock mass excavation depends primarily on its structural properties, excavation would be easier with greater fractures and rock mass discontinuities [15]. Determi- nation of ground rippability and weathering grade is performed to identify and exclude any uncertainties in selecting the most suitable method for excavation. Therefore, this study would be greatly useful in summarizing the overall cost of the project while avoiding project delays that could potentially lead to project abandonment due to problems that can be minimized or best avoided.

Fig. (1).

Location of survey area [18].

Fig. (2).

Regional geology of survey area [16].

2. SURVEY AREA AND GEOLOGY

The survey area was located at Iskandar Puteri, Johor, Malaysia, and covered 82,000 m2 area, which was cleared and flattened via cut and filled Fig. (1). The area was located on sedimentary rock types: interbedded sandstone, siltstone, and shale of Triassic age, such as shown in Fig. (2) [16]. This tallies with the sandstone and siltstone outcrops found in the survey area Fig. (3). It should be noted that geological information assists in the interpretation of geophysical data as structural complexities affect the strength of ground materials and seismic signals, which may lead to masking of the targeted layers in seismic processing and inaccurate depth computations of all layers beneath it [17].

3. MATERIALS AND METHODS

3.1. Seismic Refraction

This method provides insight into the strength of the ground subsurface materials based on their densities and modulus. It is one of the common geophysical methods used in engineering and environment applications to map bedrock [19]. The method correlates the travel time of the seismic waves (P-wave) with the density and elastic moduli of the earth material in its calculations.

Three seismic refraction survey lines (S1 – S3) were conducted where 24 channels of ABEM Terraloc Pro 2 seismograph were utilised together with 24 vertical geophones with 28Hz frequency each. Selection of geophone frequency plays a role in improving signal-to-noise ratio, as it can cut out unwanted noise, in addition to multiple shot point stacking during data acquisition [20]. A seismic source (18 lb sledgehammer) was used to generate strain energy by imposing it onto the ground to produce roughly 150 joules of energy. The energy propagates radially to produce direct, reflection, and refraction waves. At a particular distance, the waves reached the ground surface, were detected by geophones, and were recorded as a function of time by the seismograph, as depicted in Fig. (4).

All spreads used 5 m geophone spacing and seven (7) shot points. Five (5) shot points were located along the spread, while the other 2 were offset shots (before and beyond the spread) to enable deeper ground subsurface mapping. All survey lines had positive and negative 50 m offsets to increase penetration depth [21]. Fig. (5) shows the seismic lines conducted. Two survey lines, S1 and S2, crossed each other at 54 m distance on SL1 and at 90 m distance on SL2. The lines were designed in a way that the two crossed lines (SL1 and SL2) would be used to correlate with each other, while the other survey line (SL3) was conducted to map and cover more area.

Fig. (3).

Sandstone outcrop found along the survey line.

Fig. (4).

Seismic wave propagation from the shot point as a direct wave, reflection, and refraction rays [22].

Fig. (5).

Seismic and electrical resistivity tomography (ERT) survey lines conducted [18].

The seismic refraction method has a limitation in detecting thin layers and hidden layers [17, 23]. A thin layer is defined as the minimum thickness of wavelength divided by two (λ/2), which is the maximum amplitude [24]. A layer is considered thin when its thickness is less than the dominant wavelength of the seismic wavefield that illuminates the bed. When the layer is too thin to produce first arrivals, the head wave arrives later in the time-distance graph, therefore making the thin layer (e.g., interbedded soils) a ‘blind zone’ or hidden layer that cannot be resolved [25].

The seismic refraction method is dependent on the basic principle of geological layering, where the density of earth material should increase with depth [24]. In the case where the upper layer (e.g., weathered rock Class II) has similar or greater velocities than the lower layer (e.g., weathered rock Class III), this will result in the lower layer (Class III) being masked/hidden by the upper layer (Class II) as the seismic waves cannot detect it and would cause depth calculations error, especially at greater depths [26, 27]. This is also called a hidden layer. It is vital to understand the concepts of thin and hidden layers for seismic refraction surveys, especially in structurally complex areas.

Data processing was divided into two stages;

3.1.1. Stage 1

The first arrivals of the P-waves were manually picked using SeisOptPicker for all spreads, and the first arrival times were plotted against the geophone distance (travel time graph) for each seismic shot point.

3.1.2. Stage 2

The data for first arrival times, shot positions, and geophone positions, including elevation, were imported into SeisOpt @2D v6.0 commercial software, which inverted the data to produce a P-waves velocity tomo- graphy profile. SeisOpt @2D is a refraction velocity optimization software that only requires the first-arrival picks and array geometry to derive velocity structural information, making the tool ideal in an area with limited information on subsurface velocity structure.

The velocity of seismic pulses depends on the densities and elastic moduli of the materials through which the waves propagate [28]. Therefore, the seismic velocity of a material is not fixed but varies depending on geology, fracturing, density, and elastic moduli [29]. Table 1 shows the seismic velocity values of common ground materials where their values were inferred during interpretation. To achieve an accurate interpretation, a comprehensive analysis was executed using seismic refraction results, geology, and neighbouring information.

Table 1.
Seismic velocities of common ground materials [30, 31].
Material Seismic Velocity (m/s)
Igneous / Metamorphic Granite 4,000 – 5,800
Weathered granite 1,000 – 4,000
Basalt 5,400 – 6,400
Sedimentary rock Sandstone 1,830 – 3,970
Shale 2,750 – 4,270
Limestone 2,140 – 6,100
Unconsolidated sediment Clay 915 – 2,750
Alluvium 500 – 2,000
Sand 200 – 2,000
Groundwater Fresh water 1,430 – 1,680
Salt water 1,460 – 1,530
Fig. (6).

Rippability prediction based on seismic velocity chart of bulldozer D10R of Caterpillar [16].

3.2. Ground Rippability

Ground rippability for the survey area was determined by using the guideline on the correlation between the values of P-wave velocity (Vp) and the Caterpillar D10R rippability chart shown in Fig. (6) produced by Caterpillar Incorporation [32].

3.3. Weathering Grade

Due to the lack of borehole records in the study area, weathering grades of the ground were estimated by referring to the correlation between seismic Vp values and weathering grades from a journal paper published by a previous study [14] with a similar geological setting (sedimentary rock). Referring to Table 2, the seismic profiles of S1 – S3 were classified into five groups of weathering grades: VI, VI-V, V-IV, IV-III-II, and III-II-I.

4. RESULTS AND DISCUSSION

The P-waves refraction first arrivals for all seismic lines S1, S2, and S3 were picked and plotted in distance-travel time graphs as shown in Fig. (7). To process the seismic profiles, seismic refraction tomography was

Table 2.
Correlation between seismic Vp values and weathering grades from borehole records produced by Ismail et al. [14].
Velocity (m/s) Ground Material Description Weathering Grade
< 300 Subsurface soil and much loosened ground. Very dry condition. VI
300 – 800 Gravelly to clayey silty sand. VI – V
800 – 1,800 Gravelly to clayey, silty sand; completely weathered to highly weathered with evidence of highly fractured rock. V – IV
1,800 – 2,400 Highly weathered to Moderately weathered rock. IV – III – II
> 2,400 Moderately weathered to slightly weathered zone and fresh rock. III – II – I
Fig. (7).

Distance-travel time graphs for seismic lines a) S1, b) S2, and c) S3.

Fig. (8).

Seismic refraction profile for seismic lines a) S1, b) S2, and c) S3.

conducted instead of the conventional processing method, Generalised Reciprocal Method (GRM). GRM is good in mapping laterally by assuming the ground subsurface is in layered media with increasing seismic velocities [33, 34]. However, the ground subsurface produced from GRM often contradicts observed ground characteristics due to its limitation in detecting heterogeneity, lateral discontinuities, and gradients [35]. Seismic refraction tomography is superior to GRM as it is capable of resolving both vertical and lateral velocity changes and velocity gradients [7, 36, 37]. Hence, this study applied seismic refraction tomography in processing seismic results to produce seismic velocity profiles, as shown in Fig. (8).

The results show that the ground subsurface in the study area had a velocity of P-waves (Vp) that ranged between 200 – 4,400 m/s with a maximum penetration depth of 26.3 – 39.4 m, as shown in Table 3. The difference in penetration depth between each profile is due to the density and velocity of the ground subsurface beneath the survey lines, where the low-velocity layer in S2 (Vp< 2,000 m/s) is approximately twice thicker than present in S1 and S3. This shows that less competent ground mass would lead to shallower penetration depth, as stated by Alsamarraie [38] and Imani et al. [5], due to seismic signal attenuation. Obermann et al. [39] also stated that the depth sensitivity of seismic waves is affected by the degree of ground heterogeneity and velocity changes, which explains the difference in penetration depth between survey lines S1, S2, and S3.

At the intersection point between lines S1 and S2, the seismic values Vp are consistent in both profiles (54 m distance on S1 and 90 m distance on S2), which signifies good and reliable first arrival pickings. Considering sandstone outcrops can be found at some parts of the survey lines, as in Fig. (3), by referring to Table 1, it can be surmised that the seismic profiles have penetrated bedrock. This is proven by the presence of Vp values that are greater than 1,830 m/s (minimum sandstone value) in all survey lines.

Table 3.
Vp and maximum penetration depth for S1 – S3.
Line Vp (m/s) Depth (m)
S1 200 - 4400 31.9
S2 200 - 4400 26.3
S3 200 - 4400 39.4

From the outcrop on site, certain parts of the seismic survey lines cut across the sandstone layer. Therefore, the ground subsurface is classified based on its rippability using the correlation between P-wave velocity values (Vp) of sandstone and the Caterpillar D10R rippability chart. The correlation produces a three-layer case of the ground, which are rippable (Vp of <2,500 m/s), marginal (Vp of 2,500 – 3,200 m/s), and non-rippable (Vp of >3,200 m/s), as shown in Table 4 and Fig. (9). The rippable layer is thinnest along line S3, which is approximately 15m from the surface, followed by lines S1 and S2. Breaks in the continuous layering seen in lines S1 and S3 are due to the presence of fracture, which causes a decrease in the density of rock mass [40]. This fracture causes a discontinuity in the rock mass, which is a common characteristic of sedimentary rock and affects the change in seismic waves’ travelling velocity via dispersion, in addition to geometric and intrinsic attenuation factors of the fracture structures [41, 42]. This consequently results in a decrease of Vp values in the fracture zone, which is sandwiched between high Vp values. Meanwhile, a boulder can be detected in line S2 at a distance of 0 m, having a higher seismic Vp value than the surroundings [43]. This proves that the more weathered a rock mass is, the lower its velocity. Therefore, layer 1 with the lowest Vp is the easiest to excavate, while layer 3 with the highest Vp is not excavatable using Caterpillar bulldozer D10R due to its solid rock mass.

Fig. (9).

Rippability of the ground based on correlated seismic refraction velocity (Vp) values for lines a) S1, b) S2, and c) S3.

Table 4.
Validated Vp and subsurface layer rippability.
Vp (m/s) Layer Rippability
<2,500 1 Rippable
2,500 – 3,200 2 Marginal
>3,200 3 Non-rippable

Based on Table 2 by Ismail et al. [14], weathering grades of the ground subsurface for all lines were determined as shown in Fig. (10), where the weathering grade decreases with depth. This tallies with the common geological profile where the ground becomes denser with depth due to compaction and lower soil strength [44, 29], while rock mass becomes less fractured with increasing depth [45, 38, 46]. Weathering grades of moderate to slightly weathered zone and fresh rock (III-II-I) were shallower in survey lines S1 and S3, while the weathering grades occurred much deeper in S2.

Fig. (10).

Weathering grade of the ground based on correlated seismic refraction velocity (Vp) values for lines a) S1, b) S2, and c) S3.

CONCLUSION

A seismic refraction survey was conducted at Iskandar Puteri (Johor) to investigate the rippability and weathering grades of the ground in a sedimentary rock geological setting (interbedded sandstone, siltstone, and shale). Based on seismic Vp values and the Caterpillar D-10R rippability chart, the study classified the ground into three layers: rippable, marginal, and non-rippable layers. As the study area does not have any borehole records, weathering grades of the ground were determined based on a similar case study in a published journal. From the seismic Vp values, the ground was successfully classified into five groups of weathering grades: VI, VI-V, V-IV, IV-III-II, and III-II-I.

AUTHORS’ CONTRIBUTION

It is hereby acknowledged that all authors have accepted responsibility for the manuscript's content and consented to its submission. They have meticulously reviewed all results and unanimously approved the final version of the manuscript.

ETHICS APPROVAL AND CONSENT TO PARTICIPATE

Not applicable.

CONSENT FOR PUBLICATION

Not applicable.

AVAILABILITY OF DATA AND MATERIALS

The data and supportive information are available within the article.

FUNDING

This study was financially supported by the Ministry of Higher Education under Long Term Research Grant Scheme (LRGS/1/2016/UTM/01/1/1).

CONFLICT OF INTEREST

The authors declare no conflict of interest, financial or otherwise.

ACKNOWLEDGEMENTS

The authors would like to acknowledge the financial support from the Ministry of Higher Education.

REFERENCES

1
Godio A, Strobbia C, De Bacco G. Geophysical characterisation of a rockslide in an alpine region. Eng Geol 2006; 83(1-3): 273-86. [https://doi.org/10.1016/j.enggeo.2005.06.034].
2
Mohamed AME, Abdel Hafiez HE, Taha MA. Estimating the near-surface site response to mitigate earthquake disasters at the October 6th city, Egypt, using HVSR and seismic techniques. NRIAG J Astron Geophys 2013; 2(1): 146-65.
3
Akingboye A S, Ogunyele A C. Insight into seismic refraction and electrical resistivity tomography techniques in subsurface investigations. Rud-geol-naft zb 2019; 34(1)
4
Washima A, Anti K, Luper Tsenum J. Advantages and limitations of seismic refraction method using hammer sources. ScienceOpen Preprints 2020.
5
Imani P, El-Raouf AA, Tian G. Landslide investigation using seismic refraction tomography method: A review. Ann Geophys 2021; 64(6): SE657-7.
6
Chen W, Hong H, Panahi M, et al. Spatial prediction of landslide susceptibility using GIS-based data mining techniques of ANFIS with whale optimization algorithm (WOA) and grey wolf optimizer (GWO). Appl Sci 2019; 9(18): 3755.
7
El Hameedy MA, Mabrouk WM, Dahroug S, Youssef MS, Metwally AM. Role of seismic refraction tomography (SRT) in bedrock mapping; Case study from industrial zone, Ain-Sokhna area, Egypt. Contrib Geophys Geod 2023; 53(2): 111-28.
8
Wang G, Li R, Carranza EJM, et al. 3D geological modeling for prediction of subsurface Mo targets in the Luanchuan district, China. Ore Geol Rev 2015; 71: 592-610. [https://doi.org/10.1016/j.oregeorev.2015.03.002].
9
Schlindwein V, Bönnemann C, Reichert C, Grevemeyer I, Flueh E. Three-dimensional seismic refraction tomography of the crustal structure at the ION site on the Ninetyeast Ridge, Indian Ocean. Geophys J Int 2003; 152(1): 171-84.
10
Singh TN, Kanchan R, Sigal K, Verma AK. Prediction of p-wave velocity and anisotropic property of rock using artificial neural network technique. J Sci Ind Res 2004; 63(1): 32-8.
11
Kahraman S, Yeken T. Determination of physical properties of carbonate rocks from P-wave velocity. Bull Eng Geol Environ 2008; 67(2): 277-81.
12
Haldar SK. Mineral exploration: Principles and applications 2018.
13
Tucker ME. Sedimentary Rocks in the Field; A Practical Guide 2011.
14
Ismail MAM, Kumar NS, Abidin MHZ, Madun A. Rippability assessment of weathered sedimentary rock mass using seismic refraction methods. J Phys Conf Ser 2018; 995(1): 012105.
15
Bieniawski ZT. Engineering rock mass classifications: A complete manual for engineers and geologists in mining, civil, and petroleum engineering 1989.
16
Geology Map of Iskandar Puteri 1985.
17
Pant PR. A simple solution to overcome large errors due to unknown masked layers in refraction seismics. J Appl Geophys 2023; 219: 105222.
19
Kearey P, Brooks M, Hill I. An introduction to geophysical exploration 2002; 4
20
Medhus AB, Klinkby L. Engineering Geophysics 2022.
21
Tungka M. Determining subsurface geology with seismic refraction tomography survey. IOP Conf Ser Earth Environ Sci 2022; 1003(1): 012037.
22
Reynolds JM. An introduction to applied and environmental geophysics 2011.
23
Yari M, Nabi-Bidhendi M, Ghanati R, Shomali ZH. Hidden layer imaging using joint inversion of P‐wave travel‐time and electrical resistivity data. Near Surf Geophys 2021; 19(3): 297-313.
24
Zelt CA. Seismic refraction methods 2022; 107-16.
25
Schmöller R. Some aspects of handling velocity inversion and hidden layer problems in seismic refraction work. Geophys Prospect 1982; 30(6): 735-51.
26
Foti S, Parolai S, Albarello D, Picozzi M. Application of surface-wave methods for seismic site characterization. Surv Geophys 2011; 32(6): 777-825.
27
Williams RA, Stephenson WJ, Odum JK. Comparison of P- and S-Wave Velocity Profiles Obtained from Surface Seismic Refraction/Reflection and Downhole Data Tectonophysics 2003; 368(1-4): 71-88.
28
Yu C, Ji S, Li Q. Effects of porosity on seismic velocities, elastic moduli and Poisson’s ratios of solid materials and rocks. J Rock Mech Geotech Eng 2016; 8(1): 35-49.
29
Kassab MA, Weller A. Study on P-wave and S-wave velocity in dry and wet sandstones of Tushka region, Egypt. Egypt J Pet 2015; 24(1): 1-11.
30
Jakosky JJ. EXPLORATION GEOPHYSICS 1950.
31
Sharma P V. Environmental and engineering geophysics 1997; 475.
32
Zabron J. Handbook of Ripping 12th ed. 2010.
33
Yufajjiru L, Maryadi M. Soil stiffness identification using fuzzy logic based on seismic tomography and its relationship with dynamic elastic moduli. IOP Conf Ser Earth Environ Sci 2021; 851(1): 012021.
34
Sheriff RE, Geldart LP. Exploration seismology 1995.
35
Sheehan JR, Doll WE, Mandell WA. An evaluation of methods and available software for seismic refraction tomography analysis. J Environ Eng Geophys 2005; 10(1): 21-34.
36
Thurber C, Ritsema J. Theory and Observations – Seismic Tomography and Inverse Methods 2007; 1: 323-60.
37
Whiteley JS, Chambers JE, Uhlemann S, et al. Landslide monitoring using seismic refraction tomography – The importance of incorporating topographic variations. Eng Geol 2020; 268: 105525.
38
Alsamarraie MM. Seismic refraction method in the determination of site characteristics. Iraqi Geol J 2020; 53(2D): 53-63.
39
Obermann A, Planès T, Larose E, Sens-Schönfelder C, Campillo M. Depth sensitivity of seismic coda waves to velocity perturbations in an elastic heterogeneous medium. Geophys J Int 2013; 194(1): 372-82.
40
Doetsch J, Krietsch H, Schmelzbach C, et al. Characterizing a decametre-scale granitic reservoir using ground-penetrating radar and seismic methods. Solid Earth 2020; 11(4): 1441-55.
41
Jug J, Stanko D, Grabar K, Hrženjak P. New approach in the application of seismic methods for assessing surface excavatability of sedimentary rocks. Bull Eng Geol Environ 2020; 79(7): 3797-813.
42
Wenning QC, Madonna C, Zappone A, et al. Shale fault zone structure and stress dependent anisotropic permeability and seismic velocity properties (Opalinus Clay, Switzerland). J Struct Geol 2021; 144: 104273.
43
Cichostępski K, Dec J, Golonka J, Waśkowska A. Shallow seismic refraction tomography images from the pieniny klippen belt (Southern Poland). Minerals 2024; 14(2): 155.
44
Uyanık O. Estimation of the porosity of clay soils using seismic P- and S-wave velocities. J Appl Geophys 2019; 170: 103832.
45
Bahmaei Z, Hosseini E. Pore pressure prediction using seismic velocity modeling: Case study, Sefid-Zakhor gas field in Southern Iran. J Pet Explor Prod Technol 2020; 10(3): 1051-62. [https://doi.org/10.1007/s13202-019-00818-y].
46
Boadu FK. Predicting the transport properties of fractured rocks from seismic information: Numerical experiments. J Appl Geophys 2000; 44(2-3): 103-13.