You are on page 1of 22

Transportation Infrastructure Geotechnology (2023) 10:774–794

https://doi.org/10.1007/s40515-022-00242-6

TECHNICAL PAPER

Dynamic Slope Stability Subject to Blasting Vibrations:


a Case Study of the Jakarta‑Bandung High‑Speed Railway
Tunnel

Widodo1 · Hermawan Phanjaya1 · Simon Heru Prassetyo2 ·


Ganda Marihot Simangunsong2 · Made Astawa Rai2 ·
Ridho Kresna Wattimena2

Accepted: 28 March 2022 / Published online: 19 April 2022


© The Author(s), under exclusive licence to Springer Science+Business Media, LLC, part of Springer Nature
2022

Abstract
This paper aims to evaluate dynamic slope stability induced by tunnel blast vibra-
tion with the help of a combination of geoelectrical techniques (e.g., vertical elec-
trical sounding and electrical resistivity tomography), geotechnical methods (e.g.,
boreholes, Standard Penetration Test), and laboratory testing to characterize the sub-
surface conditions of a rock slope near residential houses in the Padalarang sub-
district, West Bandung Regency, Indonesia, that is subjected to tunnel blast vibra-
tion. The slope is part of the mountain that will be excavated by a drill-and-blast
method to provide a passageway for one of the Jakarta-Bandung high-speed railway
tunnels. A series of trial blasts have been carried out and blast vibrations have been
recorded, resulting in the linear regressions of the peak vector sum (PVS) and peak
particle acceleration (PPA). Based on the allowable charge weights per delay, the
PPA regression was then used to predict the acceleration that would result from the
tunnel blast. The predicted acceleration was then used to analyze the dynamic stabil-
ity of the slope using the pseudo-static approach. The slope stability analysis shows
that the slope has a dynamic factor of safety of 1.3–1.5, indicating that the slope will
be stable after experiencing the vibration induced by the tunnel blast. The results
of this study show that combining geoelectrical survey with geotechnical methods
could help geotechnical engineers to understand the subsurface condition and its
complexity, which play vital roles in assessing the stability of the slope, particularly
the slope that is subjected to dynamic loading from a tunnel blast.

Keywords Geoelectrical · Geotechnical investigation · Tunnel blast · Blasting


vibration, Dynamic slope stability · Factor of safety · High-speed railway tunnel

* Widodo
widodo@gf.itb.ac.id
Extended author information available on the last page of the article

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Transportation Infrastructure Geotechnology (2023) 10:774–794 775

Tunnel 11
N
(Inlet)
Tunnel 11
alignment

Slope under study


Residential houses

Mt. Bohong
Tunnel 11-1

Trial blast location

Tunnel 11-1
Tunnel 11
(outlet)

7m

8m

Fig. 1  Location of the study area in the Padalarang sub-district, West Bandung Regency

1 Introduction

A high-speed train railway project that will connect Jakarta-Bandung is ongoing in


Indonesia. One of the main railway tunnels, designated as Tunnel 11, is currently
being built in the Padalarang sub-district, West Bandung Regency, West Java Prov-
ince, Indonesia. The tunnel will be excavated through a mountain called Mount
Bohong using a drill-and-blast method. A facility tunnel called Tunnel 11–1 will
also be built using a drill-and-blast method. This tunnel will serve as one of the
entrances during the construction of Tunnel 11 (Fig. 1). Because the slope of Mount
Bohong is near residential houses, the dynamic stability of the slope needs to be
evaluated in order to identify the slope stability when experiencing vibration from

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
776 Transportation Infrastructure Geotechnology (2023) 10:774–794

the tunnel blast. In this paper, the ground vibration data are collected from the blast-
ing of Tunnel 11–1 only.
To perform a slope stability analysis due to tunnel blast, it is necessary to develop
the geotechnical model of the slope in question through subsurface soil investigation
methods. Among these investigation methods, geotechnical investigations such as
boreholes or drill-cores give the best image of the subsurface but are very expen-
sive and only provide single-point estimations, while laboratory experiments give
physical and mechanical properties of rocks and soils. Another method to investigate
the subsurface is geophysical methods. Geophysical methods have been successfully
used to investigate subsurface modeling mitigation of geological structures, hazard
mitigation, near-surface mapping study, hydrological investigation, and engineering
applications (Widodo et al. 2016, 2018; Binley et al. 2015; Devi et al. 2020). Gener-
ally, geophysical methods are less invasive, less time-consuming, and less expensive
than conventional geotechnical and/or laboratory investigations. Geophysical tech-
niques particularly help to extend geotechnical investigation and interpolation of the
area. They provide large-scale investigations to characterize the subsurface by using
physical properties.
One such geophysical method is the geoelectrical or electrical resistivity method.
This method has been a subject of interest and attention to solve engineering prob-
lems (Giocoli et al. 2019; Kneisel et al. 2014; Lech et al. 2020; Li et al. 2017; Met-
waly and AlFouzan 2013; Pazzi et al. 2018; Udphuay et al. 2011; Zakaria et al.
2021). Rapid development of instrumentation, data inversion, and advances in com-
puter technology have made geoelectrical methods popular in subsurface investi-
gation (Loke et al. 2013), and they are now well-established for use in investigat-
ing/studying landslide on an unstable slope (Jongmans and Garambois 2007). In
particular, the electrical resistivity tomography (ERT) technique has been widely
applied to define the geometry of the slope, locate possible sliding surfaces, identify
material thickness, and detect areas with high water content (Perrone et al. 2014).
By integrating geophysical and geotechnical borehole data, it is possible to calibrate
a geophysical model to extend the subsurface volume of interest (Fressard et al.
2016; Hack 2000; Perrone et al. 2008; Rezaei et al. 2019; Romero-Ruiz et al. 2018;
Rusydy et al. 2021).
Correlation between ERT and geotechnical studies becomes essential to inves-
tigate subsurface geological structures. Combinations of electrical parameters and
geotechnical tests, such as the Standard Penetration Test (SPT), have been studied
in many cases (Sudha et al. 2009; Abidin et al. 2014; Gonçalves et al. 2021). For
example, according to the study by Sudha et al. (2009) in the Uttarakhand Himalaya
region, there is a linear correlation between the SPT and resistivity (ρ). According to
Gonçalves et al. (2021), the advantage of applying the combination of ERT and SPT
is that the resistivity tomography method showed great potential for extrapolating
data from SPT surveys.
It is also possible to correlate ground quality and/or strength with electrical resis-
tivity indicated by the distribution of electrical current in the ground (Hasan et al.
2021). The parameters which control soil strength, as well as electrical resistivity,
are grain size distribution, degree of saturation, porosity, and cementation. Besides
that, clay content in the soil matrix can affect both soil strength and resistivity to

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Transportation Infrastructure Geotechnology (2023) 10:774–794 777

Fig. 2  a Method of plotting midpoint in VES with Schlumberger configuration. b Method of plotting
dipole–dipole apparent resistivity data in ERT pseudo-section. n represents the relative spacing between
the current and potential dipoles

different degrees. The ion exchange property of clay forms a mobile cloud of addi-
tional ions around each clay particle. These ions facilitate the easy flow of electri-
cal current. Thus, in fine-grained soils such as clay, electrical resistivity is always
lower than expected on the basis of chemical analysis of water extracted from the
soil (Zhdanov and Keller 1994).
This paper aims to evaluate the dynamic slope stability induced by tunnel blast
vibration using a combination of geoelectrical and geotechnical methods to charac-
terize the subsurface condition. The ERT model was generated using the resistivity
inversion method. To achieve a representative geotechnical model, the ERT model
was then correlated with geotechnical boreholes and local geology data.

2 Methods of Investigation

2.1 Electrical Resistivity Investigations

The main purpose of the geophysical investigation is to learn the clear distribu-
tion of local subsurface structures including the top of the basement in the research
area. For this purpose, vertical electrical sounding (VES) and ERT techniques were
applied. VES is a 1-dimensional (1D) technique that measures vertical resistivity
variation only, assuming there are no horizontal variations of resistivity or the sub-
surface is a horizontally layered medium. It uses the same midpoint of an electrode
array as its position of measurement. For a Schlumberger array, the larger the current
electrode spacings, the deeper the current can penetrate into the subsurface; informa-
tion for the deeper subsurface can be obtained as illustrated in Fig. 2a. The potential
electrodes remain in fixed points and are only spaced more widely if the measured
voltage falls to a very low value compared to its neighborhood. There are numerous
studies on improving the 1D resistivity technique (Zohdy 1989; Gupta et al. 1997;
Ekinci and Demirci 2008; Raflesia and Widodo 2021). In the ERT technique, the
whole set of electrodes shifts laterally each time by one electrode separation after
measurement. Once the entire array has been measured, the electrode separation is

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
778 Transportation Infrastructure Geotechnology (2023) 10:774–794

Table 1  Geometric parameters Geometric parameter Profile lines


of dipole–dipole configuration
for ERT survey A-A’ B-B’ C–C’ D-D’

Number of electrodes 48 48 48 48
Electrode spacing (m) 7.5 7.5 10 10
Length of profile (m) 332 332 438 438
Max n level 5 5 5 5

increased by increments of one electrode and the process is repeated until the appro-
priate number of levels is completed. We used a dipole–dipole configuration for our
ERT survey. In this configuration, the current electrodes are a meters apart from the
potential electrode. The distance between the nearest current and potential electrode
is n times the basic distance a as illustrated in Fig. 2b. The apparent resistivity value
is plotted at the intersection of two converging lines at 45° drawn from the centers of
the current and potential electrodes (Reynolds 2011; Binley and Slater 2020).
The resistivity survey was carried out with an Ares II multi-channel resistivity
meter for VES and ERT measurements. The VES measurements were carried out at
3 points by means of the Schlumberger method with minimum half current electrode
spacing of 1 m up to 100 m. The ERT measurements were carried out in 4 profiles
with an overlapping dipole–dipole configuration. Detailed ERT geometric param-
eters are given in Table 1. The electrodes were placed perpendicular to the slope to
maintain signal distribution. The locations of the VES points and ERT profiles are
given in Fig. 3.
Raw data from the VES and ERT surveys were processed by an inversion scheme.
Before data processing, “poor” data were removed to improve the inversion results.
We used MATLAB software to plot the VES curves and inverse modeling to repre-
sent the 1D subsurface. The damped least-squares inversion method with Singular
Value Decomposition technique (Ekinci and Demirci 2008; Raflesia and Widodo
2021) was used to invert the apparent resistivity data.
The ERT inversion was carried out using Res2dinv software based on the smooth-
ness-constrained least-square method (Loke and Barker 1996). The inversion was
taken into account in the topographic data due to significant elevation of the profile
survey. The difference between the observed and computed apparent resistivity val-
ues in both the VES and ERT inversions was given by Root Mean Square (RMS),
where lower RMS means better representation of the subsurface.

2.2 Geotechnical Investigations

To quickly characterize the strength of the soil layer during the subsurface soil
investigation, field Standard Penetration Tests (SPT) were carried out at the loca-
tions of 4 geotechnical boreholes (from BH-01 to BH-04). The locations of the bore-
holes are shown in Fig. 3. Each borehole was 20 m deep, and the SPT was carried
out at vertical intervals of 1.5–2 m in each borehole. The SPT was performed by

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Transportation Infrastructure Geotechnology (2023) 10:774–794 779

Fig. 3  Location of ERT profiles, boreholes, VES points, and blasting

freely dropping a 63.5-kg hammer from a height of 76 cm. The required number of
hammer blows to penetrate to a 30-cm deep soil column is referred to as the SPT N
Value, denoted as N-SPT in this paper. The higher the N value, the stiffer the mate-
rial. Laboratory tests were also performed to determine the lithological findings in
each borehole, and the lithological properties were then averaged.

2.3 Vibration and Acceleration Prediction Using Scaled Distance Analysis

A series of trial blasts were conducted in Tunnel 11–1 with the aim to evaluate the
impact of blast and vibration on the stability of the slope. Vibration monitoring
devices called Minimate were installed at a distance of 50 to 200 m to record the
vibration magnitude. The maximum charge per delay of the explosives varied from
0.6 to 1.4 kg under various delay intervals of 200–2000 ms. The Minimate devices
recorded the PPV (peak particle velocity) and PPA (peak particle acceleration) val-
ues. The maximum amplitude of the vibration vector sum is defined as the Peak
Vector Sum (PVS).

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
780 Transportation Infrastructure Geotechnology (2023) 10:774–794

Vibration prediction using scaled distance regression analysis is the most popular
and frequently used method to predict ground vibration due to blasting (Agrawal and
Mishra 2019). The concept of scaled distance combines the distance between the
source and measurement points and the maximum explosive charge per delay (Sis-
kind et al. 1980; Siskind et al. 1994).
In general, scaled distance is defined as:
R
SD = √ (1)
Qmax

where SD is the scaled distance (m/kg0.5), R is the absolute distance from the source
to the monitoring point (m), and Qmax is the maximum charge per delay (kg). The
relationship between PVS and scaled distance (Siskind et al. 1980; Siskind et al.
1994) is given by:

PVS = K(SD)b (2)


where K and b are the site-specific constants.
Similarly, the relationship between PPA and scaled distance (Dowding 1999) is
given by:
� �b
R
PPA = K √ (3)
3
Qmax

Note that the values of K and b in Eq. (3) are not the same as in Eq. (2), in which
both constants are derived from the trial blasting result using linear regression anal-
ysis. The predicted PPA will be set as the dynamic load for slope stability analysis.

2.4 Slope Stability Modeling: Factor of Safety Calculation

In slope stability analysis, the geometry of the slope’s internal structure and its sub-
soil physical properties, the water table, and the load over the slope are the most
important parts of measuring slope stability in terms of the safety factor. In this
study, a limit equilibrium method using Spencer’s method of slices was used to cal-
culate the factor of safety using Slide2 software version 9.020 (Rocscience. 2021).
The strength of material properties was assumed to follow the Mohr–Coulomb
failure criterion. Electrical resistivity and geotechnical data obtained from the site
investigation were then used as input in defining the slope geometry. The physical
parameter properties required to perform the analysis were the unit weight (kN/m3),
cohesion (kPa), and friction angle (°) for each lithology. These rock mass properties
were obtained by combining the results of laboratory tests on intact rock with the
rock mass quality of each lithology.

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Transportation Infrastructure Geotechnology (2023) 10:774–794 781

Fig. 4  Simplified lithological logs drilled with respect to elevation. Each log was drilled 20 m deep verti-
cally and the SPT was recorded at 1.5–2 m intervals

3 Results and Discussion

3.1 Results of Geotechnical and Electrical Resistivity Investigations

The lithological information and corrected N-SPT values by depth for each borehole
are plotted in Fig. 4. The results characterize this area by three lithological units:
silty clay, gravelly clay, and andesite breccia. The sample obtained from boring was
taken in for laboratory analysis and is summarized in Table 2. In general, the N-SPT
values increase with depth, especially when the boring reaches the gravelly clay and
andesite breccia formations.
The VES curves obtained from 3 sounding points generally have low RMS values
(1.3–4%), indicating good fitness between the observed and calculated data (Fig. 5).
The curves suggest three geoelectrical layers: a low resistivity layer (13 to 25 Ωm)
sandwiched between a resistive top layer (24 to 66 Ωm) and a resistive substratum
(34 to 64 Ωm). The data from the available borehole located near the sounding point
show that the top layer is associated with silty clay, the middle layer is associated
with gravelly clay, and the substratum layer is related to andesite breccia as revealed
from BH-01 and BH-04 (Fig. 6a and b). VES-03 does not have a borehole nearby to
correlate with, but based on the geological background in this area, it has the same
stratigraphic level as all the boreholes with different thickness (Fig. 6c). These VES
results prove the ability of the electrical technique to distinguish lithological units in
the subsurface.

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
782

13
Table 2  The average values of the physical properties from the laboratory tests for each lithology
Parameter Unit Silty clay Gravelly clay Andesite breccia

SPT-N value Blow 19 53 > 100


Density kg/m3 1668 1668 1973
Unconfined compressive strength MPa 0.08 0.35 32
Young’s modulus MPa 4.4 50 4924
Poisson’s ratio –- 0.3 0.3 0.22
Cohesion MPa 0.07 0.1 9.2
Friction angle ° 13 30 37
Tensile strength MPa 0 0 2.7
Bulk modulus MPa 4 42 2885
Shear modulus MPa 2 19 2025
General description of soil or rock material - Brown color, stiff to very Mixture of clay and gravel of weathered Gray color,
stiff andesite breccia, hard soil slightly weath-
ered, hard
Transportation Infrastructure Geotechnology (2023) 10:774–794

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Transportation Infrastructure Geotechnology (2023) 10:774–794 783

Fig. 5  Fit between measured and calculated data at a VES-01, b VES-02, and c VES-03

We verified the ERT results by integrating them with geotechnical evidence in


order to interpret the resistivity values that corresponded to each formation. The
A-A’ profile trended in the NW–SE direction and was located at the top of a farm-
ing field, an open area on top of the slope. The result of the A-A’ profile is shown in
Fig. 7a. The inversion result revealed high resistivity contrast at shallow and inter-
mediate depths, which is in good agreement with the BH-02 stratigraphical data.
The top of the model showed an intermediate resistivity zone (32 to 64 Ωm) indi-
cated as gravelly clay with depth up to 10 m from the surface. The second layer
has low resistivity value (< 16 Ωm) which corresponds to silty clay. The interface
between the silty clay and gravelly clay is clearly observed by their contrast in resis-
tivity value. Meanwhile, the last layer has relatively high resistivity (> 64 Ωm) and
is interpreted as andesite breccia all the way down. ERT results confirm the geoelec-
trical model given by VES in Fig. 6, a conductive layer between two resistive layers.
The andesite breccia lies under the gravelly clay material, and the interface between
these two materials is also well-defined.
Figure 7b shows the results for the B-B’ profile, which trended in the W-E direc-
tion. The BH-01 borehole in this profile is also in good agreement with the resistiv-
ity contrast in this ERT inversion model. However, there are very different resistivity

Fig. 6  Lithological correlation of resistivity at a BH-04 obtained from VES-01, b BH-01 obtained from
VES-02, and c geoelectrical model obtained from VES-03

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
784 Transportation Infrastructure Geotechnology (2023) 10:774–794

(a)

(b)

(c)

(d)

Fig. 7  Inverted resistivity model for dipole–dipole configuration for a A-A’ profile, b B-B’ profile, c
C–C’ profile, and d D-D’ profile. VES and boreholes are also depicted in each line for comparison

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Transportation Infrastructure Geotechnology (2023) 10:774–794 785

Table 3  Correlation between Layer Resistivity (Ωm) Lithology


ERT and geotechnical borehole
data based on the range of 1 1–32 Silty clay
resistivity values
2 33–64 Gravelly clay
3 > 64 Andesite breccia

values in this ERT compared to the VES geoelectrical model (Fig. 6). In this ERT,
silty clay is characterized by low resistivity (< 16 Ωm), gravelly clay by moderate
resistivity (16–32 Ωm), and andesite breccia by higher resistivity (> 32 Ωm). This
indicates that this area has a complex geology such that assuming the subsurface is a
horizontally layered medium could lead to misinterpretation of the area. The reverse
resistivity of silty clay and gravelly clay compared to the A-A’ ERT profile could be
explained by the fact that the A-A’ profile was located in a farming field where the
silty clay has been tilled, thus increasing its resistivity, while low-resistivity gravelly
clay has probably been affected by irrigation. Similar findings have also been found
in (Rusydy et al. 2021).
The result of the C–C’ profile (Fig. 7c), which trended in the N-S direction,
revealed a wide variety of resistivity in both the vertical and lateral boundaries,
probably due to high noise since this profile was located in a forested area. The
inversion result in this profile is generally not in good agreement with the BH-03
and BH-04 boreholes. However, the BH-02 borehole fits well within the resistiv-
ity model as it is in the A-A’ profile. The silty clay formation in this profile shows
high resistivity in the southern part, low resistivity in the central zone, and moderate
resistivity in the northern area. The low and moderate resistivity of silty clay could
be explained the same way as for the silty clay in the B-B’ and A-A’ profiles, respec-
tively. No clear boundary could be identified for the formation in this profile.
Figure 7d also trended in the N-S direction and was in a forested area. The resis-
tivity image shows quite similar results with the C–C’ profile in terms of wide resis-
tivity variation. The high resistivity (> 64 Ωm) at 200–240 m distance is consistent
with the first layer in the VES-03 top layer (Fig. 6c). High resistivity (> 128 Ωm)
was also observed at 310–330 m, which was expected to be a dry zone. Unfortu-
nately, there was no borehole available in this area for interpretation since the rail-
way tunnel was originally planned to be built almost parallel to this profile. Overall,
the ERT results produced low RMS value of less than 6%, indicating that the calcu-
lated data fit with the measured data.
Having integrated the ERT results in Fig. 7 with the four geotechnical boreholes,
it is now possible to make a lithological correlation for each slope based on the
range of resistivity values. Table 3 presents the results of such lithological correla-
tion, in which layers having resistivity of 1–32 Ωm, 32–64 Ωm, and > 64 Ωm are
defined as silty clay, gravelly clay, and andesite breccia, respectively. Geotechnical
models can now be readily generated based on this lithological correlation, as shown
in Fig. 8. The geotechnical models correspond to the A-A’ slope and the B-B’ slope,
whose stabilities are the most critical because their sliding direction is towards the
residential houses (see again Fig. 3).

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
786 Transportation Infrastructure Geotechnology (2023) 10:774–794

(a)

(a)
(b)

(b)

Fig. 8  Geotechnical model for a A-A’ slope and b B-B’ slope. The model is based on the combined use
of geotechnical investigations and ERTs

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Transportation Infrastructure Geotechnology (2023) 10:774–794 787

Fig. 9  Regression of a PVS and b PPA based on recorded data during trial blasting of Tunnel 11–1

3.2 Predicted Vibration and Acceleration

The shortest distance from the blasting point to the residential area was 200 m in the
B-B’ slope and 500 m in the A-A’ slope. Therefore, we only calculated the scaled
distance analysis for the A-A’ and B-B’ slopes. The observed vibration data from
the trial blast and their predicted PVS and PPA are shown in Fig. 9. The PVS val-
ues were taken into consideration to determine whether these values were within
the threshold, which was set at PVS = 2 mm/s (according to the Indonesian National
Standard SNI 7571:2010).

3.3 Slope Stability Analysis

For this analysis, 3 tunnel blast designs for Tunnel 11–1 were tested with the maxi-
mum charge per delay set as 10 kg, 3 kg, and 1 kg. For each design, we calculated
the vibration and acceleration in terms of PVS and PPA based on the scaled dis-
tance regression analysis derived from the trial blasting (see Fig. 9). The predicted
PVS and PPA for a distance of 200 m and 500 m from the blasting location are
given in Table 4. As can be seen in Table 4, there are no PVS values greater than
PVS = 2 mm/s. Hence, the blast designs for Tunnel 11–1 comply with the Indone-
sian National Standard SNI 7571:2010.
The predicted PPA value for each design will be set as the seismic load for the
slope stability analysis for the A-A’ and B-B’ slopes. The geotechnical models for
lines A-A’ (at 500 m from blasting location) and B-B’ (at 200 m from blasting loca-
tion) are given in Fig. 9. These geotechnical models were used to calculate the factor
of safety (FoS) based on the related geotechnical data and geometry obtained from
the ERT result. For the slope stability analysis, we make justifications as follows:

• For the A-A’ and B-B’ slopes, the water table is 5 m below the surface parallel to
the slope. We assumed that this depth is already conservative in anticipation of
the worst conditions due to heavy rainfall.

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
788

13
Table 4  Predicted PVS and PPA based on the linear regression
Slope Distance from Tun- PVS (mm/s) PPA (g)
nel 11–1 Design 1 (10 kg) Design 2 (3 kg) Design 3 (1 kg) Design 1 (10 kg) Design 2 (3 kg) Design 3 (1 kg)

A-A’ 500 m 0.31 0.11 0.04 0.0036 0.0014 0.0006


B-B’ 200 m 1.49 0.53 0.21 0.0319 0.0122 0.0051
Transportation Infrastructure Geotechnology (2023) 10:774–794

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Transportation Infrastructure Geotechnology (2023) 10:774–794 789

(a)
Dynamic
Static
Design 1 (Qmax = 10 kg),
FoS = 1.56
PPA = 0.0036g
FoS = 1.53

Dynamic Dynamic
Design 2 (Qmax = 3 kg), Design 3 (Qmax = 1 kg),
PPA = 0.0014g PPA = 0.0006g
FoS = 1.55 FoS = 1.55

(b)

(b)
Static Dynamic
FoS = 1.42 Design 1 (Qmax = 10 kg),
PPA = 0.0319g
FoS = 1.29

Dynamic Dynamic
Design 2 (Qmax = 3 kg), Design 3 (Qmax = 1 kg),
PPA = 0.0122g PPA = 0.0051g
FoS = 1.37 FoS = 1.39

Fig. 10  Results of static and dynamic stability analysis for a A-A’ and b B-B’ slopes

• The materials above the sliding surface (silty clay and gravelly clay) are low-
strength with 50% cohesion. The silty clay material in the top layer will be
assumed to have cohesion of 35 kPa and the gravelly clay material in the second
layer will be assumed to have cohesion of 50 kPa.

The limit equilibrium analysis was performed under two conditions: static and
dynamic. The static calculation used only the available strength in the geotechnical
model, while the dynamic calculation had the seismic load caused by the accelera-
tion from the tunnel blast. The results are given in Fig. 10. The critical slip surface
generated by Slide 2 fits well with our interface obtained from the ERT inversion

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
790 Transportation Infrastructure Geotechnology (2023) 10:774–794

model (Fig. 7a and b). A static FoS = 1.56 is obtained for the A-A’ slope and a static
FoS = 1.42 is obtained for the B-B’ slope. This result is larger than the minimum
required FoS under the static condition as stated by the Indonesian National Stand-
ard SNI 8460:2017, that is, static FoS = 1.3. The slope stability analysis results are
summarized in Table 5.
As for dynamic FoS, it can also be seen from Table 4 that all dynamic
FoS = 1.3–1.5, which are larger than the minimum required FoS under the dynamic
condition as stated by the Indonesian National Standard SNI 8460:2017, that is,
dynamic FoS = 1.1. This result indicates that by using a maximum explosive charge
per delay of 1 kg, 3 kg, and 10 kg, Tunnel 11–1 can be excavated by a drill-and-blast
method without causing slope failure. In addition, because the dynamic FoS in this
paper adopts a pseudo-static approach that uses PPA, its value will be more con-
servative than that calculated using a full dynamic load as performed in (Prassetyo
et al. 2020).

4 Conclusions

This study evaluated the dynamic slope stability induced by blasting in the Jakarta-
Bandung railway tunnel in West Java using a combination of geoelectrical and geo-
technical methods. Data from four boreholes as well as N-SPT value and laboratory
testing, three VES points, and four profiles of ERT were used to obtain the subsur-
face characteristics of the investigated area. Trial blasting was also carried out to
obtain the geological constant factor by varying the distances and explosive weights
in the study area. Three blasting designs were evaluated by scaled distance analysis
to predict the vibration and acceleration experienced by the slope.
Integrating the electrical resistivity and geotechnical results allows us to develop
a geotechnical model of the subsurface with less ambiguity. We also set some pre-
cautionary conditions for the geotechnical model. Static and dynamic slope stability
calculations were carried out in this geotechnical model by inputting the predicted
seismic load that will be generated by each of the blasting designs. Our slope stabil-
ity calculation indicates that all dynamic FoS = 1.3–1.5, which are larger than the
minimum required FoS under the dynamic condition as stated by the Indonesian
National Standard SNI 8460:2017, that is, dynamic FoS = 1.1. This result indicates
that by using a maximum explosive charge per delay of 1 kg, 3 kg, and 10 kg, Tun-
nel 11–1 can be excavated by a drill-and-blast method without causing slope failure.
Moreover, our vibration prediction also shows that there are no PVS values greater
than PVS = 2 mm/s. Hence, the blast designs comply with the Indonesian National
Standard SNI 7571:2010.
The results of this study show that combining geoelectrical survey with geotechnical
methods could help geotechnical engineers to understand the subsurface condition and
its complexity, which play vital roles in assessing the stability of the slope, particularly
the slope that is subjected to dynamic loading from a tunnel blast.

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Table 5  Summary of slope stability analysis (calculated using the Spencer method). PPA indicates the seismic load induced by tunnel blast
Slope Distance from Static FoS PPA (g) Dynamic FoS
Tunnel 11–1 Design 1 (10 kg) Design 2 (3 kg) Design 3 (1 kg) Design 1 (10 kg) Design 2 (3 kg) Design 3 (1 kg)

A-A’ 500 m 1.56 0.0036 0.0014 0.0006 1.53 1.55 1.55


Transportation Infrastructure Geotechnology (2023) 10:774–794

B-B’ 200 m 1.42 0.0319 0.0122 0.0051 1.29 1.37 1.39

13
791

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


792 Transportation Infrastructure Geotechnology (2023) 10:774–794

Abbreviations ERT: Electrical resistivity tomography; VES: Vertical electrical sounding; SPT: Standard
Penetration Test; N-SPT value: Number of blows to penetrate 30 cm deep into a soil column; SD: Scaled
distance; R: Absolute distance from the source to the monitoring point; Qmax: Maximum charge per
delay; PPV: Peak particle velocity; PPA: Peak particle acceleration; PVS: Peak vector sum; FoS: Factor
of safety; K, b: Site-specific constants

Acknowledgements We would like to thank the Ministry of Research, Technology, and Higher Educa-
tion (RISTEKDIKTI) of Indonesia. The authors also gratefully acknowledge a research grant from the
Program of Research, Community Service, and Innovation of the Institute Technology of Bandung (ITB).
We also would like to thank PT. KCIC, HSRCC, and CREC for their assistance during data acquisition
for this research.

Author Contribution Writing-original draft preparation, writing review and editing, formal analysis, soft-
ware, visualization, and investigation (Hermawan Phanjaya and Widodo); conceptualization, methodol-
ogy, resources, and investigation (Hermawan Phanjaya, Widodo, Simon Heru Prasetyo, and Ganda Mari-
hot Simangunsong); funding acquisition (Widodo and Simon Heru Prasetyo); supervision and project
administration (Widodo, Simon Heru Prasetyo, Ganda Marihot Simangunsong, Made Astawa Rai, and
Ridho Kresna Wattimena). All authors have read and agreed to the published version of the manuscript.

Funding This research was funded by the Ministry of Research, Technology, and Higher Education (RIS-
TEKDIKTI) of Indonesia. The authors also gratefully acknowledge a research grant from the Program of
Research, Community Service, and Innovation of the Institute Technology of Bandung (P3MI-ITB) fiscal
year 2020 that is granted to the ITB Mining Engineering Research Group.

Data Availability The measurement and analysis of geophysics data were done by Hermawan Phanjaya
and Widodo, while the analysis and material of geotechnical data were supported by Simon Heru Prase-
tyo, Ganda Marihot Simangunsong, Made Astawa Rai, and Ridho Kresna Wattimena.

Code Availability The 1D models of geoelectric data performed by own software (RESEP BANYU) and
the 2D models of geoelectric data have proceed by Res2Dinv Software (Loke and Barker, 1996). Rocsci-
ence. Slide2-2D Limit Equilibrium Analysis for Slopes [Version 9.020]. Toronto: Rocscience Inc.; 2021.

Declarations

Conflict of Interest The authors declare no competing interests.

References
Abidin, M.H.Z., Saad, R., Ahmad, F., Wijeyesekera, D.C., Baharuddin, M.F.T.: Correlation analysis between
field electrical resistivity value (ERV) and basic geotechnical properties (BGP). Soil Mech. Found. Eng.
51, 117–125 (2014). https://​doi.​org/​10.​1007/​s11204-​014-​9264-x
Agrawal, H., Mishra, A.K.: Modified scaled distance regression analysis approach for prediction of blast-
induced ground vibration in multi-hole blasting. J. Rock Mech. Geotech. Eng. 11, 202–207 (2019).
https://​doi.​org/​10.​1016/j.​jrmge.​2018.​07.​004
Binley, A., Slater, L.: Resistivity and induced polarization: theory and applications to the near-surface Earth.
Cambridge University Press, Cambridge (2020)
Binley, A., Hubbard, S.S., Huisman, J.A., Revil, A., Robinson, D.A., Singha, K., et al.: The emergence
of hydrogeophysics for improved understanding of subsurface processes over multiple scales. Water
Resour. Res. 51, 3837–3866 (2015). https://​doi.​org/​10.​1002/​2015W​R0170​16
Devi, A., Israil, M., Singh, A., Gupta, P.K., Yogeshwar, P., Tezkan, B.: Imaging of groundwater contamina-
tion using 3D joint inversion of electrical resistivity tomography and radio magnetotelluric data: a case
study from Northern India. Near Surf. Geophys. 18, 261–274 (2020). https://​doi.​org/​10.​1002/​nsg.​12092
Dowding, C.H.: Blast vibration monitoring and control: a thirty year perspective. 37th U.S. Symp. Rock
Mech., Vail, CO: Taylor & Francis; p. ARMA-99–0045 (1999)
Ekinci, Y.L., Demirci, A.: A damped least-squares inversion program for the interpretation of Schlumberger
sounding curves. J. Appl. Sci. 8, 4070–4078 (2008). https://​doi.​org/​10.​3923/​jas.​2008.​4070.​4078

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Transportation Infrastructure Geotechnology (2023) 10:774–794 793

Fressard, M., Maquaire, O., Thiery, Y., Davidson, R., Lissak, C.: Multi-method characterisation of an active
landslide: case study in the Pays d’Auge plateau (Normandy, France). Geomorphology 270, 22–39
(2016). https://​doi.​org/​10.​1016/j.​geomo​rph.​2016.​07.​001
Giocoli, A., Hailemikael, S., Bellanova, J., Calamita, G., Perrone, A., Piscitelli, S.: Site and building charac-
terization of the Orvieto Cathedral (Umbria, Central Italy) by electrical resistivity tomography and sin-
gle-station ambient vibration measurements. Eng. Geol. 260, 105195 (2019). https://​doi.​org/​10.​1016/j.​
enggeo.​2019.​105195
Gonçalves, J.T.D., Botelho, M.A.B., Machado, S.L., Netto, L.G.: Correlation between field electrical resistiv-
ity and geotechnical SPT blow counts at tropical soils in Brazil. Environ. Challenges 5, 100220 (2021).
https://​doi.​org/​10.​1016/j.​envc.​2021.​100220
Gupta, P.K., Niwas, S., Gaur, V.K.: Straightforward inversion of vertical electrical sounding data. Geophysics
62, 775–785 (1997). https://​doi.​org/​10.​1190/1.​14441​87
Hack, R.: Geophysics for slope stability. Surv. Geophys. 21, 423–448 (2000). https://​doi.​org/​10.​1023/A:​
10067​97126​800
Hasan, M., Shang, Y., Meng, H., Shao, P., Yi, X.: Application of electrical resistivity tomography (ERT) for
rock mass quality evaluation. Sci. Rep. 11, 23683 (2021). https://​doi.​org/​10.​1038/​s41598-​021-​03217-8
Jongmans, D., Garambois, S.: Geophysical investigation of landslides: a review. Bull. La Société Géologique
Fr. 178, 101–112 (2007). https://​doi.​org/​10.​2113/​gssgf​bull.​178.2.​101
Kneisel, C., Emmert, A., Kästl, J.: Application of 3D electrical resistivity imaging for mapping frozen
ground conditions exemplified by three case studies. Geomorphology 210, 71–82 (2014). https://​doi.​
org/​10.​1016/j.​geomo​rph.​2013.​12.​022
Lech, M., Skutnik, Z., Bajda, M., Markowska-Lech, K.: Applications of electrical resistivity surveys in solv-
ing selected geotechnical and environmental problems. Appl. Sci. 10, 2263 (2020). https://​doi.​org/​10.​
3390/​app10​072263
Li, S., Liu, B., Xu, X., Nie, L., Liu, Z., Song, J., et al.: An overview of ahead geological prospecting in tun-
neling. Tunn. Undergr. Sp. Technol. 63, 69–94 (2017). https://​doi.​org/​10.​1016/j.​tust.​2016.​12.​011
Loke, M.H., Barker, R.D.: Rapid least-squares inversion of apparent resistivity pseudosections by a quasi-
Newton method1. Geophys. Prospect. 44, 131–152 (1996). https://doi.org/10.1111/j.1365-2478.1996.
tb00142.x
Loke, M.H., Chambers, J.E., Rucker, D.F., Kuras, O., Wilkinson, P.B.: Recent developments in the direct-
current geoelectrical imaging method. J Appl Geophys 95, 135–156 (2013). https://​doi.​org/​10.​1016/j.​
jappg​eo.​2013.​02.​017
Metwaly, M., AlFouzan, F.: Application of 2-D geoelectrical resistivity tomography for subsurface cavity
detection in the eastern part of Saudi Arabia. Geosci. Front. 4, 469–476 (2013). https://​doi.​org/​10.​
1016/j.​gsf.​2012.​12.​005
Pazzi, V., Di Filippo, M., Di Nezza, M., Carlà, T., Bardi, F., Marini, F., et al.: Integrated geophysical survey
in a sinkhole-prone area: Microgravity, electrical resistivity tomographies, and seismic noise measure-
ments to delimit its extension. Eng. Geol. 243, 282–293 (2018). https://​doi.​org/​10.​1016/j.​enggeo.​2018.​
07.​016
Perrone, A., Vassallo, R., Lapenna, V., Di, M.C.: Pore water pressures and slope stability: a joint geophysical
and geotechnical analysis. J. Geophys. Eng. 5, 323–337 (2008). https://​doi.​org/​10.​1088/​1742-​2132/5/​3/​
008
Perrone, A., Lapenna, V., Piscitelli, S.: Electrical resistivity tomography technique for landslide investigation:
a review. Earth-Sci. Rev. 135, 65–82 (2014). https://​doi.​org/​10.​1016/j.​earsc​irev.​2014.​04.​002
Prassetyo, S.H., Simangunsong, G.M., Sadikin, A.: Dynamic stability of mine slopes due to bench blasting.
In: da Fontoura SAB, Rocca RJ, Mendoza JFP, editors. Rock Mech. Nat. Resour. Infrastruct. Dev., Lon-
don: CRC Press; p. 3506–13 (2020) https://​doi.​org/​10.​1201/​97803​67823​177.
Raflesia, F., Widodo, W.: Flower pollination algorithm for the inversion of Schlumberger sounding curve.
IOP Conf. Ser. Earth Environ. Sci. 873, 012018 (2021). https://​doi.​org/​10.​1088/​1755-​1315/​873/1/​
012018
Reynolds, J.M.: An introduction to applied and environmental geophysics, 2nd edn. Wiley-Blackwell, Chich-
ester (2011)
Rezaei, S., Shooshpasha, I., Rezaei, H.: Reconstruction of landslide model from ERT, geotechnical, and field
data, Nargeschal landslide, Iran. Bull. Eng. Geol. Environ. 78, 3223–3237 (2019). https://​doi.​org/​10.​
1007/​s10064-​018-​1352-0
Rocscience.: Slide2–2D limit equilibrium analysis for slopes [version 9.020]. Toronto: Rocscience Inc.
(2021)
Romero-Ruiz, A., Linde, N., Keller, T., Or, D.: A review of geophysical methods for soil structure characteri-
zation. Rev. Geophys. 56, 672–697 (2018). https://​doi.​org/​10.​1029/​2018R​G0006​11

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
794 Transportation Infrastructure Geotechnology (2023) 10:774–794

Rusydy, I., Fathani, T.F., Al-Huda, N., Sugiarto, Iqbal, K., Jamaluddin, K., et al.: Integrated approach in
studying rock and soil slope stability in a tropical and active tectonic country. Environ. Earth Sci. 80, 58
(2021). https://​doi.​org/​10.​1007/​s12665-​020-​09357-w
Siskind, D.E., Stagg, M.S., Kopp, J.W., Dowding, C.H.: Structure response and damage produced by ground
vibration from surface mine blasting. Twin Cities (1980)
Siskind, D.E., Stagg, M.S., Wiegand, J.E., Schulz, D.L.: Surface mine blasting near pressurized transmission
pipelines. Report of investigations. Twin Cities (1994)
Sudha, K., Israil, M., Mittal, S., Rai, J.: Soil characterization using electrical resistivity tomography and geo-
technical investigations. J. Appl. Geophys. 67, 74–79 (2009). https://​doi.​org/​10.​1016/j.​jappg​eo.​2008.​09.​
012
Udphuay, S., Günther, T., Everett, M.E., Warden, R.R., Briaud, J.-L.: Three-dimensional resistiv-
ity tomography in extreme coastal terrain amidst dense cultural signals: application to cliff sta-
bility assessment at the historic D-Day site. Geophys. J. Int. 185, 201–220 (2011). https://doi.
org/10.1111/j.1365-246X.2010.04915.x
Widodo, Gurk, M., Tezkan, B.: Multi-dimensional interpretation of radiomagnetotelluric and transient elec-
tromagnetic data to study active faults in the Mygdonian Basin, Northern Greece. J. Environ. Eng. Geo-
phys. 21, 121–33 (2016). https://​doi.​org/​10.​2113/​JEEG21.​3.​121
Widodo, W., Azimmah, A., Santoso, D.: Exploring the Japan Cave in Taman Hutan Raya Djuanda, Bandung
using GPR. J. Environ. Eng. Geophys. 23, 377–381 (2018). https://​doi.​org/​10.​2113/​JEEG23.​3.​377
Zakaria, M.T., MohdMuztaza, N., Zabidi, H., Salleh, A.N., Mahmud, N., Samsudin, N., et al.: 2-D cross-
plot model analysis using integrated geophysical methods for landslides assessment. Appl. Sci. 11, 747
(2021). https://​doi.​org/​10.​3390/​app11​020747
Zhdanov, M., Keller, G.: The geoelectrical methods in geophysical exploration. Elsevier, Amsterdam (1994)
Zohdy, A.A.R.: A new method for the automatic interpretation of Schlumberger and Wenner sounding
curves. Geophysics 54, 245–253 (1989). https://​doi.​org/​10.​1190/1.​14426​48

Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps
and institutional affiliations.

Authors and Affiliations

Widodo1 · Hermawan Phanjaya1 · Simon Heru Prassetyo2 ·


Ganda Marihot Simangunsong2 · Made Astawa Rai2 ·
Ridho Kresna Wattimena2
Hermawan Phanjaya
hermawan5751@gmail.com
Simon Heru Prassetyo
simon@mining.itb.ac.id
Ganda Marihot Simangunsong
ganda@mining.itb.ac.id
Made Astawa Rai
maderai@mining.itb.ac.id
Ridho Kresna Wattimena
rkw@mining.itb.ac.id
1
Geophysical Engineering, Faculty of Mining and Petroleum Engineering, Institute
of Technology Bandung, Bandung 40132, Indonesia
2
Mining Engineering, Faculty of Mining and Petroleum Engineering, Institute of Technology
Bandung, Bandung 40132, Indonesia

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center
GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers
and authorised users (“Users”), for small-scale personal, non-commercial use provided that all
copyright, trade and service marks and other proprietary notices are maintained. By accessing,
sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of
use (“Terms”). For these purposes, Springer Nature considers academic use (by researchers and
students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and
conditions, a relevant site licence or a personal subscription. These Terms will prevail over any
conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription (to
the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of
the Creative Commons license used will apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may
also use these personal data internally within ResearchGate and Springer Nature and as agreed share
it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not otherwise
disclose your personal data outside the ResearchGate or the Springer Nature group of companies
unless we have your permission as detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial
use, it is important to note that Users may not:

1. use such content for the purpose of providing other users with access on a regular or large scale
basis or as a means to circumvent access control;
2. use such content where to do so would be considered a criminal or statutory offence in any
jurisdiction, or gives rise to civil liability, or is otherwise unlawful;
3. falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association
unless explicitly agreed to by Springer Nature in writing;
4. use bots or other automated methods to access the content or redirect messages
5. override any security feature or exclusionary protocol; or
6. share the content in order to create substitute for Springer Nature products or services or a
systematic database of Springer Nature journal content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a
product or service that creates revenue, royalties, rent or income from our content or its inclusion as
part of a paid for service or for other commercial gain. Springer Nature journal content cannot be
used for inter-library loans and librarians may not upload Springer Nature journal content on a large
scale into their, or any other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not
obligated to publish any information or content on this website and may remove it or features or
functionality at our sole discretion, at any time with or without notice. Springer Nature may revoke
this licence to you at any time and remove access to any copies of the Springer Nature journal content
which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or
guarantees to Users, either express or implied with respect to the Springer nature journal content and
all parties disclaim and waive any implied warranties or warranties imposed by law, including
merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published
by Springer Nature that may be licensed from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a
regular basis or in any other manner not expressly permitted by these Terms, please contact Springer
Nature at

onlineservice@springernature.com

You might also like