You are on page 1of 20

HOSTED BY Available online at www.sciencedirect.

com

ScienceDirect
Soils and Foundations 61 (2021) 198–217
www.elsevier.com/locate/sandf

Technical Paper

Early warning system for rainfall- and snowmelt-induced slope failure


in seasonally cold regions
Yulong Zhu a, Tatsuya Ishikawa b,⇑, Srikrishnan Siva Subramanian c, Bin Luo d
a
Laboratory of Analytical Geomechanics, Graduate School of Engineering, Hokkaido University, Kita 13, Nishi 8, Kita-ku, Sapporo, Hokkaido
060-8628, Japan
b
Faculty of Engineering, Hokkaido University, Japan
c
State Key Laboratory of Geohazard Prevention and Geo-environment Protection, Chengdu University of Technology, China
d
Department of Road and Bridge Engineering, School of Civil Engineering, Sichuan Agricultural University, China

Received 9 January 2020; received in revised form 15 July 2020; accepted 23 November 2020
Available online 19 January 2021

Abstract

In 2005, the Japanese government launched a new nationwide early warning system for predicting debris flow and slope failure dis-
asters based on rainfall intensity and the Soil Water Index (SWI). However, the Japanese government has not set early warning criteria in
many mountain areas. In addition, the existing early warning criteria in some areas are much higher than realistic ones, and snowmelt
water is not considered in the calculation of the SWI. These two factors have been the cause of many slope failures in seasonally cold
regions, induced by rainfall and/or snowmelt, which were not predicted. Therefore, this study attempts to propose a new determination
method for setting early warning criteria for rainfall- and/or snowmelt-induced slope failures in seasonally cold regions. For this purpose,
the study firstly proposes a combination model for estimating snow density that incorporates the hourly snowmelt water into the Japa-
nese early warning system more accurately by using meteorological monitoring data and modeled snow density. Next, based on case
studies and parametric analyses for slope stability assessment, new early warning criteria are proposed for predicting three different pat-
terns of slope failures under two typical types of precipitation (rainfall and snowmelt) conditions. Finally, a new determination method
for setting early warning criteria in seasonally cold regions is proposed by referring to the existing early warning criteria near the target
area, in accordance with the precipitation types and the local ground conditions of the slopes. Since the existing early warning criteria
near the target area already take the effects of the variations in local geology and geography into account, the new determination method
for early warning criteria can be applied to arbitrary areas in seasonally cold regions, without directly considering the local soil prop-
erties, in the actual design and maintenance works.
Ó 2020 Production and hosting by Elsevier B.V. on behalf of The Japanese Geotechnical Society. This is an open access article under the CC BY-
NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

Keywords: Early warning system; Soil Water Index; Snowmelt water estimation; Slope stability assessment; Seasonally cold regions

1. Introduction most instances, it is the only suitable option (Glade and


Nadim, 2014), as it gives people sufficient time to relocate
The implementation of a territorial early warning system themselves after receiving early warning messages before
is a cost-effective non-structural risk mitigation measure. In debris flows or slope failures occur. In the past few decades,
various systems have been designed in many countries. For
example, the Landslip Warning System was set up in Hong
Peer review under responsibility of The Japanese Geotechnical Society. Kong, China, the Real-time Monitoring System (RTMS)
⇑ Corresponding author.
was installed in Malaysia, and the national early warning
E-mail addresses: zhuyulong@eis.hokudai.ac.jp (Y. Zhu), t-ishika@
eng.hokudai.ac.jp (T. Ishikawa), srikrishnan@frontier.hokudai.ac.jp
system (SANF), designed for rainfall-induced landslides,
(S. Siva Subramanian), bin-luo@sicau.edu.cn (B. Luo). was established in Italy (Piciullo, 2016). In Japan,

https://doi.org/10.1016/j.sandf.2020.11.009
0038-0806/Ó 2020 Production and hosting by Elsevier B.V. on behalf of The Japanese Geotechnical Society.
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

Kuramoto et al. (2001) proposed the RBFN (Radial Basis season, this study proposes a new determination method
Function Network) method to set the non-linear Critical for setting early warning criteria in seasonally cold regions.
Line (hereinafter, CL) for sediment-related disasters in 5- At first, in order to incorporate the hourly snowmelt water
km-square grids covering all of Japan. In 2005, the Japa- into the Japanese early warning system, referring to several
nese government launched a new nationwide early warning well-known existing models and measurement data, a com-
system based on this method (Osanai et al., 2010). The sys- bination model is newly proposed for more accurately esti-
tem utilizes two rainfall indices, 60-min cumulative rainfall mating the snow density during the midwinter and early
(rainfall intensity) and the Soil Water Index (hereinafter spring periods. Next, based on case studies and parametric
referred to as the SWI), as warning information for heavy analyses, new early warning criteria are proposed for pre-
rainfall-induced sediment-related disasters. dicting three different patterns of slope failures under two
In seasonally cold regions, the major factors that influ- typical types of precipitation (rainfall and snowmelt) con-
ence soil slope stability are heavy rainfall, freeze-thaw ditions observed in Hokkaido, Japan. Finally, the new
action, and snowmelt water infiltration. Recent studies on determination method is proposed for setting early warn-
geohazards that have occurred on volcanic soil slopes, by ing criteria in arbitrary areas of seasonally cold regions
Ishikawa and Miura (2011), Ishikawa et al. (2015), and by referring to the existing early warning criteria near the
Ishikawa et al. (2016), have described how there is a differ- target area, in accordance with the precipitation types
ence in the failure mechanisms of slopes between cold and the local ground conditions of the slopes.
regions and warm-temperate regions due to additional fac-
tors, i.e., freeze-thaw action causing residual displacement
parallel to the slope surface and shear strain at the subsur- 2. Soil Water Index for seasonally cold regions
face layer. Accordingly, Ishikawa et al. (2015) divided slope
failures in cold regions into two main classes: slope failures 2.1. Soil Water Index (SWI) incorporating snowmelt water
that occur during the snow-melting season (March-April-
May) and slope failures that occur during the heavy rainfall Sugawara et al. (1974) proposed a physical runoff model
season (August-September). Moreover, in recent decades, (tank model) using multi-layered tanks to simulate surface
snowmelt water has become the major contributor to disas- water infiltration, surface runoff, and subsurface runoff. In
ters occurring during the snow-melting season. For exam- the three-layer tank model, the soil moisture is represented
ple, according to statistics on sediment-related disasters as the water depth in each layer. The sum of the water
happening along national highways in Hokkaido for a per- depths in the three layers is named the ‘‘Soil Water Index
iod of 16 years, from 1998 to 2013 (Yajima and Kurahashi, (SWI)”. The tank model has been proven capable of mod-
2015), about 40% of the slope failures occurred during the eling the hydrologic responses from a wide range of water-
snow-melting season. Among them, 41% were induced by sheds (Hossain, 1989; Chen and Adams, 2006; Hong et al.,
snowmelt, 12% were induced by snowmelt and rainfall, 2015; Chen et al., 2017; Nie et al., 2017). In analyzing the
and the remaining 47% were induced by rainfall, snow runoff and infiltration of basins, satisfactory results can
slides, etc. It is worth noting that about 37% of the slope be obtained by calibrating the model parameters. For
failures occurring in the heavy rainfall season were not pre- example, the tank model has been verified for modeling
dicted, while as many as 96% of the slope failures occurring the rainfall-runoff process of the Brue catchment in the
in the snow-melting season were not predicted (Kurahashi South West of England (Cerda-Villafana et al., 2008), the
et al., 2017). Therefore, the precipitation thresholds of the Hieu River Basin in Vietnam (Phuong et al., 2018), and
Japanese early warning criteria are much higher than the the Tandula catchment in Chhattisgarh state of India
realistic ones in some areas, and the Japan Meteorological (Jaiswal et al., 2020). However, in the Japanese early warn-
Agency (JMA) has been tentatively operating with lower ing system, the SWI only represents the conceptual ground-
thresholds of the SWI of up to 50% to 80% for its heavy water level, and the time series change in the SWI exhibits a
rain warning/advisories (JMA, 2019). On the other hand, similar trend when the parameters are adjusted for different
the existing precipitation threshold was determined only areas (Chen et al., 2017). Therefore, the Japanese govern-
by using the occurrence of rainfall-induced debris flows ment applies the SWI to all of Japan using fixed parameters
and slope failures and did not consider the long-term snow- in spite of the various geological conditions that exist
melt water infiltration. Due to these facts, the rainfall (Chen et al., 2017). The differences in geological character-
threshold is not applicable to predictions of snowmelt- istics and material properties are taken into account in the
induced soil slope failures during the snow-melting season local early warning criteria. In this study, the calculation of
(Siva Subramanian et al., 2018). Therefore, it is indispens- the SWI is based on the parameters suggested by Okada
able that new early warning criteria be proposed for pre- (2001), as presented in Fig. 1 (a).
dicting rainfall- and/or snowmelt-induced slope failures The tank model assumes that when the depth of the
in seasonally cold regions, like Hokkaido, Japan. water storage in each layer (H1, H2, and H3) exceeds the
In order to develop and improve on the existing Japa- height of the outlet of the sidewall (L11, L12, L2, and L3),
nese early warning system for predicting rainfall- and/or groundwater will outflow from the outlets by the constant
snowmelt-induced slope failures during the snow-melting coefficients (a11, a12, a2, and a3). The infiltration (I1, I2, and
199
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

Fig. 1. (a) Three-layer tank model for calculating Soil Water Index (adapted from Okada, 2001) and (b) Prediction of occurrence of slope failure based on
CL and SL.

I3) from the upper tank to the lower tank is assumed to be pensable that the hourly snowmelt water be considered in
proportional (b1, b2, and b3) to the height of the water stor- the calculation of the SWI during the snow-melting season.
age. The outflows and infiltrations are calculated based on
the following equations: 2.2. Estimation of hourly snowmelt water

a11  ðH1 ðtÞ  L11 Þ; if H1 > L11
q11 ¼ ð1Þ A basic approach is commonly used to estimate the
0; if H1  L11 amount of snowmelt water, which is based on meteorolog-

a12  ðH1 ðtÞ  L12 Þ; if H1 > L12 ical monitoring data, i.e., monitored snow depth and snow-
q12 ¼ ð2Þ fall. Siva Subramanian et al. (2018) estimated the hourly
0; if H1  L12
 snowmelt water using this approach. However, they did
a2  ðH2 ðtÞ  L2 Þ; if H2 > L2 not consider the changes in snow density during the estima-
q2 ¼ ð3Þ
0; if H2  L2 tion process. Therefore, for more precise predictions, this
 study takes into account the changes in snow density in
a3  ðH3 ðtÞ  L3 Þ; if H3 > L3
q3 ¼ ð4Þ the estimation of the hourly snowmelt water. The amount
0; if H3  L3 of hourly snowmelt water can be estimated according to
I1 ¼ b1  H1 ðtÞ; I2 ¼ b2  H2 ðtÞ; I3 ¼ b3  H3 ðtÞ ð5Þ the following relationship (Kazama et al., 2008):
After considering the snowmelt water with rainfall as SM ¼ DSWE  SF ð10Þ
the water supply to the model, the water storage in each where SWE is the snow water equivalent (mm), DSWE is
tank and the SWI are represented as the hourly change in the snow water equivalent (mm/h),
H1 ðt þ DtÞ ¼ H1 ðtÞ  q11  q12  I1 þ R þ SM ð6Þ and SF is the snowfall (mm/h).
The following relationship allows for the conversion of
H2 ðt þ DtÞ ¼ H2 ðtÞ  q2  I2 þ I1 ð7Þ
the snow depth to the SWE by the given snow density
H3 ðt þ DtÞ ¼ H3 ðtÞ  q3  I3 þ I2 ð8Þ (qn) (Mizukami and Perica, 2008):
SWIðtÞ ¼ H1 ðtÞ þ H2 ðtÞ þ H3 ðtÞ ð9Þ q
SWE ¼ hs n ð11Þ
qw
where R is the rainfall intensity (mm/h), SM is the amount
of hourly snowmelt water (mm/h), t is the current time, and where hs is the snow depth (m), qn is the snow density (kg/
Dt is the time step with the value of 60 min. In order to pre- m3), and qw is the density of the liquid water (approxi-
dict the occurrence of debris flows and slope failures, the mately 1000 kg/m3 at 0C). Grounded in the fact that the
Japanese government set a CL for the relationship between snowmelt is directly estimated from the depth of the accu-
the 60-minute cumulative rainfall and the SWI in a 5-km- mulated snow, the estimation of the snowmelt water
square grid, i.e., when the snake line (hereafter, SL, indicat- derived from the meteorological monitoring data and mod-
ing the relationship between the 60-min rainfall and the eled snow density (snow density is estimated from a snow
SWI) exceeds the CL, debris flows and slope failures can density model that considers snow depth, snow aging,
occur, as shown in Fig. 1 (b). However, the current Japa- etc.) usually has high accuracy. Therefore, this study incor-
nese early warning system does not take snowmelt water porates the hourly snowmelt water into the Japanese early
into account, resulting in the inaccurate calculation of the warning system by employing the meteorological monitor-
SWI during the snow-melting season. Therefore, it is indis- ing data and modeled snow density.
200
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

3. Snow density estimation 3.2. Existing models

3.1. Field monitoring data of snow density in northern Japan Sturm et al. (2010) developed a nonlinear approach for
estimating the corresponding snow density by the observed
The snow density is influenced by the snowmelt water, snow depth (abbreviated as the S10 model). In the model,
snow depth, snow aging (older snow is denser than younger the snow density is a function of the snow depth (hs), the
snow), and meteorological variables like air temperature, day of the year (DOY), and four general fitting parameters
wind speed, and solar radiation, etc. (Sturm et al., 2010). (qmax, q0, k1, and k2) for each of the five broad snow classes
For example, the densification of an underlying snow layer given in Table 1.
proceeds with time and depth due to compaction by the
qh;DOY ¼ ðqmax  q0 Þ½1  exp ðk1  hs  k2  DOYÞ þ q0
mechanical loading of the overlying snow. As a result,
snow density increases with time. The typical seasonal ð12Þ
snow density in Japan lies in the range of 30 to 600 kg/ where qh,DOY indicates the snow density (kg/m3) related to
m3 (Motoya et al., 2017). According to the field monitoring the snow depth and the DOY (DOY is a number from 92
data (monitoring period: February 20th–March 10th in (1st October) to + 181 (30th June); the DOY is 1 on 1st
2012, 2014, 2016, and 2017) of the snow depth and the January), qmax is the maximum snow density (kg/m3), q0
SWE in Hokkaido, Japan, referenced from Arakawa is the initial snow density (kg/m3), k1 (1/m) and k2 are
(2012), Shirakawa et al. (2014), Shirakawa (2016), and the densification parameters for depth and the DOY,
Shirakawa (2017), the snow density at the beginning of respectively, and hs indicates the snow depth (m).
March throughout Hokkaido is about 300 kg/m3. Further- Meloysund et al. (2007) reported a method derived from
more, Abe et al. (1989) measured the snow depth, SWE, the former USSR in which the snow density can be repre-
and snow density in Shinjo City, Japan for 15 years (during sented by the following expression (abbreviated as the
the winter seasons from 1973/74 to 1987/88). Motoya et al. USSR model):
(2017) measured the snow density around Akita Prefecture, pffiffiffiffiffi p ffiffiffiffi pffiffiffi
qn ¼ ð90 þ 130 hs Þð1:5 þ 0:17  TÞð1 þ 0:1 vÞ ð13Þ
3
Japan from December 20th, 2011 to April 22nd, 2012.
Shinme and Yamashita (2008) measured the snow density
from December 10th, 2006 to May 10th, 2007, and In Japan, Endo (1986) fitted the following relationship
Yajima and Kurahashi (2015) measured the snow density between the snow density and the date based on the mon-
from March 5th, 2015 to May 14th, 2015 in the Jozankei itoring data of the snow density from 1964 to 1979 in Sap-
area of Hokkaido, Japan, respectively. Fig. 2 illustrates poro, Hokkaido, Japan (abbreviated as the E86 model):
the results of the measured density of snow cover vs. time qn ¼ 239 þ 0:861  t1 þ 0:01  t21 ð14Þ
referred from Abe et al. (1989), Shinme and Yamashita
(2008), Motoya et al. (2017), and Yajima and Kurahashi Shinme and Yamashita (2008) proposed that the snow
(2015) for the chosen area. From Fig. 2, it is recognized density is a linear equation of the snow depth and time
that, in northern Japan, the snow density is about based on the monitoring data of the snow depth and snow
150 kg/m3 in the first part of winter (December-January) density from 2006 to 2007 in the Jozankei area of Hok-
and increases to 300 kg/m3 in the middle part of winter kaido (abbreviated as the SY08 model).
(February-March), then reaches about 600 kg/m3 in the qn ¼ 15:77  hs þ 0:11  t2 þ 128:1 ð15Þ
last part of winter (April-May).
where m is the wind velocity (m/s), t1 denotes DOY + 17 (t1
is 1 on 15th December), and t2 is the time in hours mea-
sured from the existing snow. Apart from this, many stud-
ies (e.g., Tomomura et al., 1982; Abe and Shimizu, 2004)
have proposed a variety of different models for estimating
the snow density based on the snow depth or snow aging.
The existing models are summarized in Table 2.
For discussing the suitability and applicability of the
above-mentioned existing models in Hokkaido, Japan,
changes in the snow density in the Jozankei area of Hok-
Table 1
Model parameters by snow class (Sturm et al., 2010).
Snow class qmax (kg/m3) q0 (kg/m3) k1 (1/m) k2
Tundra 363.0 242.5 0.0029 0.0049
Maritime 597.9 257.8 0.0010 0.0038
Prairie 594.0 233.2 0.0016 0.0031
Alpine 597.5 223.7 0.0012 0.0038
Taiga 217.0 217.0 0.0000 0.0000
Fig. 2. Field monitoring data of snow density in northern Japan.
201
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

kaido are calculated by the existing models from December


10th, 2006 to May 15th, 2007, and then compared with the
measured values of the snow density measured by Shinme
and Yamashita (2008), shown in Figs. 3 and 4. It is recog-
nized from these figures that, as the AS04 model and the
USSR model only use the snow depth as the variable, they
can only describe the changes in snow density with the
snow depth between December and March, while the
changes in snow density with the snow aging between
March and May cannot be captured. Moreover, as the
T82 and E86 models only use the cumulative time after
the snow has accumulated on the ground as the variable,
they can only describe the changes in snow density with Fig. 3. Snow density vs. snow depth estimated by existing models and field
the snow aging between March and May, while the impact monitoring data (calculation and measurement period: 12/10/2006–05/15/
of the snow depth on the snow density between December 2007).
and March is not considered. On the other hand, although
the S10 and SY08 models can consider the influences of
both the snow depth and snow aging, it is recognized that
the snow densities estimated by these two models are smal-
ler than those measured between April and May. This sug-
gests that none of these models can be used alone to
accurately estimate the snow density between December
and May in Hokkaido, Japan. Therefore, this study combi-
nes the SY08 model and the E86 model to propose a com-
bination model, named SY08-E86, that can more
accurately estimate the snow density in Hokkaido, Japan.

3.3. Combination model


Fig. 4. Snow density vs. time estimated by existing models and field
Mizukami and Perica (2008) separated the time series in monitoring data (calculation and measurement period: 12/10/2006–05/15/
two periods, midwinter (1st December-1st March) and 2007).
early spring (1st March-30th April), by comparing the
interannual variations in the densification rates in four dis- model is newly proposed for a precise calculation of the
tinct regions (i.e., coastal region, continental region, high snow density in Hokkaido, Japan.
elevation region, and region between the coastal and conti-
nental regions). They found that the increased snow depth (1) During the midwinter period (SY08 model):
and increased snow aging are the key factors influencing
the snow density during the midwinter period (1st
December-1st March. On the other hand, during the early qn ¼ 15:77  hs þ 0:11  t2 þ 128:1 ð15Þ
spring period (1st March-30th April), snowmelt water
causes faster destructive metamorphism and finally
increases the snow density. Referring to the above study,
(2) During the early spring period (E86 model):
the SY08 model has higher precision during the midwinter
period and the E86 model has higher precision during the
early spring period, as shown in Fig. 4. By combining the
advantages of these two models, an SY08-E86 combination qn ¼ 239 þ 0:861  t1 þ 0:01  t21 ð14Þ

Table 2
Summary of existing snow density estimation models.
Model name Formula Reference
S10 model qh;DOY ¼ ðqmax p ffiffiffiffiqffi 0 Þ½1  exp ðk1p
ffiffiffiffi hs  k2 pDOY Þ þ q0 Sturm et al. (2010)
ffiffiffi
USSR model qn ¼ ð90 þ 130 hs Þð1:5 þ 0:17  3 TÞð1 þ 0:1 vÞ Meloysund et al. (2007)
SY08 model qn ¼ 15:77  hs þ 0:11  t2 þ 128:1 Shinme and Yamashita (2008)
E86 model qn ¼ 239 þ 0:861  t1 þ 0:01  t21 Endo (1986)
T82 model qn ¼ 2  t1 þ 150
pffiffiffiffi
ffi Tomomura et al. (1982)
AS04 model qn ¼ 109  hs þ 177 Abe and Shimizu (2004)
Note: t1 denotes DOY and t2 is time in hours measured from existing snow.
202
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

The snow density calculated by the newly proposed and snowmelt water were found to be the major contribu-
combination model is shown in Fig. 5. It is recognized that tors to these disasters (Nakatsugawa et al., 2015). By using
the combination model provides more robust estimations meteorological monitoring data obtained from the Auto-
of the snow density. Therefore, this study proposes the matic Meteorological Data Acquisition System, named
use of the newly proposed combination model for estimat- AMeDAS, and the modeled snow density, the amount of
ing the snow density. hourly snowmelt water is calculated for Nakayama Pass
during two long-term periods, from November 8th, 2011
to May 5th, 2012 and from November 8th, 2012 to April
4. Prediction method for rainfall- and/or snowmelt-induced 8th, 2013. As discussed in the previous chapter, the combi-
slope failures with SWI nation model has higher precision in the estimation of
snow density. Therefore, calculating the SWI by the combi-
4.1. Applicability of SWI incorporating snowmelt water nation model is more effective than calculating it by the
existing models. After integrating the hourly snowmelt
This section examines the applicability of the Soil Water water, estimated by the combination model, into the calcu-
Index (SWI) incorporating snowmelt water, which is calcu- lation of the SWI, the SLs from the day before the occur-
lated using the snow density estimated by the newly pro- rence to the day after the occurrence of the Nakayama Pass
posed combination model. As a case study, this section slope failure are plotted in Fig. 7. From Fig. 7, it is recog-
focuses on the slope failures occurring successively on nized that the starting points of the SLs, considering the
May 4th, 2012 and April 7th, 2013 at Nakayama Pass snowmelt water (the line with white points), are much lar-
along National Highway Route 230 in Hokkaido, Japan, ger than those that do not consider the snowmelt water (the
as shown in Fig. 6. On May 4th, 2012, within the range line with black points), meaning that the snowmelt water
of about 40 m in length and up to 110 m in transverse significantly increased the SWI during the long-term period
width, approximately 13,000 m3 of soil moved downward before the occurrence of the slope failure. It is noted that
to the slope toe and caused a 20-day road closure. On April the slope failure occurred at Nakayama Pass in 2012 (lo-
7th, 2013, at 11:20 AM, another slope failure occurred cated on the grid) without setting a CL, as shown in
within the range of 44 m in length and 19 m in height, Fig. 8(a). Therefore, the applicability of a cluster of CLs,
and more than 11,000 m3 of sediment collapsed. Rainfall defined as a collection of all CLs (blue dashed curves) set
in a 5-km-square grid (blue dashed grid) in Sapporo-
Minami-ku by the Hokkaido government in 2012 for the
target area is discussed. However, it is worth noting that
even though the SWI was calculated using the hourly rain-
fall and snowmelt water, the maximum point of the SL
(near the slope failure time) had not yet exceeded the lower
boundary for the cluster of CLs in Sapporo-Minami-ku,
when the slope failure occurred on either May 4th, 2012
or April 7th, 2013 at Nakayama Pass. Moreover, even if
other calculation methods are adopted, the estimated SL
does not reach the lower boundary for the cluster of CLs
regardless of the estimation model used for the snow den-
sity. This implies that the existing CLs (cluster of CLs)
are not applicable and are not realistic for predicting the
rainfall- and/or snowmelt-induced soil slope failures at
Nakayama Pass. Due to these facts, Miyazaki et al.
(2017) proposed a new CL (black dash-dot-dot line in
Fig. 7) for predicting rainfall-and/or snowmelt-induced
slope failures at Nakayama Pass. The earlier proposed
CLs are found to be too conservative for predicting the
rainfall- and/or snowmelt-induced slope failures. In consid-
eration of the above, it is indispensable that a new CL be
proposed for predicting the rainfall- and/or snowmelt-
induced slope failures at Nakayama Pass.

4.2. Early warning criteria proposed by slope stability


assessment approach

Fig. 5. Snow density estimated by combination model and field monitor- Freeze-thaw action and rainfall and/or snowmelt water
ing data: (a) snow density vs. snow depth and (b) snow density vs. time. infiltration have a significant effect on the water content
203
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

Fig. 6. Location map of disaster sites at Nakayama Pass.

fluctuation of soil slopes (Ishikawa et al., 2015). Accord- and validity of the improved determination method. A
ingly, Siva Subramanian et al. (2017) proposed a slope sta- two-dimensional numerical model for a non-isothermal
bility assessment approach for simulating the long-term coupled transient seepage analysis (see Fig. 10) was built
fluctuation in the water content and factor of safety for vol- to simulate the long-term fluctuations in the water content
canic soil slopes subjected to freeze-thaw action and rain- and factor of safety of soil slopes subjected to freeze-thaw
fall and/or snowmelt water infiltration. The approach action and rainfall and/or snowmelt water infiltration. The
consists of three parts, namely, initial analysis, soil water non-isothermal coupled transient seepage analysis was per-
content simulation, and slope stability analysis, as shown formed by VADOSE/W (GeoStudio International, 2007)
in Fig. 9. The soil water content simulation includes two and the slope stability analysis was performed by
calculation conditions, the long-term soil water content SLOPE/W (GeoStudio International, 2007). During the
simulation (the days before the slope failure) and the initial analysis and soil water content simulation, a climate
short-term soil water content simulation (on the day of boundary (air temperature, precipitation, humidity, and
the slope failure). In order to predict rainfall-and/or wind speed) was applied to the slope surface (ab, bc, and
snowmelt-induced slope failures, Siva Subramanian et al. cd in Fig. 10). The adiabatic boundary, displacement con-
(2018) proposed a new CL (blue short-dash line in Fig. 7) straint boundary, and no-flow boundary were set on the
for all of Hokkaido based on this approach. However, as right side (de), bottom side (ef), and left side (af) of the
the properties of the volcanic soil that is widely distributed model, as shown in Fig. 10. The simulation period of the
in Hokkaido were used for the numerical simulation, initial analysis and long-term soil water content simulation
instead of the local soil properties, and the difference in was from November 8th, 2012 to April 6th, 2013 with a
CLs due to the different local geology and geography in time step of 1 day, and the simulation period of short-
the different areas was not considered, the accuracy of term soil water content simulation and slope stability anal-
the new CL proposed by Siva Subramanian et al. (2018) ysis was from 00:00 to 24:00 on April 7th, 2013 with a time
is not high at Nakayama Pass, as shown in Fig. 7. There- step of 1 hour. The soil properties of Nakayama Pass and
fore, in order to propose new early warning criteria for Nissho Pass, listed in Table 3, were used to discuss the
greater accuracy in predicting rainfall- and/or snowmelt- influence of the soil properties on the new CL. The param-
induced slope failures, this section attempts to improve eters, i.e., dry density (qd), porosity (n), saturated hydraulic
the determination method for setting the CL based on conductivity (ks), effective cohesion (c’), and effective angle
the slope stability assessment approach proposed by Siva of internal friction (/’), were obtained from laboratory ele-
Subramanian et al. (2017). ment tests (Sato et al., 2017; Fukutsu and Kawamura,
In this section, the slope failures occurring along 2016). The parameters for which no laboratory measure-
Nakayama Pass and Nissho Pass in Hokkaido, Japan are ments are available, i.e., thermal conductivity (k), volumet-
taken as typical examples for checking the appropriateness ric heat capacity (f), and residual volumetric water content
204
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

Fig. 7. Soil Water Index calculated by combination model in calculation period: (a) 05/03/2012–05/05/2012 and (b) 04/06/2013–04/08/2013.

(hr), were estimated based on the grain size curve of the soil noted that Saito et al. (2010) examined the rainfall intensity
(SoilVision, 2018). According to the results of an investiga- and duration conditions of 1174 shallow landslides that
tion into the slope failures in Hokkaido from 1996 to 2016 occurred from 2006 to 2008, and found that rainfall for
(Yoshino et al., 2018), the slope angles of the slope failures shallow landslide initiation in Japan can be objectively clas-
occurring on cut slopes or embankment slopes were mainly sified into two typical types: the short-duration high-
in the range of 30–49. Therefore, the numerical simulations intensity (SH) and the long-duration low-intensity (LL)
were performed with different initial Ground Water Levels types of precipitation. Chen et al. (2017) highlighted that
(GWL: 5 m, 7.5 m, and 10 m below the slope surface), dif- the SH type of precipitation is associated with a rapid
ferent slope angles (30, 35, 40, and 45), and different slope increase in SWI, while the LL type of precipitation is asso-
heights (5 m, 10 m, 15 m, and 20 m) to reflect the various ciated with a gradual rise in SWI. For example, as seen in
in situ conditions of the embankment slopes. It should be Fig. 7, the precipitation type is determined by the relation-

205
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

Fig. 8. Cluster of CLs in local region: (a) Sapporo-Minami-ku, (b) Setana-cho, (c) Shinhidaka-cho, and (d) Kaminokumi-cho.

ship between the general overall trend of the snake line of the same GWLs and slope heights, more slope failures
(blue dashed arrow) and the threshold line (red solid arrow occurred with larger slope angles. And, in case of the same
with the slope of 0.04 proposed by Saito et al. (2010)). It GWLs and slope angles, large slope heights increased the
can also be seen in this figure that the SLs during the potential instability of the soil slope. The reason is that
snow-melting season include both the SH and the LL types the shear force is primarily related to the slope angle and
of precipitation. The main difference between the LL type the slope height. Additionally, with a rise in the GWL,
and the SH type is the slope of the SL. This is because the reduced suction causes a drop in shear strength; and
the LL type of precipitation during the snow-melting sea- thus, the slope stability becomes unstable. For example,
son is caused by slow snowmelt, whereas the SH type of in the case of Nakayama Pass, for gentle slopes (30and
precipitation is caused by rapid snowmelt, as seen in the 35), no slope failures occurred when the GWL was low
meteorological monitoring data for both years. In this due to high suction, while the slope stability was greatly
study, therefore, two types of rainfall conditions are affected by the slope height when the GWL was high.
assumed: SH type (30 mm/h for 6 h) and LL type For steep slopes (40and 45), since the slope stability was
(10 mm/h for 24 h). The snowmelt rate is assumed to be originally very low and slight changes in shear strength
a sinusoidal function with a peak of 4 mm/h (slow snow- or shear stress can cause slope failure, both the GWL
melt rate) and 15 mm/h (rapid snowmelt rate) at noon, and the slope height are seen to have a strong impact on
as listed in Table 4. Accordingly, fifteen combinations of the slope stability.
precipitation (rainfall and snowmelt) conditions were Furthermore, the SLs corresponding to the stable cases
adopted, as summarized in Table 5. and the slope failure cases for Nakayama Pass and Nissho
In total, 720 numerical models were simulated for Pass are plotted in Fig. 12. It is noted that the average CL
Nakayama Pass and Nissho Pass. Based on the results of for the cluster of CLs is set as RBFN 0.1 to determine the
the numerical simulations, the numbers of slope failures RBFN value of the new early warning criterion based on
for all the analytical conditions are summarized in the RBFN method (Kuramoto et al., 2001). In Fig. 12,
Fig. 11. By analyzing the results, it is found that, in the case the white points indicate the SLs for the numerical stable
206
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

Fig. 9. Slope stability assessment approach (). adapted from Siva Subramanian et al., 2017

Fig. 10. Two-dimensional numerical model with applied boundary conditions and FEM mesh.

cases, while the black points indicate the SLs for the tation are RBFN 0.2 and RBFN 0.3, respectively, for the
numerical slope failure cases (the first points with factors two areas. This is because the difference in the location of
of safety of less than 1.0). The margin between the stable the RBFN 0.1 curves obtained from the cluster of CLs
cases and the failure cases can be regarded as the new has already taken the variations in of the local geology
CL. From Fig. 12, it is recognized that there is a difference and geography into consideration. Moreover, the newly
in the location of the margin caused by the difference in the proposed CLs for the SH (blue dashed line) and the LL
strength parameters of the soil slope in the numerical mod- (red dash-dot line) types of precipitation at Nakayama Pass
els. Accordingly, when setting the new CL in different are depicted in Fig. 7. It is recognized that the newly pro-
areas, the differences in local geology and geography posed CLs are applicable for precisely predicting the SH
should be considered. On the other hand, in Fig. 12, the and LL types of precipitation-induced slope failures at
RBFN values of the new CLs, determined by the slope sta- Nakayama Pass, and that they are more reasonable than
bility assessment approach and the average CL for the clus- those proposed in past researches (black dash-dot-dot line
ter of CLs, are the same even though the soil properties and blue short-dash line in Fig. 7). These results indicate
between Nakayama Pass and Nissho Pass are different. that, as existing early warning criteria near the target area
For example, the CLs for the SH and LL types of precipi- have already have taken the influences of the variations in
207
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

Table 3
Soil properties used in numerical simulations for Nakayama Pass and Nissho Pass.
Property name Nakayama Pass Nissho Pass
Dry density, qd (kg/m )
3
1423 1695
Porosity, n 0.47 0.36
Saturated hydraulic conductivity, ks (m/s) 5.62  107 1.12  105
Residual volumetric water content, hr (m3/m3) 0.05 0.035
Saturated volumetric water content, hs (m3/m3) 0.47 0.36
Unfrozen thermal conductivity of saturated soil, ku (kJ/(Daym°C)) 127 107.8
Frozen thermal conductivity of saturated soil, kf (kJ/(Daym°C)) 132.2 171.7
Unfrozen volumetric heat capacity, fu (kJ/(m3°C)) 2237.7 3372
Frozen volumetric heat capacity, ff (kJ/(m3°C)) 1624.2 1193
Effective cohesion, c’ (kPa) 0 0
Effective angle of internal friction, /’ (°) 35 37

Table 4
Five types of rainfall conditions and three snowmelt rates used in numerical simulations.
Name Rainfall intensity (mm/h) Duration Type
LL 10 24 h Continuous
SH1 30 6 h (00:00–06:00) Continuous
SH2 30 6 h (06:00–12:00) Continuous
SH3 30 6 h (12:00–18:00) Continuous
SH4 30 6 h (18:00–24:00) Continuous
Name Snowmelt rate (mm/h) Duration Type
SM1 0 24 h Continuous
SM2 0–4 24 h Sinusoidal (Max. at 12:00)
SM3 0–15 24 h Sinusoidal (Max. at 12:00)

Table 5 Hokkaido, Japan, and classified them into three different


Combination conditions of rainfall and snowmelt. patterns of slope failures, as shown in Fig. 13. Table 6 pro-
Combination condition Snowmelt rate vides information on the disasters occurring across all of
SM1 SM2 SM3 Hokkaido over the last twenty years (Iwakura et al.,
2010; Kobayashi et al., 2014). The three different patterns
Rainfall SH1 SH1 + SM1 SH1 + SM2 SH1 + SM3
SH2 SH2 + SM1 SH2 + SM2 SH2 + SM3 of slope failures proposed by Kobayashi et al. (2014) are
SH3 SH3 + SM1 SH3 + SM2 SH3 + SM3 identified for the 12 regions presented in Fig. 14. The figure
SH4 SH4 + SM1 SH4 + SM2 SH4 + SM3 indicates that the three patterns of slope failures can reflect
LL LL + SM1 LL + SM2 LL + SM3 the slope failures in all the regions of Hokkaido. The three
different patterns have the following characteristics: In Pat-
tern A, the soil deposited on the top terrace of the rock
the local geology and geography into account, the new CL slope collapses with a small volume. From the upper part
determined by the slope stability assessment approach and of the slope, the debris mixes with the soil and sand flows
the average CL for the cluster of CLs can be adopted as a to the roadway. In Pattern B, the failure happens on the
more rational early warning criterion for local areas, in cut slope of the thick soil sediment. It collapses from a rel-
accordance with the precipitation types and the local atively deep depth with a large scale to reach the roadway.
ground conditions of the slope, regardless of the soil In Pattern C, topsoil is distributed on a steep rock slope.
properties. The depth of the collapse is shallow and the scale is small.
Among the three patterns, Pattern B is the same as the
4.3. Applicability of proposed stability assessment approach slope failure at Nakayama Pass.
to natural slopes and artificial slopes over all of Hokkaido To discuss the applicability of the proposed stability
assessment approach for the three different patterns of
The above parametric studies were based on homoge- slope failures, a series of numerical simulations for the
neous embankment slopes with different slope angles and three patterns are performed using the soil properties for
heights, while the stability assessment approach proposed Nakayama Pass employed in the last section. The analyti-
in the present study should also be applicable for setting cal method as well as the simulation period of the initial
early warning criteria for natural slopes and artificial slopes analysis, soil water content simulation, and slope stability
(cut slopes and embankment slopes) throughout Hok- analysis are the same as those used in Section 4.2. Accord-
kaido. Kobayashi et al. (2014) investigated the slope fail- ing to the results of the investigation into the slope failures
ures in the Oshima, Hidaka, and Rumoi districts of in Hokkaido from 1996 to 2016 (Yoshino et al., 2018), the
208
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

slope angles of the slope failures occurring on natural and the soil properties are the same as those given in
slopes are mainly in the range of 30–59, while those occur- Table 3. A total of 60 numerical simulations using four
ring on cut slopes or embankment slopes are mainly in the slope models (Table 7) under fifteen combinations of pre-
range of 30–49. For both natural slopes and artificial cipitation conditions (Table 5) were performed for the pat-
slopes, the analysis of the relatively steep slopes (slope terns of slope failures. Fig. 15 shows the new CLs for the
angles are greater than 45) is of little significance. This is three different patterns of slope failures determined by
because relatively steep slopes beside roads are rare or the numerical results. The occurrence of slope failure
because they have had supporting structures applied to (black point) indicates the first point with a factor of safety
them to ensure the safety of the roads. Therefore, the three less than 1.0 along with the SL for the numerical slope fail-
different patterns of slope failures were separately simu- ure case. Moreover, Fig. 15 also shows the slip surfaces of
lated using four slope models, namely, the high-steep slope the short-steep slope models under the LL and SH types of
(20 m, 45), high-gentle slope (20 m, 30), short-steep slope precipitation for each pattern of slope failure. From the
(5 m, 45), and short-gentle slope (5 m, 30) models, as shapes of the slip surfaces shown in Fig. 15, it can be seen
shown in Table 7. The boundary conditions are the same that the slope failures are shallow for the Pattern A and C
as for the model shown in Fig. 10. For the Pattern A and slopes, and that the positions of the slip surfaces are almost
Pattern C slopes, the GWL was set at the interface between the same under the SH and LL types of precipitation. For
the bedrock and the soil. For the Pattern B slope, the GWL the Pattern B slope, the SH type of precipitation rapidly
was set to 3 m below the slope surface. The bedrock increases the saturation of the surface layer causing shallow
properties used in the simulations are listed in Table 8 slope failures. On the other hand, as there is no bedrock,

Fig. 11. Numbers of slope failures for (a) Nakayama Pass and (b) Nissho Pass.
209
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

Fig. 12. Possible revisions of new CLs based on numerical simulation results for (a) Nakayama Pass and (b) Nissho Pass.

the rainwater and/or snowmelt water in the case of the LL 5. Proposal of new determination method for early warning
type of precipitation can slowly infiltrate into deeper soil criteria
layers resulting in an increase in the saturation of the dee-
per layer, which induces the occurrence of a deep-seated 5.1. New determination method for early warning criteria
slope failure. The simulation results suggest that the CL
needs to be set individually for each combination of three In the previous section, a new early warning criterion
different patterns of slope failures under the two typical was proposed for each combination of three different pat-
types of precipitation conditions summarized in Table 9. terns of slope failures under two typical types of precipita-
In other words, for the Pattern A and C slope failures, tion conditions based on case studies and parametric
the existing early warning criteria under the SH and LL analyses. Furthermore, it was found that by using the clus-
types of precipitation can be revised to RBFN 0.3 and ter of CLs in the local region as an index for explaining the
RBFN 0.5, respectively, while for the Pattern B slope fail- local geology and geography, the new early warning crite-
ures, the existing early warning criteria under the SH and rion (Table 9) can predict the occurrence of rainfall- and/or
LL types of precipitation can be revised to RBFN 0.2 snowmelt-induced slope failures, without directly consider-
and RBFN 0.3, respectively, as the new early warning cri- ing the local soil properties. Therefore, the newly proposed
teria for the early warning of rainfall- and/or snowmelt- method of determination for early warning criteria is
induced slope failures in Hokkaido, Japan. shown in Fig. 16. Firstly, the hourly snowmelt water is
210
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

Fig. 14. Locations of subprefectures with historical records of rainfall-


and/or snowmelt-induced slope failures.

Table 7
Four slope models under three different patterns of slope failures.
Slope model Slope angle, a (°)
30 45
Slope height, H (m) 5 Short-gentle slope Short-steep slope
20 High-gentle slope High-steep slope

of the new early warning criteria based on the RBFN


method (Kuramoto et al., 2001). Finally, new early warn-
ing criteria are proposed for predicting three different pat-
Fig. 13. Schematic models and characteristics for the three different terns of slope failures under the two typical types of
patterns of slope failures.
precipitation conditions by considering the local ground
conditions determined from the disaster prevention inspec-
incorporated into the Japanese early warning system using tion records and the cluster of CLs in the local area.
meteorological monitoring data and modeled snow density.
Then, from the calculated SL, the precipitation conditions 5.2. Applicability and reliability of new determination
are divided into two typical types: the SH type and the LL method for early warning criteria
type. Next, the cluster of CLs is plotted using all the exist-
ing CLs in the local area, and the average CL for the cluster The applicability and reliability of the new determina-
of CLs is set as RBFN 0.1 to determine the RBFN values tion method for early warning criteria are discussed against
Table 6
List of rainfall- and/or snowmelt-induced sediment disaster events in Hokkaido, Japan (after Iwakura et al., 2010; Kurahashi et al., 2018).
No. Disaster area Date Characteristic of the disaster Slope pattern Cause
1 Shimokawa-cho 04-18-1999 17:00 Soil collapse deposited on top of rock slope Pattern A Rainfall and snowmelt
2 Shinhidaka-cho 05-15-2000 01:00 Soil collapse deposited on top of rock slope Pattern A Rainfall and snowmelt
3 Yakumo-cho 02-22-2004 19:00 Soil collapse deposited on top of rock slope Pattern A Rainfall and snowmelt
4 Kaminokumi-cho 08-03-2008 17:28 Cut slope failure Pattern B Rainfall
5 Setana-cho 07-29-2010 Topsoil collapse distributed on roadside slope Pattern C Rainfall
6 Setana-cho 07-16-2011 Soil collapse deposited on top of rock slope Pattern A Rainfall
7 Tomamae-cho 04-26-2012 Topsoil collapse distributed on roadside slope Pattern C Rainfall and snowmelt
8 Yubari City 04-27-2012 Embankment slope failure Pattern B Rainfall and snowmelt
9 Rausu-cho 04-24-2015 18:00 Topsoil collapse distributed on roadside slope Pattern C Rainfall and snowmelt
10 Rikubetsu-cho Early April-2016 Topsoil collapse distributed on roadside slope Pattern C Snowmelt
11 Hiroo-cho 03-09-2018 Topsoil collapse distributed on roadside slope Pattern C Rainfall and snowmelt

211
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

Fig. 15. New CLs and simulated slip surfaces for the three different patterns of slope failures.

actual cases of rainfall- and/or snowmelt-induced slope Setana-cho, most of the CLs (blue line) are concentrated
failures in the subprefectures listed in Table 6. At first, in a small area of the cluster of CLs, while a few CLs
Setana-cho, Shinhidaka-cho, and Kaminokumi-cho are (red solid line, the CL set in the red solid grid) significantly
taken as examples for explaining the determination process increase the variation in clusters of CLs, as shown in Fig. 8
of the cluster of CLs, as shown in Fig. 8. During the deter- (b). Therefore, the red solid CLs are excluded because they
mination process of the cluster of CLs in Setana-cho, are not considered to be representative CLs describing the
Shinhidaka-cho, and Kaminokumi-cho, for Shinhidaka- soil properties and ground conditions of the local area.
cho and Kaminokumi-cho, all of the CLs (blue dashed line, Fig. 17 shows the SLs when slope failures occurred in
the CL set in the blue dashed grid) are collected to plot the Setana-cho, Shinhidaka-cho, and Kaminokumi-cho during
cluster of CLs shown in Fig. 8(c) and (d). However, for the heavy rainfall season (non-snow-melting season). It is
212
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

Fig. 16. New method of determination for early warning criteria.

Table 8
Bedrock properties used in numerical simulations for natural slopes. early warning criteria are much higher than the realistic
Property name Value
ones in these areas. Subsequently, a new CL is proposed
(red dashed line in Fig. 17) based on the new determination
Dry density, qd (kg/m )
3
2000
Porosity, n 0.2
method for early warning criteria for the slope failures at
Saturated hydraulic conductivity, ks (m/s) 3.47  109 Setana-cho, Shinhidaka-cho, and Kaminokumi-cho. It is
Residual volumetric water content, hr (m3/m3) 0.008 recognized that the newly proposed CLs can successfully
Saturated volumetric water content, hs (m3/m3) 0.2 predict the occurrence of slope failures in these areas dur-
Unfrozen thermal conductivity of saturated soil, ku (kJ/ 117.6 ing the heavy rainfall season, meaning that the newly pro-
(Daym°C))
Frozen thermal conductivity of saturated soil, kf (kJ/ 173
posed determination method can be applied to the
(Daym°C)) prediction of slope failures during the heavy rainfall season
Unfrozen volumetric heat capacity, fu (kJ/(m3°C)) 2490 in arbitrary areas of Hokkaido. The main reason is that the
Frozen volumetric heat capacity, ff (kJ/(m3°C)) 1567 distribution of soil saturation is mainly affected by the pre-
Effective cohesion, c’ (kPa) 37 cipitation conditions, as discussed in Section 4.3. There-
Effective angle of internal friction, /’ (°) 21
fore, when using SWI to predict the occurrence of slope
failures in the mountain areas shown in Fig. 17, it can be
considered that the slope failure mechanism only depends
recognized that the maximum point of the SL (near the on the slope of the SL (rainfall intensity), and that the tim-
slope failure time) has not yet exceeded the lower boundary ing of the occurrence of slope failures is mainly affected by
for the cluster of CLs, meaning that the existing Japanese the difference in the starting points (initial conditions of the
water content due to the former rainfall) and slopes of the
SL. However, it is worth noting that there is a difference in
the CL locations between the cluster of CLs and the newly
proposed CL. The main reason is that the target areas in
Table 9
this study are mainly mountain areas while, as shown in
Summary of newly proposed early warning criteria.
Fig. 8, the Hokkaido government has not set the CLs in
Rainfall type Pattern A slope Pattern B slope Pattern C slope
many mountain areas, so they differ from the setup areas
SH type RBFN 0.3 RBFN 0.2 RBFN 0.3 for the existing CLs. Since mountain areas have steeper
LL type RBFN 0.5 RBFN 0.3 RBFN 0.5
water catchment terrain and weaker disaster-prevention
213
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

considered, the peak value of the snake line (the line with
white points) has not yet exceeded the lower boundary
for the cluster of CLs, which is inconsistent with the occur-
rence of the disasters. Accordingly, the new CLs for pre-
dicting rainfall- and/or snowmelt-induced slope failures
are proposed in these subprefectures based on the newly
proposed determination method for early warning criteria
shown in Fig. 16. It is recognized that the new CLs (red
dashed lines in Fig. 18) can successfully predict the occur-
rence times of rainfall- and/or snowmelt-induced slope fail-
ures under each combination of the three different patterns
of slope failures under the two typical types of precipitation
conditions in all these areas. Through the comparisons seen
in Figs. 17 and 18, it seems reasonable to conclude that the
newly proposed determination method is effective for
rationally setting early warning criteria not only for
rainfall-induced slope failures without freeze-thaw actions,
but also for rainfall- and/or snowmelt-induced slope fail-
ures with freeze-thaw actions in arbitrary areas of Hok-
kaido, Japan. The reason is that the frozen soil layer
disappears at the end of March in Hokkaido, Japan (Zhu
and Ishikawa, 2018), while most slope failures occur in
April and May, as shown in Fig. 18. If some time passes
after the frozen soil layer disappears, the effects of freeze-
thaw action can hardly be seen. Therefore, the influence
of freeze-thaw action on the setting of early warning crite-
ria is ignorable, which agrees well with the findings by Siva
Subramanian et al. (2017), namely, that freeze-thaw action
has a very small impact on the soil water content and the
factor of safety. On the other hand, Rahardjo et al.
(2007) found that the soil properties and rainfall intensity
are the primary factors controlling the instability of slopes,
while the initial GWL and slope geometry only play a sec-
ondary role. In this study, the above four factors are con-
sidered in the new determination method for early warning
criteria, i.e., the rainfall intensity and/or snowmelt rate, ini-
tial GWL, soil properties, and slope geometry are repre-
Fig. 17. Newly proposed CL for subprefectures in Hokkaido during
sented by the slope of the SL, starting point of the SL,
heavy rainfall season.
local existing early warning criteria, and slope pattern,
respectively. According to the verification of the newly pro-
measures as compared with town areas, the newly pro- posed determination method for early warning criteria by
posed CL is smaller than the cluster of CLs. This indicates the case studies shown in Figs. 7, 17, and 18 (five of the
that although clusters of CLs are effective in town areas, six kinds of cases shown in Table 9 are verified), it can be
they are not applicable in mountain areas. considered that the newly proposed determination method
Furthermore, the newly proposed CLs determined by is effective for rationally setting early warning criteria for
the new determination method for early warning criteria arbitrary areas of Hokkaido during both the snow-
against slope failures, occurring in the subprefectures of melting season and the heavy rainfall season. The advan-
Hokkaido during the snow-melting season, are shown in tage of the newly proposed determination method is that,
Fig. 18. According to this figure, before the disasters hap- since the existing early warning criteria near the target area
pened, a large amount of snowmelt water increased the have already taken the influence of the variations in local
SWI, i.e., the moisture content of the soil ground. There- geology and geography into account, the new determina-
fore, it is impossible to use a snake line that does not con- tion method for early warning criteria can be applied to
sider the snowmelt water (the line with black points) when arbitrary areas in seasonally cold regions without directly
predicting the occurrence of slope failures during the snow- considering the local soil properties, in actual design and
melting season. However, even if the snowmelt water is maintenance works.

214
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

Fig. 18. Newly proposed CL in Hokkaido during snow-melting season for (a) Shimokawa-cho, (b) Shinhidaka-cho, (c) Yakumo-cho, (d) Kiritachi Pass,
(e) Yubari City, (f) Rausu-cho, (g) Rikubetsu-cho, and (h) Hiroo-cho.

6. Conclusions the hourly snowmelt has high accuracy. For this reason,
a combination model was proposed in this study for esti-
The main findings derived from this study are summa- mating the snow density more accurately in seasonally
rized as follows: cold regions by adopting the SY08 model in the midwinter
period and the E86 model in the early spring period.
– ‘When the snow density during the midwinter (1st – The Hokkaido government does not set the CLs in
December-1st March) and early spring (1st March-30th many mountain areas, and the Japanese early warning
April) periods is estimated separately, the calculation of system does not consider the long-term snowmelt water
215
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

infiltration. Therefore, based on the results of case stud- Chen, C.W., Saito, H., Oguchi, T., 2017. Analyzing rainfall-induced mass
ies and parametric analyses for slope stability assess- movements in Taiwan using the soil water index. Landslides 14, 1031–
1041.
ment, new early warning criteria for rainfall- and/or Chen, J., Adams, B.J., 2006. Semidistributed form of the tank model
snowmelt-induced slope failures in Hokkaido, Japan coupled with artificial neural networks. J. Hydrol. Eng. 11 (5), 408–417.
were proposed in this study individually for each combi- Endo, A., 1986. A question of snow density and maximum snow depth. J.
nation of three different patterns of slope failures under Jpn. Assoc. Snow Eng. 1986 (2), 23–30 (in Japanese).
two typical types of precipitation conditions in moun- Fukutsu, K., Kawamura, S., 2016. Field monitoring of road embank-
ments in snowy cold regions and its evaluation. 56th Annual Meeting
tain areas. of Hokkaido Branch Japanese Geotechnical Society, pp. 319–328 (in
– This study proposed a new determination method for Japanese).
setting early warning criteria of rainfall- and/or GeoStudio International, 2007. Calgary, Alberta, Canada.
snowmelt-induced slope failures by considering the local Glade, T., Nadim, F., 2014. Early warning systems for natural hazards
existing early warning criteria, in the case of the same and risks. Nat. Hazards 70, 1669–1671.
Hong, N., Hama, T., Kitajima, T., Aqili, S.W., Huang, X., Wei, Q.,
precipitation (rainfall and snowmelt) types and local Kawagoshi, Y., 2015. Simulation of groundwater levels using tank
ground conditions of the slope. Since the existing early model with consideration of mixed hydrological structure in Kuma-
warning criteria near the target area have already taken moto city. J. Water Environ. Technol. 13 (4), 313–324.
the influence of the variations in the local geology and Hossain, A.F.M., 1989. Comparison of the HYSIM and the tank models
geography into account, the new determination method in Cho Shui River Basin, Taiwan. J. Kejuruteraan 1, 65–86.
Ishikawa, T., Miura, S., 2011. Influence of freeze-thaw action on
for early warning criteria can be applied to arbitrary deformation-strength characteristics and particle crushability of vol-
areas of seasonally cold regions without directly consid- canic coarse-grained soils. Soils Found. 51 (5), 785–799.
ering the local soil properties, in the actual design and Ishikawa, T., Tokoro, T., Miura, S., 2015. Geohazard at volcanic soil
maintenance works. slope in cold regions and its influencing factors. Jpn. Geotech. Soc.
Spec. Publ. 1, 1–20.
Ishikawa, T., Tokoro, T., Miura, S., 2016. Influence of freeze-thaw action
From the research findings, an effective method has been on hydraulic behavior of unsaturated volcanic coarse-grained soils.
proposed for determining early warning criteria of rainfall- Soils Found. 56 (5), 790–804.
and/or snowmelt-induced slope failures in seasonally cold Iwakura, T., Kamihara, T., Tanimoto, H., Otani, K., Tokioka, S.,
regions. The method can be used in both the wide-range Watanabe, T., Onishi, M., 2010. Study of Warning and Evacuation
assessment of slope stability and in the implementation of Standard-setting Approach to the Sediment-related Disasters that
Occur in the Snowmelt Season. JSECE Publication, pp. 520–521 (In
structural measures for three different patterns of slope fail- Japanese).
ures during heavy rainfall and snow-melting seasons, in the Jaiswal, R.K., Ali, S., Bharti, B., 2020. Comparative evaluation of
actual design and maintenance works. conceptual and physical rainfall-runoff models. Appl. Water Sci. 10
(1), 1–14.
Acknowledgments Japan Meteorological Agency (JMA), 2019. Talas Soil Water Index (in
Japanese). <http://www.jma.go.jp/jma/en/News/Talas_soil_water_in-
dex.html> (date accessed: March 16th, 2019).
We gratefully acknowledge the Sapporo Road Office, Kazama, S., Izumi, H., Sarukkalige, P.R., Nasu, T., Sawamoto, M., 2008.
Bureau of Hokkaido Development and Docon Co., Ltd. Estimating snow distribution over a large area and its application for
for their surveyed data and comments. This research was water resources. Hydrol. Process. 22, 2315–2324.
supported in part by Grants-in-Aid for Scientific Research Kobayashi, S., Kon, H., Kohata, Y., Tajika, J., Tanaka, H., Mikita, M.,
2014. Rainfall types, geomorphological and soil physical characteris-
(A) (16H02360) from the Japan Society for the Promotion tics of slope disasters in Oshima, Hidaka and Rumoi districts, in
of Science (JSPS) KAKENHI. The weather station data Hokkaido. In: 54th Annual Meeting of Hokkaido Branch Japanese
and CL information used in this research are publicly avail- Geotechnical Society, vol. 54, pp. 143–148 (in Japanese).
able at the website of the Japan Meteorological Agency Kurahashi, T., Yajima, Y., Tsunoda, F., 2017. Investigation and
(http://www.data.jma.go.jp/gmd/risk/obsdl/index.php) Evaluation Methods of Road Slope Stability due to Snowmelt. 1–6
(in Japanese).
and the Hokkaido Sediment Disaster Warning Information Kurahashi, T., Ishimaru, S., Ito, Y., 2018. Geological disasters in the last
System (http://www.njwa.jp/hokkaido-sabou/) 10 years in Hokkaido. Monthly Report of Civil Engineering Research
Institute for Cold Region. No. 784, 57–66 (in Japanese).
References Kuramoto, K., Tetsuga, H., Higashi, N., Arakawa, M., Nakayama, H.,
Furukawa, K., 2001. A study on a method for determining non-linear
Abe, O., Higashiura, M., Numano, N., Sato, A., 1989. Profiles of snow critical line of slope failures during heavy rainfall based on RBF
cover at Shinjo city for 15 years (1973/74-1987/88 winter seasons). Res. network. Proceed. Jpn. Soc. Civil Eng. 50 (672), 117–132 (in Japanese).
Rep. Natl. Res. Center Disaster Prevent. 44, 123–136 (in Japanese). Meloysund, V., Leira, B., Hoiseth, K.V., Liso, K.R., 2007. Predicting
Abe, O., Shimizu, M., 2004. Equivalent snow density to estimate snow snow density using meteorological data. Meteorol. Appl. 14 (4), 413–
load in extremely heavy snow areas. J. Jpn. Soc. Snow Ice 66 (1), 11–16 423.
(in Japanese). Miyazaki, T., Nakatsugawa, M., Usutani, T., 2017. Risk Evaluation of
Arakawa, H., 2012. Data Report of Snow Survey in Hokkaido, 2011–2012 landslide disaster occurrence using quantification of soil moisture
Winter. Ann. Rep. Snow Ice Stud. Hokkaido 31, S1–S14 (in Japanese). considering snowmelt. J. Jpn. Soc. Hydrol. Water Resour. 30 (2), 89–
Cerda-Villafana, G., Ledesma-Orozco, S.E., Gonzalez-Ramirez, E., 2008. 101 (in Japanese).
Tank model coupled with an artificial neural network. Mexican Mizukami, N., Perica, S., 2008. Spatiotemporal characteristics of snow-
International Conference on Artificial Intelligence. Springer, Berlin, pack density in the mountainous regions of the Western United States.
Heidelberg, pp. 343–350. J. Hydrometeorol. 9, 1416–1426.

216
Y. Zhu et al. Soils and Foundations 61 (2021) 198–217

Motoya, K., Kawashima, K., Matsumoto, T., Iyobe, T., 2017. Analysis of Shirakawa, T., 2016. Characteristics of spatial distribution of snow water
average snowpack densities around Akita pref., estimated from the equivalent based on the recent snow surveys, central and eastern
routinely observed snow depth data and the snow water estimation Hokkaido. Ann. Rep. Snow Ice Stud. Hokkaido 35, 39–42 (in
model. Snow Life Tohoku Reg. 32, 9–14 (in Japanese). Japanese).
Nakatsugawa, M., Usutani, T., Miyazaki, T., 2015. Risk assessment of Shirakawa, T., 2017. Report of snow survey of 32 sites in the central and
sediment disaster based on watershed-wide hydrologic processes. E- eastern region of Hokkaido, 2017. Ann. Rep. Snow Ice Stud.
proceeding of the 36th IAHR World Congress, pp. 1–9. Hokkaido 36, 117–120 (in Japanese).
Nie, W., Krautblatter, M., Leith, K., Thuro, K., Festl, J., 2017. A Siva Subramanian, S., Ishikawa, T., Tokoro, T., 2017. Stability assess-
modified tank model including snowmelt and infiltration time lags for ment approach for soil slopes in seasonal cold regions. Eng. Geol. 221,
deep-seated landslides in alpine environments (Aggenalm, Germany). 154–169.
Nat. Hazards Earth Syst. Sci. 17 (9), 1595–1610. Siva Subramanian, S., Ishikawa, T., Tokoro, T., 2018. An early warning
Okada, K., 2001. Soil Water Index. Meteorol. Soc. Jpn. 48 (5), 59–66 (in criterion for the prediction of snowmelt induced soil slope failures in
Japanese). seasonal cold regions. Soils Found. 58 (3), 582–601.
Osanai, N., Shimizu, T., Kuramoto, K., Kojima, S., Noro, T., 2010. SoilVision, 2018. Version, 4.23. SoilVision Systems Ltd. Saskatoon,
Japanese early-warning for debris flows and slope failures using Saskatchewan, Canada.
rainfall indices with Radial Basis Function Network. Landslides 7, Sturm, M., Taras, B., Liston, G.E., Derksen, C., Jonas, T., Lea, J., 2010.
325–338. Estimating snow water equivalent using snow depth data and climate
Phuong, H.T., Tien, N.X., Chikamori, H., Okubo, K., 2018. A hydro- classes. J. Hydrometeorol. 11, 1380–1394.
logical tank model assessing historical runoff variation in the Hieu Sugawara, M., Ozaki, E., Watanabe, I., Katsuyama, Y., 1974. Tank
river basin. Asian J. Water Environ. Pollut. 15 (1), 75–86. model and its application to Bird Creek, Wollombi Brook, Bikin
Piciullo, L., 2016. Performance analysis of landslide early warning systems River, Kitsu River, Sanaga River, and Nam Mune. Research Note of
at regional scale. Ph.D. Thesis, UNISA, Salerno, Italy, 186 pp. the National Research Center for Disaster Prevention 11, Tsukuba,
Rahardjo, H., Ong, T.H., Rezaur, R.B., Leong, E.C., 2007. Factors Japan, pp. 1–64.
controlling instability of homogeneous soil slopes under rainfall. J. Tomomura, M., Fukushima, Y., Suzuki, M., Kubota, J., Ohta, T., 1982.
Geotech. Geoenviron. Eng., 1532–1543 Distribution of water equivalent of snow cover by altitude in the
Saito, H., Nakayama, D., Matsuyama, H., 2010. Two types of rainfall northwestern mountains of Lake Biwa. Bullet. Kyoto Univ. Forests
conditions associated with shallow landslide initiation. SOLA 6, 57–60. 54, 106–120 (in Japanese).
Sato, A., Hayashi, T., Hayashi, H., Yamaki, M., 2017. On the geotech- Yajima, Y., Kurahashi, T., 2015. The factor of road slope disasters in
nical properties of decomposed granite soil in Hokkaido. 57th Annual Hokkaido during snowmelt period. Proc. Res. Meet. Jpn. Soc. Eng.
Meeting of Hokkaido Branch Japanese Geotechnical Society, pp. 145– Geol. 2015, 67–68 (in Japanese).
148 (in Japanese). Yoshino, K., Agui, K., Asai, K., 2018. Geological issues in slope failures.
Shinme, R., Yamashita, S., 2008. Hydrologic measurement using snow Monthly Rep. Civil Eng. Res. Inst. Cold Reg. 785, 44–55 (in Japanese).
cover weight meter in winter. Proc. Hydraulic Eng. 52, 493–498 (in Zhu, Y.L., Ishikawa, T., 2018. Effects of Climate Change on Slope Failure
Japanese). in Snowy Cold Regions. In: Proceedings of the 61st Symposium of
Shirakawa, T., Ogura, M., Ozeki, T., Takahashi, S., 2014. Report of snow Japanese Geotechnical Society. 6-5, 1-6.
survey in the central and east region of Hokkaido, 2014. Ann. Rep.
Snow Ice Stud. Hokkaido 33, 19–22 (in Japanese).

217

You might also like