Mechanisms of Shallow Landslides On Soil-Mantled Hillslopes With Permeable and Impermeable Bedrocks in The Boso Peninsula, Japan

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Geomorphology 76 (2006) 92 – 108

www.elsevier.com/locate/geomorph

Mechanisms of shallow landslides on soil-mantled hillslopes with


permeable and impermeable bedrocks in the Boso Peninsula, Japan
Yuki Matsushi *, Tsuyoshi Hattanji, Yukinori Matsukura
Graduate School of Life and Environmental Sciences, University of Tsukuba, Tenno-dai 1-1-1, Tsukuba, Ibaraki 305-8572, Japan
Received 5 July 2005; received in revised form 19 October 2005; accepted 20 October 2005
Available online 13 December 2005

Abstract

Rainfall-induced shallow landslides play a vital role in hillslope denudation processes in humid temperate regions. This study
demonstrates the contrasting mechanisms of landslides in adjoining hills with permeable (sandstone) and impermeable (mudstone)
bedrock in the Boso Peninsula, Japan. The characteristics of slope hydrology were inferred from pressure-head monitoring and
rainfall–runoff observations. An analysis of slope stability provided critical conditions for several previously occurring landslides.
The results are as follows. (1) In slopes with the permeable sandstone, infiltrated rainwater percolates through the bedrock as an
unsaturated gravitational flow. The wetting front migration results in a decrease of soil cohesion and causes landsliding at the steep
lower part of the hillslopes. (2) In contrast, the impermeable mudstone beneath a thin soil layer causes a transient positive pressure
head that generates a saturated subsurface storm flow. The reduction in effective normal stress triggers shallow soil slipping at the
uppermost part of a hollow.
D 2005 Elsevier B.V. All rights reserved.

Keywords: Landslides; Hillslope hydrology; Slope stability; Permeability; Rock control

1. Introduction Physically motivated models were developed in the


late 1980s and 1990s to explain landslide initiation
Rainfall-induced shallow landslides frequently occur and distribution in mountain drainage basins. The
on soil-mantled steep hillslopes in humid temperate models coupled an infinite slope stability analysis
regions as a dominant denudation process. Repetition with the concept of steady-state saturated throughflow
of such landslides deform hillslopes and in turn drain- (Montgomery and Dietrich, 1989, 1994; Wu and
age basins. Several studies have established that hydro- Sidle, 1995). These steady-state models achieved suc-
logical processes affect the landslide initiation on cess in assessing topographic control on landslide
hillslopes (Onda, 1992, 1994; Van Asch et al., 1996; susceptibility (Iida, 1999; Borga et al., 2002). How-
Terlien, 1997; Onda et al., 2004). These studies empha- ever, two issues remain unsolved regarding steady-
sized the importance of a comprehensive understanding state models. One is a time scale discrepancy in the
of the relation between subsurface water behavior and supposed hydrological process. The concept of steady
slope destabilization mechanisms. groundwater flow parallel to the slope above an im-
permeable bed can predict only the long-term distri-
* Corresponding author. Tel.: +81 29 853 5691; fax: +81 29 853
bution of groundwater pressure, which should be
6879. identified as a predisposition to landsliding. The sec-
E-mail address: matsushi@atm.geo.tsukuba.ac.jp (Y. Matsushi). ond concern is that the model cannot apply to hill-
0169-555X/$ - see front matter D 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.geomorph.2005.10.003
Y. Matsushi et al. / Geomorphology 76 (2006) 92–108 93

slopes underlain by highly permeable bedrocks, where The present study focuses on slope hydrological
the near-surface lateral water movement becomes an processes and landslide mechanisms in two hilly loca-
unfeasible proposition. The aforementioned shortcom- tions with contrasting morphology and bedrock. We
ings prevent steady-state models from providing a measured geotechnical and hydraulic properties of
rationale for simultaneous landslide occurrence at a slope materials. Pressure head fluctuations in hillslopes
rainfall peak associated with intense rainwater infiltra- as well as rainfall–runoff responses in small watersheds
tion and transient pore-pressure generation (Iverson, were also monitored. Slope stability analysis revealed
2000; D’Odorico et al., 2005). critical conditions for several past landslides. The
Recent slope stability analyses have examined such results indicate different triggering mechanisms induce
a realistic hydrological response and proved rain in- landsliding on hillslopes with disparate lithology. This
filtration and subsequent redistribution of groundwater study also postulates that the difference in the landslide
pressure are sufficiently competent to trigger land- mechanisms account for varying hill morphologies in
slides (Gasmo et al., 2000; Cho and Lee, 2001; Col- this region.
lins and Znidarcic, 2004; Kim et al., 2004). However,
despite these improved models and simulations there 2. Study area
are still few quantitative studies of hillslope hydrolog-
ical processes and landslide mechanisms based on 2.1. Tectonic, climatic and geologic settings
field observations. We cannot expect more pertinent
discussion about the evolution of landscapes in steep The study area is in the south-west Boso Peninsula,
terrains unless the actual slope hydrology is linked central Japan. The area is located on the edge of the
with the critical condition for landsliding. Certain North American plate, bordered on the south by the
field evidence is needed to understand how subsur- Philippine Sea plate (Fig. 1). The area has been uplifted
face-water dynamics influence the critical condition rapidly at a rate of N 1 m/kyr throughout the Quaternary
for landsliding. (Kaizuka et al., 2000).

Fig. 1. Location of the study area and shaded relief map of the hilly lands studied. Broken lines indicate geological boundaries.
94 Y. Matsushi et al. / Geomorphology 76 (2006) 92–108

The study area has a humid temperate climate with a 2.2. Topography
mean temperature of 158C and annual rainfall of 1500–
2000 mm (1490 mm in Kisarazu; 1760 mm in Sakuma; The study site exhibits distinct topographic charac-
and 2050 mm in Sakahata; the locations of these cli- teristics that vary from hills in the northwest to hills in
matological stations are shown in Fig. 1). Seasonal the southeast. The central dashed line in Fig. 1 demar-
fronts in early summer and late fall as well as occa- cates this boundary between the varying morphometric
sional typhoons contribute to 40–50% of the total rain- characteristics. Hills in the northwestern section ex-
fall. The most intensive 1-day rainfall over the last 30 hibit relatively high rounded crests (relative relief of
years is 296 mm in Kisarazu, 426 mm in Sakuma, and 150–200 m) with low drainage density (5–8 km 1).
364 mm in Sakahata. The southeastern hills display low rugged ridges (rel-
Brown forest soil covers the hillslopes, except for ative relief of 50–100 m) with high drainage density
narrow ridges or steep slip scars with bedrock expo- (15–22 km 1).
sures. The majority of the study area is covered with a The topographic difference corresponds to the geo-
planted forest of cypress (Chamaecyparis obtusa) and logy of the area (Fig. 2). Coarse sandstone and con-
cedar (Cryptomeria japonica), although hardwood and glomerates (the Ichijuku and Nagahama Formations;
various understory species also coexist within the co- ca. 600–700 ka) comprise the northwest high terrain,
nifer stands. The forest age varies from several years to whereas muddy sandstone and sandy mudstone (the
several decades. No large-scale timber harvesting has Iwasaka and Awakura Formations; ca. 700–800 ka)
been conducted in recent decades. make up the southeast lower terrain. Overall, this
Bedrock in this area belongs to the middle Pleisto- landscape comprises cuesta-like landforms with back-
cene forearc basin fill, referred to as the Kazusa Group slopes facing northwest. Hereafter, we call these two
(Ito, 1995, 1998). Submarine fan successions generated areas the sandstone area and the mudstone area (Figs. 1
by glacioeustasy cycles under the influence of paleo- and 2).
ocean currents produced repeated sandy and muddy
depositional sequences (Ito, 1998; Ito and Horikawa, 2.3. Shallow landslide
2000). The emergence of these submarine deposits in
the last 500 ka has led to the development of small hills Rainfall-induced shallow landslides in the two areas
ranging in elevation from 100–300 m. are of particular interest to this study. Torrential rainfall

Fig. 2. Geologic map and topographic cross-sections of the hilly lands studied. Nh: Nagahama Formation; Ij: Ichijuku Formation; Is: Iwasaka
Formation; Ak: Awakura Formation. The geological map corresponds to the area of Fig. 1.
Y. Matsushi et al. / Geomorphology 76 (2006) 92–108 95

Fig. 3. Landslide density for the past two decades in the hills studied. Bar charts represent number of newly emerged landslide scars on vertical air
photographs taken in 1981, 1984, 1990, 1995 and 2000. The area underwent a torrential rainfall on August 1, 1989 (total rainfall N300 mm).

on August 1, 1989 (total rainfall N 300 mm) produced The number of the landslide scars was identified
many slides along the mudstone slopes, but only a few from vertical air photographs with various scales
landslides occurred in the sandstone area (Furuya and from 1 : 20 000 to 1 : 40 000 for the following years:
Ohkura, 1992). Landslides occurred only on steep 1981, 1984, 1990, 1995 and 2000 (Fig. 3). The
lower hillslopes adjacent to major valleys in the sand- landslide density over the past two decades in the
stone area. While the landslides in the mudstone area mudstone area (127.6 km 2) is about 22 times larger
took place mainly on uppermost hollows near slope than that in the sandstone area (5.7 km 2). The
crests (Matsushi and Matsukura, 2004). adjoining hills have no great variation in climatic

Fig. 4. Locations of the selected slopes including landslide scars (S-1, 2, 3 for the sandstone area (A) and M-1, 2, 3 for the mudstone area (C)), and
detailed topographic maps of the experimental watersheds (S-watershed for the sandstone (B) and M-watershed for the mudstone (D)). See Figs. 1
and 2 for locations. Contours are in meters.
96 Y. Matsushi et al. / Geomorphology 76 (2006) 92–108

conditions, vegetation cover, and tectonic settings. 5 m upstream from the catchment outlet (Fig. 4B). The
Therefore, the difference in the density and locations channel floor is inundated with saturated sediment
of the landslides must reflect hydrological processes resulting from the groundwater seepage. Behind the
and critical limits for landsliding, controlled by the spring, soil-mantled slopes with no slip scars form an
hydraulic and mechanical properties of the hillslope unchanneled dry hollow.
materials. This paper focuses on the validation of this In the mudstone watershed (M-watershed), a channel
hypothesis. with well-defined banks incises the bottom of the wa-
tershed and is dry except for a short period during
3. Methods rainfall events (Fig. 4D). The upstream end of the
channel branches into tributaries and they extend to
3.1. Selection of slopes and watersheds the upper convergent hollows. Several hillslopes near
the deep channel underwent slipping, exposing the
Field and laboratory experiments were designed to mudstone bedrock. Most slopes in the watershed have
understand hillslope hydrological processes and obtain thin soil cover, and some have only organic deposits on
parameter values relevant to slope stability analyses. the bedrock.
For these purposes, three slipped slopes and a small Fig. 5 shows profiles of the six selected slopes, as
watershed were selected from each area (Fig. 4). We well as the plan views and cross-sections of the slip
assumed that hillslope materials on the same substrates scars on the slopes. All the scars have a shallow platy
have similar homogeneous geotechnical properties. form bounded by a small scar step, showing general
The sandstone watershed (S-watershed) maintains a geometry of a translational landslide (Selby, 1993).
stable base flow of 0.2–0.3 L/s welling from a spring at Those landslides seem to have resulted from the

Fig. 5. Profiles of the selected slopes (S-1, 2, 3 and M-1, 2, 3), and plan views and cross-sections of slip scars. The subsurface layers of S-1 and M-1
are inferred from soil soundings. Enlarged views of the upper parts of S-1 and M-1 show tensiometer nests for pressure head monitoring (ST1, 2 and
MT1, 2, 3). The schematic illustration in the lower right shows soil pit scraping at scar heads.
Y. Matsushi et al. / Geomorphology 76 (2006) 92–108 97

heavy rainfall on August 1, 1989, since they were first corresponds to the original height of the scar step
visible on the aerial photograph taken in 1990. (Fig. 5).
Landslides on the same substrates have similar slip Undisturbed soil cores of 5.0 cm diameter and 5.1
depths and slope angles. Table 1 describes the dimen- cm height were extracted at every 10–15 cm depth
sions of the landslides. The listed slip depths measured intervals. These samples provided the depth-averaged
vertically from the original surface as well as slope physical properties and hydraulic characteristics of the
angles represent mean values estimated from the soil. Samples for soil strength determination were also
cross-sections and longitudinal profiles of the land- obtained from the pits 1 and 3 using a trimming ring of
slides. Volumes of the landslides were calculated from 6.0 cm bore diameter and 2.0 cm height. The dimension
the mean slip depth and the sliding area. The landslides of the soil specimens is identical to the shear instru-
on the sandstone slopes are deeper, steeper, and larger ment. Sampling depths are 70–80 cm for the pit 1 (24
than those on the mudstone slopes. cores) and 30–40 cm for the pit 3 (20 cores), which
roughly correspond to the depth of the potential failure
3.2. Investigation of soil thickness planes.

On the S-1 and M-1 slopes, we conducted soil 3.4. Soil tests
soundings along the scar profiles using a simplified
dynamic cone penetrometer. The penetrometer consists The dry unit weight, porosity, grain-size distribution,
of a penetration rod with a cone tip (25 mm diameter and depth profiles of saturated hydraulic conductivity
with 608 tip angle), guide rods and a 5-kg weight. The were obtained from the cored materials. In-situ testing
weight drives the cone into subsoil falling from a of the infiltration capacity was also conducted at the
height of 50 cm along the guide rod. Here, the resis- bottom of the mudstone pit because of difficulty in
tance value for the cone penetration N c is defined as undisturbed sampling. Drying soil–water characteristic
the number of impacts needed for every 10-cm pene- curves were also measured for the soil cores from 30
tration. Using N c, Wakatsuki et al. (2005) distinguished cm depth in the pit 1 (sandstone) and 15 cm depth in the
four subsurface layers: 0 V N c b 5 (upper); 5 V N c b 10 pit 3 (mudstone). Hysteresis between drying and wet-
(middle); 10 V N c b 30 (lower) and N c z 30 (bedrock). ting processes and spatial variance in soil–water reten-
Accordingly, the thickness of soil mantle is defined as tion characteristics were not taken into account.
the depth attained when the N c reaches 30. Therefore, only one soil–water characteristic curve
could be defined for each soil.
3.3. Soil sampling Basic box shear tests using the undisturbed soil
specimens were performed to determine soil strength
In addition to the soil soundings, we scraped soil parameters. The moisture contents of the specimens
pits in order to inspect subsurface structure (Fig. 5; were adjusted stepwise from an air-dried condition to
pits 1 and 2 on the S-1 slope, pit 3 on the M-1 a capillary saturated condition. These specimens were
slope). Subsurface soil profiles were recorded using delicately placed into an isometric shear box after seal-
the exposed upslope face of the pits. The soil profile ing them in a plastic bag for at least a week to allow
at the scar head includes a potential failure plane that moisture equilibration. Single stage direct shear tests at
a strain rate of 1 mm/min were conducted under four
different normal stresses (10, 20, 30 and 40 kPa). The
Table 1 volumetric water content of the specimens was calcu-
Dimensions of the landslides lated from the weight difference between the specimen
Slip depth Slope angle Sliding area Volume just after the shear test and that in completely dried
(m) (degrees) (m2) (m3) condition.
Sandstone
S-1 1.6 38.4 620 990 3.5. Hydrological observation
S-2 1.7 37.1 570 980
S-3 1.4 38.4 230 330
Tensiometers were set-up within 10 m upslope from
Mudstone the scar head on S-1 and M-1 to monitor the pressure
M-1 0.7 32.2 70 50 head (Fig. 5). The tensiometers were established at two
M-2 0.6 35.4 70 40 nests on the S-1 slope (ST1 and ST 2, with ceramic
M-3 0.6 34.7 100 40 cups at depth of 30, 75, and 120 cm), and at three nests
98 Y. Matsushi et al. / Geomorphology 76 (2006) 92–108

on the M-1 slope (MT1, MT 2 and MT 3; the depths of intervals was also observed at the points denoted with
ceramic cups were 30 cm for MT1, 23 cm and 46 cm stars in Fig. 4A and C. The period of these observations
for MT2, and 30, 60, and 90 cm for MT3). A data was from May to October in 2004, including a respite
logger connected to the tensiometers recorded the sub- during dry summer season.
soil pressure head at every 10 min.
Runoff from the S- and M-watersheds (Fig. 4B and 4. Subsurface structure of the slopes
D) was measured at the watershed outlet by means of a
weir and a water-level gauge. The flow depths in the 4.1. Soil thickness
weir were recorded with a data logger at 10-min inter-
vals. The observed water level was converted into a The sandstone and mudstone slopes have signifi-
discharge using a calibration curve. Rainfall in 10-min cantly differing soil thicknesses. The value of N c for

Fig. 6. Subsurface soil profiles with depth variations of the cone penetration resistance (N c values), in the sandstone slope (A), (B) and the mudstone
slope (C). Horizontal narrow bands represent the depth of potential failure planes.
Y. Matsushi et al. / Geomorphology 76 (2006) 92–108 99

the sandstone slope increases gradually and reaches 30 soil profiles. Table 2 presents the depth-averaged values
(bedrock) at depths of 6–7 m. The slip plane of the of these soil properties.
landslide perches on the boundary of the upper The physical properties reflect the grain-size distri-
(0 V N c b 5) and middle (5 V N c b 10) layers. Beneath bution of the soils. Soil originating from the sandstone
the scar head of the landslide, thick soil mantles (4–5 has a large fraction of sand, whereas soil originating
m of N c b 30 layers) remain on the bedrock (Fig. 5). from the mudstone contains more fine materials. Cor-
In contrast, the soil thickness on the mudstone slope respondingly, the soil from the sandstone slope has a
is up to a meter except at the bottom of the slope greater dry unit weight and a smaller porosity than the
covered with colluvial deposits. On the intact part of soil from the mudstone slope (Table 2).
the slope upward from the scar head, soft upper soil
(0 V N c b 5) translates sharply into the bedrock 5.2. Permeability
(N c z 30) at the base of the soil layer. The slip plane
of the landslides lies just above this soil–bedrock Fig. 7 shows depth profiles of the saturated hydrau-
boundary. lic conductivity of the soils. The box-plots in the upper
diagram represent annual maximum 1-h rainfalls from
4.2. Soil profiles 1979 to 2004 at the Kisarazu, Sakuma, and Sakahata
climatological stations. Horizontal whiskers, shaded
The subsurface soil profiles of the slopes S-1 and M- boxes, and vertical lines within the boxes represent
1 have contrasting appearances. Fig. 6 shows the soil the extremes, interquartile range, and median values,
profiles at the pits with depth variations of the N c respectively. The intensity of annual maximum 1-h rain-
values. Horizontal narrow bands on the diagrams rep- falls ranges from 5  10 6 to 2  10 5 m/s (Fig. 7).
resent the potential failure planes (~75 cm deep for S-1, The variation of the saturated hydraulic conducti-
and ~55 cm deep for M-1). vity with depth strongly depends on geology. The
The soil profile of the slope S-1 consists of homo- hydraulic conductivity of the sandstone slope has a
geneous sand (Fig. 6A). The N c value in the pit 1 nearly straight vertical profile, with values around
increases slightly with depth, but no apparent disconti- 10 4 m/s from the shallow soil layer (pit 1; including
nuity is found around the potential failure depth. The the potential failure plane) down to the bedrock (pit 2).
sand at the bottom of the pit 2 on the landslide scar In contrast, the hydraulic conductivity of the mudstone
(Fig. 6B) remains the original beddings of the stratum slope changes significantly with depth. The value on
(the Ichijuku Formation). This observation indicates the the order of 10 5 m/s near the surface falls abruptly to
slope material in this part originates from in-situ weath- 8  10 7 m/s at a depth of 0.7 m. Field measurements
ering of the sandstone bedrock. of the infiltration capacity gave a minimum hydraulic
The soil profile on slope M-1 (Fig. 6C) exhibits conductivity of 5  10 8 m/s at a depth of 0.9 m. The
sharply demarcated weathering horizons that contrast hydraulic discontinuity in the profile at a depth of
the finding from the S-1 profile. Silty organic aggre- about 0.6 m corresponds to the potential failure
gates make up a soft loose soil (N c b 5) that includes plane. This implies a strong relation of rainwater ac-
scattered gravels at depths less than 0.3 m. From 0.3 to cumulation upon the impervious bedrock with the
0.6 m deep, interlocking rock fragments form a dense conformation of the slip surface.
residual horizon (10 V N c b 20). The residual horizon
sharply switches into the parent material (i.e., the silty 5.3. Soil–water retention characteristics
mudstone of the Awakura Formation) below 0.6 m
depth. The soils originating from the two parent materials
have differing soil–water retention characteristics (Fig.
5. Slope materials 8). The volumetric water content of the sandstone slope

5.1. Physical properties Table 2


Depth-averaged physical properties of soils
The soils on the sandstone and mudstone slopes Specific Dry Porosity Grain-size distribution (%)
have different physical characteristics. Eight samples gravity unit wt. (m3/m3) Clay Silt Sand
() (kN/m3)
from the sandstone slope (pit 1) and five samples from
the mudstone slope (pit 3) provided values of soil unit Sandstone 2.69 12.7 0.52 5.6 10.1 84.3
Mudstone 2.68 11.1 0.58 12.4 43.7 43.9
weight, porosity and grain-size distribution through the
100 Y. Matsushi et al. / Geomorphology 76 (2006) 92–108

Table 3
Optimum parameter values for the model fitted to soil–water charac-
teristic curves
Saturated Residual Inflection Parameter r2
v.w.c.* h s v.w.c.* h r point w 0 m ()
(m3/m3) (m3/m3) (m)
Sandstone soil 0.465 0.171 0.186 0.569 0.97
Mudstone soil 0.530 0.061 0.001 0.085 0.81
*v.w.c. = volumetric water content.

pressure head decreases to  0.5 m. However the soil


maintains relatively high volumetric water content (0.4
to 0.35) even when the pressure head is  4 m, since the
small pores within the aggregates may retain immobile
water.
The broken and solid lines in Fig. 8 show the best-fit
model curves. An equation of Van Genuchten’s (1980)
model modified by Kosugi (1994) was adopted:
"   1m 1
#m
w
h ¼ hr þ ð hs  hr Þ 1 þ m ð1Þ
w0

where h is the volumetric water content, h s is the


saturated volumetric water content, h r is the residual
volumetric water content, w is the pressure head, w 0 is
the pressure head at the inflection point, and m is a
constant (0 b m b 1). In this case, the air-entry value of
the soil is assumed to be zero, because the specimens
Fig. 7. Depth profiles of saturated hydraulic conductivity in the
sandstone and mudstone slopes. Box-plots in the upper diagram
are undisturbed natural soils having sufficiently large
represent the annual maximum 1-h rainfalls from 1979 to 2004 at maximum pore sizes to bubble in a near-zero negative
the Kisarazu, Sakuma and Sakahata climatological stations (see Fig. 1 pressure head.
for their locations). Eq. (1) was fitted to the data in Fig. 8 using the non-
linear least squares optimization of Marquardt (1963).
soil falls smoothly from 0.45 to 0.2 as the pressure head The value of h s used was the volumetric water content
decreases from 0 to  1 m and then reaches a residual at zero pressure head, and the other parameters (i.e., h r,
state as the soil contains only suspended water. The soil w 0 and m) were estimated by the regression. The fitting
on the mudstone slope rapidly drains gravitational analysis provided the optimum parameter values as in
water from its inter-aggregate macro pores before the Table 3. The values accurately represent the soil–water

Fig. 8. Soil–water characteristic curves for soils originating from the sandstone (A) and the mudstone (B). The broken and solid curves are best-fit
Kosugi’s (1994) model curves, whose optimum parameter values are listed in Table 3.
Y. Matsushi et al. / Geomorphology 76 (2006) 92–108 101

retention characteristics of each soil (r2 = 0.97 and Table 4


0.81). Optimum parameter values for the regressions for reduction in soil
shear strength
The values of h s and h r are similar to those in a
Ultimate Reduction Angle of shearing r2
previously published catalog (Mualem, 1976); w 0 and
soil cohesion coefficient resistance
m for the soil from the sandstone are also similar to those C (kPa) l () / (degrees)
in Kosugi (1994). However, w 0 and m for the soil from
Sandstone soil 35.8 4.81 28.3 0.88
the mudstone are smaller than the previously reported Mudstone soil 192.9 6.94 27.7 0.93
values. This difference may be a result of the aggregated
structure of the soil, which allows rapid water drainage
from the inter-aggregate large pores. creasing soil moisture (Fredlund and Raharjo, 1993).
However, some previous studies employed a function
5.4. Shear strength of matric suction (soil pore-air pressure minus pore-
water pressure) as a parameter of soil moisture condi-
Shear tests revealed the characteristics of decline in tion (Fredlund et al., 1978; Gan et al., 1988; Vanapalli
shear strength with increasing volumetric water con- et al., 1996), which cannot be determined accurately in
tent (Fig. 9). An analysis of dry to saturated condi- our shear tests. Therefore, we use Eq. (2) as an approx-
tions provides evidence that the soils reduced their imate formula of soil shear strength to evaluate stability
shear strength by 60–70% in the sandstone samples of unsaturated slopes.
(Fig. 9A), and by 70–80% in the mudstone samples The unknowns in the equation (i.e., C, l and /)
(Fig. 9B). were determined by a multiple linear-regression analy-
The solid and broken lines in Fig. 9 represent the sis (Table 4). The variables in the equations reflect
best-fit regression curves that express the relation be- liquefaction characteristics of the soils and susceptibil-
tween the shear strength, normal stresses, and volumet- ity to changes in pore-water pressure during shear
ric water content based on the following formula: deformation. Therefore, we regard these values as pure-
ly empirical rather than physically significant. Instead,
s ¼ Celh þ rtan/ ð2Þ the equations using these soil-specific parameters accu-
where, s is the shear stress, C is the ultimate increment of rately represent the observed strength reduction char-
apparent soil cohesion (when h = 0), l is the reduction acteristics of each soil (Fig. 9).
coefficient, r is the applied normal stress, and / is the
apparent angle of shearing resistance. The equation 6. Subsurface water movement and runoff response
involves an exponential function for the simplest regres-
sion of the change in cohesive component to the soil 6.1. Long-term fluctuations in pressure head and rain-
shear strength, assuming that the angle of shearing re- fall–runoff characteristics
sistance takes a value content.
Eq. (2) is consistent with the existing general remark The pressure heads in the sandstone slope show a
that only the cohesive component decreases with in- small range of fluctuation throughout the observation

Fig. 9. Reduction of shear strength with respect to increasing soil volumetric water content, from air-dried to capillary saturated conditions. (A) Soil
originating from the sandstone, (B) soil originating from the mudstone. The solid and broken lines represent best-fit regression curves for each soil,
whose optimum parameter values are listed in Table 4.
102
Y. Matsushi et al. / Geomorphology 76 (2006) 92–108
Fig. 10. Pressure head fluctuations at ST2 and MT3 and runoff from the experimental watersheds in early summer and fall in 2004.
Y. Matsushi et al. / Geomorphology 76 (2006) 92–108 103

period (Fig. 10A). The pressure head at deeper points up to 0.5 m (Fig. 10B). The pressure head displayed
maintained a negative value even during intensive rain- positive values during almost all distinct rainfall
fall events. Pressure head at the shallowest monitoring events at every depth. Then a marked decrease
depth (30 cm) momentarily exceeded zero only at rain- in the pressure heads continued until the next
fall peaks. rainfall. The pressure heads between the events os-
Runoff from the S-watershed did not increase mea- cillate diurnally because of the evapotranspiration
surably in response to individual storms. The watershed cycles.
maintained a stable discharge less than 0.5 L/s, with Runoff from the M-watersheds responded sharply
slight fluctuations on a rather long time scale. We veri- to rainfall events. The accurate magnitudes of peak
fied the low discharge by field observations during a flows could not be measured because of overflow.
rainfall event. Following these extremes the discharge fell rapidly,
The pressure head on the mudstone slope varied and then the watershed yielded a small discharge
over a wide range during the observation period during the slope maintained the positive pressure
from negative values of  2 m to positive values head.

Fig. 11. Pressure head responses in the sandstone slope (A) and the mudstone slope (C) from May 19 to 21, 2004 (cf. Fig. 10). Slope profiles (B)
and (D) illustrate the time-series distributions of the hydraulic head and of the saturated zone within the slopes. The numbers in boxes in (B) and (D)
correspond to the arrows in (A) and (C), respectively.
104 Y. Matsushi et al. / Geomorphology 76 (2006) 92–108

6.2. Short-term responses of the pressure head Assuming an equable wetting of homogeneous soil,
i.e., uniform soil water content and pressure heads from
The fluctuations from May 19 to May 21 are similar the land surface to the slip plane, FS can be expressed
to those observed during the other rainfall events (Fig. as:
11A, C). Fig. 11B and D illustrate the distributions of
the hydraulic head and the saturated zone at the four tan/ Celh
FS ¼ þ
moments shown by the arrows in Fig. 11A and C. tanb ðcd þ hcw ÞZsinbcosb
The pressure heads in the sandstone slope responded maxðw; 0Þcw tan/
to rainfall sequentially from shallow to deep monitoring  ð4Þ
ðcd þ hcw ÞZsinbcosb
depths (Fig. 11A). The time lag between the rising
edges of the pressure head at shallow and deep loca- where Z is vertical soil depth, b is slope angle, c d is the
tions was about 6 h for the first event on May 19, and dry unit weight of soil, and c w is the unit weight of
became shorter for the subsequent rainfall on May 21. water (9.8 kN/m3).
In addition, the pressure heads rose abruptly at shal- On the right side of Eq. (4), the first term represents
lower depths but more gradually at deeper points. The the frictional component of slope stability; the second
equipotential hydraulic head lines in the sandstone term expresses the change in cohesive resistance in an
slope tended to be horizontal in the absence of signifi- unsaturated state (on the basis of Eq. (2)); and the third
cant saturated zones (Fig. 11B). term refers to slope instabilization by the positive pres-
The pressure heads rose nearly concurrently with sure head in the saturated state (w N 0 m). Substitution
rainfall and reach positive values at all the monitoring of Eq. (1) into Eq. (4) generates the factor of safety
depths in the mudstone slope (Fig. 11C). The sharp mediated by the pressure head, and allows a simulation
increases in the pressure heads began from the deepest of slope instabilization for both unsaturated and satu-
point and rapidly moved to shallower depths. The rated conditions.
positive pressures persisted for several hours at 30 cm The moisture condition of soil between the land
depth, but several days at 90 cm depth. The hydraulic surface and the slip plane is not actually uniform. Soil
head lines under an expanded saturated zone tended to moisture decreases with increasing depth because of
be perpendicular to the slope surface (Fig. 11D). rainwater infiltration from the land surface during an
intensive storm event. Thus the weight of soil mass in
7. Slope stability analysis the unsaturated condition may be slightly underesti-
mated. However, this under estimate has much smaller
7.1. Formulation of the slope stability model influence on FS than that of the other parameters
(Borga et al., 2002; Dykes, 2002). The effect of the
We employed a limit equilibrium method to analyze moisture heterogeneity in the soil column was ignored
the critical conditions for landsliding. In this method, based upon the findings from these previous studies.
materials on a hillslope are subject to two opposing We also did not take into account the stabilizing effect
forces: a downslope component of soil weight as a of tree roots, because the potential failure plane in the
driving force, and shear strength of the soil providing actual soil profiles lies well below the major root zone
a resisting force. The factor of safety FS is: (Fig. 6).

Resisting force 7.2. Critical conditions for past landslides


FS ¼
Driving force
Shear strength of soil The slope stability analysis was applied to specify
¼ : ð3Þ the critical conditions of some past landslides whose
Downslope component of soil weight
scars are described in Fig. 5. We obtained slip depths
If FS b 1, the slope fails along a potential failure plane. and slope angles of the landslides (Table 1) and also
Three instability factors were taken into account: input-variables for the analysis using Eqs. (1) and (4)
reduction of soil cohesion in response to soil wetting, (Tables 2–4). We can therefore simulate the factor of
weight increase of soil resulting from water absorption, safety in those slipped slopes as a function of the
and decrease in effective stress derived from positive pressure head.
pore-water pressure. The former two were considered Fig. 12 juxtaposes curves showing the results of the
only in an unsaturated state, because their values are simulation. FS for the sandstone slopes falls below
fixed at the saturated condition. unity just before the saturation (w =  0.10 m for S-1,
Y. Matsushi et al. / Geomorphology 76 (2006) 92–108 105

Fig. 12. Simulation of the reduction in the factor of safety with respect to the pressure head in hillslopes, using actual slip depths and slope angles of
the six landslides.

 0.05 m for S-2, and  0.03 m for S-3). FS for the drains rainwater through the soil layer out to channel
mudstone slope becomes 1.7–1.8 at the saturation point systems, giving acute runoff peaks in response to every
(w = 0 m), and sinks below unity when the pressure rainfall event (Fig. 10B). In fact, discharge from the
head reaches approximately 0.7 m (w = 0.76 m for M- watershed continues while the pressure head in the
1, 0.69 m for M-2, and 0.71 m for M-3). slope maintains a positive value, and declines as the
pressure head reverts to a negative value.
8. Discussion
8.2. Hydrological triggering mechanisms of landslides
8.1. Slope hydrology
The thick soil layer on the permeable sandstone
Relations between pressure head fluctuations in the bedrock cannot reach a saturated state (Fig. 10A). In
slopes and runoff from the watersheds in each area can this case, wetting from land surface migrates into sub-
be explained as follows. In the sandstone slope, infil- soil without a positive pressure head (Fig. 11A). This
trated rainwater can seep down vertically though the redistribution of incoming rainwater causes reduction in
soil layers into bedrock, because there is no hydraulic cohesive strength of unsaturated slope materials (Fig.
discontinuity in the subsurface profile (Fig. 7). Accord- 9). The critical pressure heads for the sandstone slopes
ingly, equipotential lines of the hydraulic head in the were indeed calculated to be negative values (Fig. 12).
sandstone slope were horizontal (Fig. 11B), indicating Consequently, the wetting front migration by intense
vertical rainwater percolation. This unsaturated gravita- rainwater infiltration causes a landslide associated with
tional water flow should recharge the deep groundwater reduction of soil cohesion (Fig. 13A). This concept of
body in this permeable bedrock. Indeed, no perceptible landslide initiation is supported by a simulation of pore-
responses in runoff were observed (Fig. 10A). Infiltrat- pressure diffusion associated with slope-normal rainwa-
ed rainwater discharges to the stream through the deep ter infiltration (Collins and Znidarcic, 2004). Forefront
aquifer on a long-term basis, giving the observed con- of the down-seeping wetting band locates the slip plane
stant base flow. of this type of landslide.
The impermeable bedrock beneath a rather perme- The thin overburden of the impermeable mudstone
able soil layer prevents the percolation of incoming becomes saturated with a relatively small amount of
rainwater in the mudstone slopes (Fig. 7). The resultant rainfall (Fig. 10B). This results in generation of positive
equipotential lines of the hydraulic head perpendicular pressure heads, building up a shallow transient ground-
to the slope and expansion of the saturated zone dem- water table from soil bottom to upwards (Fig. 11D).
onstrate saturated downslope water movement upon the The mudstone slopes became unstable when the posi-
impermeable bedrock (Fig. 11D). This movement tive pressure head reaches to ~0.7 m (Fig. 12). This
106 Y. Matsushi et al. / Geomorphology 76 (2006) 92–108

Fig. 13. Schematic illustrations showing time-domain alterations in moisture depth profile and subsurface stress states in the sandstone slope (A)
and the mudstone slope (B). The left part of the diagrams indicates wetting processes of slope materials, and the right part represents the change in
shear stress (broken lines) and shearing resistance (solid curves) within the slopes.

indicates that a decrease in effective stress as well as the potential for landsliding. Thus the sandstone area
reduction in soil cohesion triggers a soil slip upon the rarely experience landslide events as shown in Fig. 3.
bedrock (Fig. 13B). The less the landslides occur, the more thick and hardly
The pressure head in the mudstone slope responds saturated soil layers will be preserved on the hillslope
promptly to rainfall, reaching positive states in a short (Fig. 5).
period (Fig. 11C). This lends support to the concept of The landsliding with negative pressure heads usually
efficiency of pore-pressure diffusion associated with occur on hillslopes whose angles are steeper than the
slope-normal rainwater infiltration (Iverson, 2000). It angle of shearing resistance of slope material (Rao,
also implies the concept of the steady-state groundwa- 1996). Slopes gentler than the angle of shearing resis-
ter flow is not appropriate to assess the direct triggering tance can slip only when the ascending force acts on the
of landslides, even where the impermeable bedrock potential failure material. However, such a buoyancy
slopes. In fact, landslides in the mudstone area took effect is not expected for the sandstone slopes, because
place on uppermost hollows (Matsushi and Matsukura, of the absence of the positive pressure head (Figs. 10A
2004) without large contributing areas theorized in the and 11A). The angle of shearing resistance is 28.38 for
steady-state model. the soil from the sandstone (Table 4). Indeed, landslide
locations in the sandstone area are confined to the steep
8.3. Landslide mechanisms, slope conditions, and land- (N308) lower parts of the slopes adjacent to main
scape evolution valleys (Matsushi and Matsukura, 2004). Thus, land-
slides do not erode the gentle (typically b208) upper
Landslides on the sandstone slope are controlled slopes, resulting in the relatively high hillcrests in the
whether wetting front reaches deep enough to form a sandstone area (Fig. 2).
failure plane. The highly permeable condition of the The subsurface water dynamics and frequency of
sandstone slopes disperses soil moisture and reduces landslide events on the mudstone slopes are in direct
Y. Matsushi et al. / Geomorphology 76 (2006) 92–108 107

opposition to these findings. The low permeability of area has then been eroded at a high rate, forming the
the mudstone bedrock permits generation of positive low and rugged landscapes.
pressure head that decreases shear resistance and pro-
motes the frequent landsliding identified in Fig. 3. The Acknowledgments
critical pressure heads for landsliding (~0.7 m) roughly
corresponds to thickness of the landslides (Table 1). We thank Dr. Thad Wasklewicz for his technical
This indicates that the maximum possible pressure head support to improving the manuscript. This research
causes a landslide of minimum thickness. Consequent- was financially supported by the Science Research
ly, the mudstone area has been eroded at a maximum Fund of the Ministry of Education, Science and Culture
possible rate maintaining only thin overburden on the (B-16300292, principal investigator: Y. Matsukura) and
hillslope (Fig. 5). The similar slope condition was the JSPS Research Fellowship for Young Scientists in
reported in the undisturbed tropical rainforests in Bru- 2003–2004 held by T. Hattanji.
nei, where intensive rainstorms often hit hillslopes
(Dykes, 2002). References
Rainfall often triggers landslides because of the low
permeability of the bedrock, provided that adequately Borga, M., Fontana, G.D., Gregoretti, C., Marchi, L., 2002. Assess-
thick soil has developed upon the bedrock. Convergent ment of shallow landsliding by using a physically based model of
hollows may accumulate sediment from surrounding hillslope stability. Hydrological Processes 16, 2833 – 2851.
ridges or side slopes, in addition to the in-situ disinte- Cho, S.E., Lee, S.R., 2001. Instability of unsaturated soil slopes due
to infiltration. Computers and Geotechnics 28, 185 – 208.
gration of the bedrock. Readying sufficient soil thick- Collins, B.D., Znidarcic, D., 2004. Stability analyses of rainfall
ness for a failure event, the hollows become the most induced landslides. Journal of Geotechnical and Geoenvironmen-
feasible locations for landsliding. Landslides on the tal Engineering 130, 362 – 372.
uppermost hollows probably spread their effect toward D’Odorico, P., Fagherazzi, S., Rigon, R., 2005. Potential for landslid-
ing: dependence on hyetograph characteristics. Journal of Geo-
the hillcrest, evolving the low and rugged hilly land-
physical Research 110, F01007. doi:10.1029/2004JF000127.
scapes (Figs. 1 and 2). Dykes, A.P., 2002. Weathering-limited rainfall-triggered shallow mass
movements in undisturbed steepland tropical rainforest. Geomor-
9. Conclusions phology 46, 73 – 93.
Fredlund, D.G., Raharjo, H., 1993. Soil Mechanics for Unsaturated
This study has focused on landslide mechanisms on Soils. John Wiley & Sons Inc, New York.
Fredlund, D.G., Morgenstern, N.R., Widger, R.A., 1978. The shear
hillslopes underlain by permeable sandstone and imper- strength of unsaturated soils. Canadian Geotechnical Journal 15,
meable mudstone in the Boso Peninsula, central Japan. 313 – 321.
Pressure head monitoring and rainfall–runoff observa- Furuya, T., Ohkura, H., 1992. Some geological and geomorphological
tion revealed contrasting hydrological processes in hill- characteristics of the slope failures on the Mt. Kano-zan and its
slopes with the different substrates. Slope stability environs in the Boso Peninsula, Japan. Journal of Japan Landslide
Society 28-4, 29 – 36 (in Japanese with English abstract).
analysis, including strength reduction of the slope ma- Gan, J.K.M., Fredlund, D.G., Rahardjo, H., 1988. Determination of
terial, was also used to specify the critical conditions for the shear strength parameters of an unsaturated soil using the
landsliding. direct shear test. Canadian Geotechnical Journal 25, 500 – 510.
In the case of hillslopes with permeable sandstone, Gasmo, J.M., Rahardjo, H., Leong, E.C., 2000. Infiltration effects on
stability of a residual soil slope. Computers and Geotechnics 26,
infiltrated rainwater percolates though the bedrock as
145 – 165.
an unsaturated gravitational flow. The reduction of soil Iida, T., 1999. A stochastic hydro-geomorphological model for shal-
cohesion resulting from wetting front migration causes low landsliding due to rainstorm. Catena 34, 293 – 313.
landslides on the steep lower parts of the hillslopes Ito, M., 1995. Volcanic ash layers facilitate high-resolution sequence
when deep soil becomes sufficiently wet to form a stratigraphy at convergent plate margins: an example from the
failure plane. This type of landsliding contributes to Plio-Pleistocene forearc basin fill in the Boso Peninsula, Japan.
Sedimentary Geology 95, 187 – 206.
the preservation of the relatively high hillcrests. Ito, M., 1998. Submarine fan sequences of the lower Kazusa Group, a
The impermeable mudstone, by contrast, allows oc- Plio-Pleistocene forearc basin fill in the Boso Peninsula, Japan.
currence of saturated subsurface storm flow draining Sedimentary Geology 122, 69 – 93.
incoming rainwater though a thin permeable soil layer Ito, M., Horikawa, K., 2000. Millennial- to decadal-scale fluctuation
upon the bedrock. The resulting decrease in the effec- in the paleo-Kuroshio Current documented in the Middle Pleisto-
cene shelf succession on the Boso Peninsula, Japan. Sedimentary
tive stress at the soil–bedrock boundary causes a land- Geology 137, 1 – 8.
slide. In this case, if soil is sufficiently thick, one can Iverson, R.M., 2000. Landslide triggering by rain infiltration. Water
often expect rainfall-triggered landslides. The mudstone Resources Research 36, 1897 – 1910.
108 Y. Matsushi et al. / Geomorphology 76 (2006) 92–108

Kaizuka, S., Koike, K., Endo, K., Yamazaki, H., Suzuki, T., 2000. and granite. Transactions, Japanese Geomorphological Union 15,
Regional geomorphology of the Japanese Islands. Geomorpholo- 49 – 65.
gy of Kanto and Izu-Ogasawara, vol. 4. University of Tokyo Onda, Y., Tsujimura, M., Tabuchi, H., 2004. The role of subsurface
Press, Tokyo (in Japanese). water flow paths on hillslope hydrological processes, landslides
Kim, J., Jeong, S., Park, S., Sharma, J., 2004. Influence of rainfall- and landform development in steep mountains of Japan. Hydro-
induced wetting on the stability of slopes in weathered soils. logical Processes 18, 637 – 650.
Engineering Geology 75, 251 – 262. Rao, S.M., 1996. Role of apparent cohesion in the stability of
Kosugi, K., 1994. Three-parameter lognormal distribution model for Dominican allophane soil slopes. Engineering Geology 43,
soil water retention. Water Resources Research 30, 891 – 901. 265 – 279.
Marquardt, D.W., 1963. An algorithm for least-squares estimation of Selby, M.J., 1993. Hillslope Materials and Processes, 2nd edition.
non-linear parameters. Journal of the Society for Industrial and Oxford University Press, Oxford.
Applied Mathematics 11, 431 – 441. Terlien, M.T.J., 1997. Hydrological landslide triggering in ash-
Matsushi, Y., Matsukura, Y., 2004. Mechanism and location of slope covered slopes of Manizales (Colombia). Geomorphology 20,
failures in hilly terranes with different bedrock permeability. 165 – 175.
Transactions, Japanese Geomorphological Union 25, 139 – 159 Vanapalli, S.K., Fredlund, D.G., Pufahl, D.E., Clifton, A.W., 1996.
(in Japanese with English abstract). Model for the prediction of shear strength with respect to soil
Montgomery, D.R., Dietrich, W.E., 1989. Source areas, drainage suction. Canadian Geotechnical Journal 33, 379 – 392.
density, and channel initiation. Water Resources Research 25, Van Asch, Th.W.J., Hendriks, M.R., Hessel, R., Rappange, F.E.,
1907 – 1918. 1996. Hydrological triggering conditions of landslides in
Montgomery, D.R., Dietrich, W.E., 1994. A physically based model varved clays in the French Alps. Engineering Geology 42,
for the topographic control on shallow landsliding. Water 239 – 251.
Resources Research 30, 1153 – 1171. Van Genuchten, M.Th., 1980. A closed-form equation for predicting
Mualem, Y., 1976. A new model for predicting the hydraulic con- the hydraulic conductivity of unsaturated soils. Soil Science So-
ductivity of unsaturated porous media. Water Resources Research ciety of America Journal 44, 892 – 898.
12, 513 – 522. Wakatsuki, T., Tanaka, Y., Matsukura, Y., 2005. Soil slips on
Onda, Y., 1992. Influence of water storage capacity in the regolith weathering-limited slopes underlain by coarse-grained granite
zone on hydrological characteristics, slope processes, and slope or fine-grained gneiss near Seoul, Republic of Korea. Catena
form. Zeitschrift für Geomorphologie N. F. 36, 165 – 178. 60, 181 – 203.
Onda, Y., 1994. Contrasting hydrological characteristics, slope pro- Wu, W., Sidle, R.C., 1995. A distributed slope stability model for
cesses and topography underlain by Paleozoic sedimentary rocks steep forested basins. Water Resources Research 31, 2097 – 2110.

You might also like