Open access
Technical Papers
Nov 10, 2020

Stability Monitoring of Embankments Constructed by Volcanic Coarse-Grained Soil in Snowy-Cold Regions

Publication: Journal of Cold Regions Engineering
Volume 35, Issue 1

Abstract

This study aims to identify the characteristics of a large-scale embankment constructed from volcanic soil in snowy-cold regions during all seasons and to confirm the validity of the prediction method for rainfall-induced failure that has been proposed in previous studies. To realize the aims, a series of laboratory-based model tests will be conducted initially on small-scale embankments constructed with a representative volcanic soil. Thereafter, a large-scale embankment was constructed in Sapporo City, Hokkaido, Japan. The same volcanic soil as that used in the model testing was utilized as the construction material. Taking into consideration the results of the field monitoring, model test, and stability analysis, the aspects of slope failure were clarified. It was shown that the rainfall-induced failure of an embankment could be uniquely evaluated by estimating the changes in water content in zones exposed to freezing and thawing. Finally, a stability monitoring protocol based on infiltration theory is discussed in detail.

Introduction

Geodisasters generated the most serious damage in volcanic ground and slopes in Hokkaido, Japan (Kawamura et al. 2019). It is known that the failures were caused by the specific performance and topography of volcanic soils defined as problematic soils. In addition, it has been reported that many cases of slope failure and landslide that were caused by rainfall and snowmelt occurred mainly from spring to summer, for example, the slope failure of a cut slope on the Hokkaido Expressway in 1999 (Kuromatsunai town, Hokkaido, Japan) and the landslide of a natural slope on the Kiritachi Mountain Pass on Route 239 in 2012 (Tomamae town, Hokkaido, Japan). In particular, the failure of a cut slope on the Hokkaido Expressway in 1999 was caused by rainfall and snowmelt after freezing and thawing. In this study, the authors focus on the rainfall-induced failure of embankments subjected to freezing and thawing among several patterns of slope failure.
A significant amount of research has been reported on the slope instability and the failure mechanism due to rainfall. Rahimi et al. (2011) clarified the effect of antecedent rainfall patterns on rainfall-induced slope failure. Kim and Lee (2010) monitored rainfall infiltration characteristics of a compacted roadside slope and evaluated this using examples from several one-dimensional infiltration models.
In addition, recent studies have provided some results on volcanic soils and slopes that have the particular characteristics of their mechanical behavior. Cascini et al. (2010) proposed a geotechnical model for the failure and postfailure stages of landslides of the flow type in the Campania region, Italy. Comegna et al. (2016) carried out a field hydrological monitoring of pyroclastic deposits for mountainous areas in the Campania region and investigated the seasonal variation in soil suction and water content.
However, although a number of geotechnical studies on issues caused by freezing and thawing and frost heaving have been conducted (e.g., Xu et al. 2011; Zhang and Hulsey 2015), in particular, analytical and in situ experimental approaches that investigated the instability of slopes by freezing and thawing are limited (e.g., Harries and Lewkowicz 2000; Yu et al. 2016; Zwissler et al. 2016; Flynn et al. 2016). Among this research, for example, Xu et al. (2011) investigated the effects of frozen soils on seismic site responses in Anchorage, Alaska. Zwissler et al. (2016) presented that the main control mechanism for slope recession in Calvert Cliffs was freeze–thaw events, and presented a method to quantify freeze–thaw effects.
The authors have investigated the failure of volcanic slopes induced by rainfall and its mechanisms in previous studies (e.g., Kawamura and Miura 2013; Kawamura and Miura 2014a, b). By considering of a series of model test results, a prediction method for slope stability was proposed based on the water retention characteristics (e.g., the changes in water content) of slopes consist of volcanic soils. However, it has been highlighted that the results obtained from a series of small-scale model tests might change with variations in inherent errors, such as the scale effect.
Thereafter, the authors conducted field monitoring from September 2011 to July 2012 on a large-scale embankment that was constructed to 5 m in height, 2 m in crown width, 4 m × 3 (=12 m) in length, and at a 45° angle by using a representative volcanic soil in Hokkaido, Japan. Details of the results of the large-scale embankment were reported by Kawamura et al. (2013).
This study aims to identify the characteristics of a large-scale embankment reconstructed by a representative volcanic soil in cold regions during all seasons and to confirm the validity of the prediction method that has been proposed in previous studies. Taking into consideration the results of the field monitoring, model test, and stability analysis, the aspects of slope failure were clarified. In addition, it was found that rainfall-induced failure of embankments could be uniquely evaluated by the changes in water content in zones exposed to freezing and thawing. Finally, a stability monitoring protocol is discussed based on Horton’s (1939) infiltration theory in detail.

Materials and Procedures of Small-Scale Tests

Test Materials

Volcanic coarse-grained soil that was sampled from the Shikotsu caldera in Hokkaido was used in this study (e.g., Kawamura and Miura 2013). The sampling site is shown in Fig. 1. This soil material is referred to as Komaoka volcanic soil. The physical properties and grain size distributions of the samples adopted for the construction of a large-scale embankment (typical two points) and model test are given in Table 1 and Fig. 2, and are compared with the characteristics of Toyoura sand. As presented in Table 1 and Fig. 2, the finer contents were from 26.0% to 42.6%. The fines of Komaoka volcanic soil was judged to be nonplastic material (NP) based on Atterberg limits (liquid limit = 44.5%, plastic limit = NP). Previously, volcanic soils, including Komaoka volcanic soil (Spfl), which have been utilized for residential embankments, have repeatedly liquefied due to the strong seismic loadings, for example, Tokachi-oki (1968 and 2003) and Hokkaido Eastern Iburi (2018), Japan.
Fig. 1. Locations of sampling and monitoring site of embankment.
Fig. 2. Grain size distributions of Komaoka volcanic soil for embankment and model test.
Table 1. Physical properties of Komaoka volcanic soil
Sample nameρs (g/cm3)ρd in situ (g/cm3)ρd max (g/cm3)ρd min (g/cm3)wn (%)D50 (mm)UcFc (%)
Komaoka2.501.120.76430.2743.026.0–42.6
Toyoura sand2.681.631.370.181.50
Note: wn = natural water content; D50 = mean grain size; Uc = coefficient of uniformity; and Fc = finer content.
Ongoing research focuses on the effect of the initial water content on the mechanical behavior of compacted volcanic soils (e.g., Matsumura et al. 2015; Kawamura and Miura 2014a). Fig. 3 shows the compaction curves of Komaoka volcanic soil that was obtained from the test method for soil compaction [A-c method (JGS 2009a) and JGS 0711-2009] and that of small-scale embankments (the dotted line) by another compacting method discussed in the following sections. The compactive effort was equal to 550 kJ/m3 for the A-c method. In the compaction test, the maximum dry density [ρdmax (g/cm3)] and the optimum water content [wopt (%)] of 1.059 g/cm3 and 40.5%, respectively, were obtained. A degree of compaction [Dc (%)] was estimated based on ρdmax = 1.059 g/cm3. Details of the mechanical behavior of Komaoka volcanic soil were described by Miura et al. (2003) and Matsumura et al. (2015).
Fig. 3. Compaction curves of Komaoka volcanic soil for A-c method, JGS, and model test.

Procedures of Small-Scale Model Tests

Rocha (1957) is known as a pioneer that considered the prospects of using models for the solution of soil mechanics problems based on the stress–strain relationship between prototype and model. The fundamental scaling (model/prototype) is given in Table 2. If the stress–strain relationship of soil material had a different scale in model and prototype, the relationships between stress (σp) and strain (ɛp) in the prototype, and stress (σm) and strain (ɛm) in the model are given as follows:
σmσp=1α
(1)
εmεp=1β
(2)
Table 2. Correlation between model and prototype
ParameterScale (model/prototype)
Length (L)Lm/Lp = 1/λ
Time (t)tm/tp = 1/τ
Density (ρ)ρm/ρp = 1/η
Gravity acceleration (g)gm/gp = 1/γ
Note: m = model; and p = prototype.
By synchronizing the ratios of force on influence factors (e.g., self-weight, inertia force, deformation modulus, damping constant, cohesion, and internal friction angle) for each other, physical modeling was obtained. In addition, if the stress–strain relationship is expressed by a hyperbolic model, the scaling of strain can be theoretically derived as follows:
1β=1λ0.5
(3)
The details of the procedure were described by Kokusho (2014). Scales of physical properties for 1 g model testing are summarized in Table 3 if the same material as that in the prototype was adopted (gm/gp = 1/γ = 1, ρm/ρp = 1/η = 1).
Table 3. Typical scaling of physical modeling for 1 g model test
ParameterScale (model/prototype)
Length (L)1/λ
Time (t)1/λ0.75
Density (ρ)1
Gravity acceleration (g)1
Stress (σ)1/λ
Strain (ɛ)1/λ0.5
Deformation (u)1/λ1.5
Deformation modulus (E)1/λ0.5
Friction (tanϕ)1
Cohesion (c)1/λ
Permeability (k)1/λ0.25
In this study, Komaoka volcanic coarse-grained soil was used as a test material. Fig. 4 shows the results of consolidated undrained triaxial tests (CU¯ tests) for Komaoka volcanic soils in terms of the relationship between the stress ratio, σ1/σ3 versus shear strain (γ). In the Fig. 4(b), shear strain for undrained triaxial tests was calculated as Poisson's ratio = 0.5. Therefore, an angle of internal friction of 38° and cohesion of 0 kN/m2 was estimated for CU¯ tests. The stress–strain relationship was affected by the difference in the effective confining pressure [Fig. 4(a)]. However, the relationship can be expressed approximately by a hyperbolic one under the square root of σc, as shown by a dotted line [Fig. 4(b)]. In this study; therefore, it is hypothesized that shear strength of soil (τ) varies proportionally by normalizing with effective confining pressure (σc) based on the previous discussions (e.g., Kawamura and Miura 2013, 2014a). Therefore, simulation by model testing was theoretically enabled for cohesionless soil materials, such as Komaoka volcanic soils, if the infiltration speed and rainfall intensity in the model were consistent with those in the field (i.e., permeability of Komaoka volcanic soil had a high value of >10−5 m/s). In addition, the effect of suction on mechanical behavior until failure was small, because the degree of saturation was >60% in test ranges, as discussed in the following sections.
Fig. 4. Results of consolidated undrained triaxial tests (CU¯ tests) for Komaoka volcanic soils: (a) σ1σ3γ relationship in triaxial test; and (b) σ1σ3γ(σc)0.5 relationship in modeling.
Fig. 5 shows an overall view of the apparatus used in rainfall testing. The soil container was 2,000 mm in length, 700 mm in depth, and 600 mm in width. Model embankments were prepared by compacting (weight of roller = 127.4 N, the roller was reciprocated 4 times/layer of 70 mm). Variation of the single model in dry density was within 5%. Particle crushing of soil particles did not occur by compactions under the initial water contents. Thereafter, the embankment surface was carefully trimmed to a 45° angle (relative to horizontal) using a straight edge to eliminate surface disturbance. To appropriately simulate freezing and thawing, the surface of the embankment was frozen with dry ice over 8 h and was then thawed at 20°C (over a basic thawing period of 8 h) after the model embankment was set up. In this procedure, frozen layers of approximately 50 mm thickness were visually observed for the model embankment, as shown in the following sections. Slope shapes with 45° and positions of measurement instruments and test conditions are given in Fig. 6 and Table 4, respectively.
Fig. 5. Overall view of apparatus.
Fig. 6. Model shapes and setting positions of measurement instruments: (a) without freeze-thaw action; and (b) with freeze-thaw action.
Table 4. Test conditions for this study
Slope conditionNo freeze–thaw actionFreeze–thaw action
Slope angle (°)4545
Length of base [B (mm)]750750
Initial water content (%)38, 4338, 43
Dry density [ρd (g/cm3)]0.900.90
Rainfall intensity [R (mm/h)]100100
Freeze–thaw action cycles1
A rainfall intensity of 100 mm/h was adopted and was accurately simulated using a spray nozzle. During the rainfall testing, changes in deformation behavior, saturation degree, and temperature were monitored using digital video cameras, soil moisture meters, and thermocouple sensors, respectively. Specifically, deformation behavior was estimated based on particle image velocimetry (PIV) analysis (White et al. 2003). Pore water pressure was monitored simultaneously. Operative range (rated capacity) of transducers of pore water pressure = 50 kPa. In this study, it was regarded as the plastic equilibrium state when a shear strain by PIV analysis was 4%–6%, specifically, the mechanical behavior at slope failure (e.g., Kawamura and Miura 2013). The rainfall model testing was performed for 3 h, or until the slope failure was induced. In this study, each test was performed three times.

Test Results of Small-Scale Embankments and Discussions

Figs. 7(a and b) show the slip lines of failed slopes and the shear strain distribution at failure for embankments constructed using Komaoka volcanic soil (Kawamura and Miura 2014a), respectively. For the case with a low water content of w0 = 38% [Fig. 7(a)], it was shown that Slip line 1 (the first failure) was triggered around the toe of embankments, and then Slip line 2 (the final failure) rapidly developed with the increase of weight due to rainfall and of pore water pressure around the toe of the embankments, as shown in following sections. Therefore, the slope failed at 3,650 s.
Fig. 7. Slip lines of failed slopes and shear strain distributions at failure: (a) w0 = 38%; and (b) w0 = 43%.
In the case of a high water content of w0 = 43%, surface flow due to erosion proceeded until the Slip line shown in Fig. 7(b), and the elapsed time reached the same depth as Slip line 2 for w0 = 38%, which was 9,000 s. As shown in Fig. 3, the optimum water content of Komaoka volcanic soil was 40.5%. The failure mode differs for w0 = 38% and 43% based on the developments of the saturation degree and in the excess pore water pressure (Figs. 8 and 9). For instance, pore water pressure for w0 = 38% increased rapidly compared with that for w0 = 43%. Due to the changes in the initial water content, a surface failure with a circular slip was caused for w0 = 38%, and a surface flow failure was induced for w0 = 43%. This was because the permeability of compacted soils generally decreased with an increase in water content and indicates the minimum value for higher water content over the optimum water content. Matsumura (2014) confirmed this reduction of permeability with increasing water content for compacted Komaoka volcanic soils, although it was presumed that the effect of particle breakage on permeability was reflected in the results. By considering the test results, slope failure appears to be induced by the expansion of areas with high water retention and the increase of pore water pressure around the toe of embankments.
Fig. 8. Changes in saturation degree with elapsed time: (a) w0 = 38%; and (b) w0 = 43%.
Fig. 9. Changes in pore water pressure with elapsed time: (a) w0 = 38%; and (b) w0 = 43%.
A series of rainfall model tests with the freezing and thawing was conducted (e.g., Kawamura and Miura 2014a, b). During freezing and thawing, the slope surface of the model was frozen with dry ice over 8 h, and was thawed at a temperature of 20°C over 8 h after the model was set up; subsequently, a model test with rainfall continued until the model embankment failed. Representative changes in temperature (T1–T9) in a model embankment during freezing and thawing are shown in Fig. 10. This figure shows that the temperatures decreased during freezing action (until approximately 28,800 s), and then increased over time. Specifically, temperatures at the depth of 25 mm (T1, T4, and T7) indicated approximately < 0°C. A freezing depth of approximately 50 mm was visually observed in model tests.
Fig. 10. Representative variations in temperature during freezing and thawing.
Variations in the degree of saturation for model embankments composed of Komaoka volcanic soil are shown in Fig. 11. By comparing the cases without freezing and thawing (Fig. 8), no difference was observed in the behavior of saturation between both cases, but the difference in elapsed time to reach failure was observed. For instance, the elapsed time until failure for the case of freezing and thawing for w0 = 43% was 765 s, which collapsed approximately 3.5 times faster than the case without freezing and thawing (2,690 s). To clarify the influence of freeze–thaw action on the deformation behavior of model embankments, Figs. 12(a and b) show the failed shape and deformation behavior (shear strain distributions) with freezing and thawing for w0 = 43%. In comparison with cases not subjected to freezing and thawing [Figs. 7(b) and 12], the shape at the final slip line differed for each test case. For instance, the slip line depth with freezing and thawing was shallower than that without. The result means that hollows due to freezing and thawing produce loose structures in the frozen layer compared with before the freeze–thaw action. The difference in deformation behavior by the changes in the initial water content was not confirmed for Komaoka volcanic embankments with freezing and thawing. A similar tendency was observed for other volcanic soils (Kawamura and Miura 2013, 2014b).
Fig. 11. Changes in saturation degree with elapsed time for the cases subjected to freezing and thawing: (a) w0 = 38%; and (b) w0 = 43%.
Fig. 12. Failed shape and shear strain of model embankment with freeze–thaw action: (a) failed shape; and (b) distribution of shear strain.
To confirm this behavior, comparisons of the frozen layer (Df) and the depths of the final slip line (Ds) measured in the model test are shown in Fig. 13. In this figure, test data was additionally plotted as reported by Kawamura and Miura (2014b) and illustrates that both depths were approximately the same for the model test, as shown by a dotted line. In addition, Kawamura and Miura (2013) clarified that rainfall-induced failure of unsaturated slopes formed from crushable particles was derived from loose structures generated by freezing and thawing and by the reduction of shearing resistance due to particle crushing during freezing and thawing. Of note, the influence of particle breakage on slope stability was small for Komaoka volcanic soil. Kawamura et al. (2013, 2016) indicated that model and large-scale embankments subjected to freezing and thawing for one cycle (one season) have already been in the plastic equilibrium. Specifically, normal displacements on the frozen surface were 5.4 mm (normal strain ɛ = 5.4%) for w0 = 43% and 1 mm for w0 = 37% (ɛ = 1.0%) in model tests. Converting these values into shear strain results in the range of γ = 1.5%–8.3%, which means that the embankment in high water content becomes the plastic equilibrium at one cycle of freeze–thaw actions. In addition, Matsumura et al. (2015) elucidated cyclic mechanical behavior of compacted Komaoka volcanic soils, and that cyclic strengths of specimens with a freeze–thaw sequence were lower than those with no freeze–thaw sequence. By taking into consideration the results, it was highlighted that areas of embankments affected by freezing and thawing for one cycle have already been in plastic equilibrium; therefore, estimation of an influenced area subjected to freezing and thawing is important for slope stability in cold regions.
Fig. 13. Comparisons of the depths of frozen layer (Df ) with slip line (Ds) after rainfall test.

Prediction Method of Surface Slope Failure

In this study, a prediction method that considers the water retention characteristics for surface slope failure of such embankments was proposed. Fig. 14 shows the soil–water characteristics curve of Komaoka volcanic soil examined by test method for water retentivity of soils according to JGS 0150-2009 (JGS 2009b). In Fig. 14, a fitted curve was indicated by the van Genuchten model (1980). A test specimen was compacted at water content w = 38.2% and the degree of compaction Dc = 88.0% (Matsumura 2014). It was recognized that the specimen was drastically saturated at <10 kPa in suction. In the model tests, the initial water content of w0 = 38%–43% corresponded to the degree of saturation of Sr = 53.4%–60.4%. The suction value (s) was <5 kPa when the water content was >w = 39.8%. Therefore, the contribution of suction on shear strength appears to be very small for Komaoka compacted soils. Due to these factors, the authors focused mainly on hydraulic mechanical behavior at failure for the following discussion.
Fig. 14. Soil–water characteristics curve of Komaoka volcanic soil.
It is well-known that Horton (1939) investigated infiltration phenomena in the ground during rainfall and proposed an empirical expression. In this model testing, a similar tendency where the degree of saturation increases with rainfall and converges at a certain value was observed (Fig. 8). For example, the following equation can be expressed by substituting the degree of saturation with water content based on Fig. 8 as follows:
w=wc+(w0wc)ent
(4)
where w = water content (%) at time; w0 = initial water content at time t = 0; wc = final water content; and n = a constant for a given curve for the water content–time relationship. For example, the parameters w0 and wc, and n for the data of sm1 in Fig. 8(a) is 40.7% and 60.7%, and 0.001, respectively. Therefore, changes in water content, which could be easily measured, were available for stability evaluation.
To estimate water content at failure (wf) the relationships between water content at the initial state (w0) and the failure (wf) (=final water content wc) are shown in Fig. 15. In this figure, all test data are additionally depicted (including those from previous studies). The water retention capacity was evaluated as water content on the slip surface. From Fig. 15, it can be seen that there are unique relationships between both water contents. Their relationships can be estimated as follows:
wf=βw0γ
(5)
where β and γ = coefficients of a given curve whose values are shown in Fig. 15, where it can be seen that the water content at failure decreased for the case with freezing and thawing.
Fig. 15. Relationship between water content at initial condition and the failure for Komaoka volcanic soil.

Field Monitoring of Large-Scale Embankment Exposed to Freezing and Thawing

By examining the previous results, the authors conducted field monitoring from November 2012 to November 2013 for a large-scale embankment reconstructed to a height of 5 m, width 2.7 m, length 4 m, and an angle of 45° using the same soil as that of the small-scale model tests.
The field monitoring site was located in Sapporo, Japan, as shown in Fig. 1. A large-scale embankment was supported by lateral confining plates (the right and left sides), covered with plastic sheets (back side and bottom), and was thereafter built using a hand-guide roller (weight = 5.88 kN) by compacting Komaoka volcanic soil to achieve more than the degree of compaction (Dc) of 85% for each layer of 0.25 m. According to in situ density and water content tests, the compaction degree (on average) and water content (on average) were Dcave = 95.5% and wave = 42.5%, respectively. Details of the construction procedure are described by Matsumura (2014) and Kawamura et al. (2015).
The following instruments were installed to monitor soil behavior and temperature in the air and the embankment: (1) soil moisture meter [Time Domain Reflectometry type (TDR)], (2) tensiometer, (3) thermometer, (4) rainfall gauge, (5) snow gauge, (6) accelerometer, and (7) Anemovane, as shown in Fig. 16. The soil moisture meters and thermocouple sensors were set up at interval depths of 0.2 m until a depth of 1.5 m. The specifications of the instruments are reported in Kawamura et al. (2013). All data were collected within the sampling period of <10 min. In this study, volumetric water content (θ) from the TDR devices was converted by water content (w) by the relation w = (ρwd in situ) θ, where ρw is the density of water and ρd in situ is the in situ dry density of soil respectively.
Fig. 16. Setting positions of monitoring instruments: (a) whole view; (b) C section; and (c) L section and R section.
After the construction of the embankment, 10 pipes for water supply (Fig. 17) and 3 tanks were set in order to follow infiltration and freezing and thawing under climatic conditions and to cause slope failure. The tanks were located 3 m higher than the embankment crown, and the amount of water was controlled by each valve equipped with the tanks (Kawamura et al. 2013). The amount of water supply was approximately 250 L/day from November 20, 2012, to November 30, 2012. Thereafter, it was gradually increased to approximately 1,000–3,000 L/day (from May 7 to November 15, 2013). Water supply stopped for the winter months (from December 2012 to April 2013), and stability monitoring continued from November 1, 2012, to November 15, 2013. During monitoring, the effective responses of acceleration due to earthquakes were not observed. Therefore, the data from the accelerometers were omitted from this paper.
Fig. 17. Locations of water supply pipes: (a) plane view; and (b) side view of embankment.
Fig. 18 shows variations in temperature in the embankment from November 1, 2012, to November 1, 2013. During monitoring, snow removal was carried out during the winter season (from December to March 2013). In Fig. 18, temperature for each depth varies daily; in particular, the value at the depth range of 0–0.2 m was <0°C from December 2012 to March 2013. Therefore, it was confirmed that the embankment at this location would be affected by freezing and thawing, because Komaoka volcanic soil contains a significant amount of fine particles, such as silt.
Fig. 18. Variations in temperature in the embankment during monitoring.
After the winter season, water supply resumed from May 7, 2013. During monitoring, surface failure with a large deformation was induced on October 17–18, 2013 due to the increase of rainfall and water supply. In subsequent investigations, the characteristics of the embankment slope before failure was investigated (Kawamura et al. 2015).
Figs. 19(a and b) show changes in the aspect of the slope surface with elapsed days based on surveying. For example, the toe of the slope was eroded on August 1, 2013, its area gradually extending proceeding slope failure due to the increase of water supply. Fig. 19(b) shows the slope shape after failure on October 19, 2013. According to field investigations, the depth of the failed area was approximately in the range of 0.6–0.9 m. The failed soil volume on November 15, 2013 was estimated to be approximately 7.5 m3.
Fig. 19. Changes in aspect of slope surface with elapsed days and locations of sensors: (a) August 1, 2013; and (b) October 19, 2013.
Figs. 20(a and b) show changes in water content for SMR2 and SMR3 sensors in the failed area during monitoring (Fig. 19). It can be seen that the water content rapidly increased with the increase of rainfall and water supply, and decreased at slope failure; the depths of 0.2 and 0.4 m were particularly significant. Therefore, it was assumed that slip lines were generated from a depth of 0.2–0.4 m. Due to the development of slip lines, the failed area appeared to become deeper (to a depth of 0.6–0.9 m). Therefore, the change in water content is a key factor to evaluate slope stability.
Fig. 20. Changes in water content for SMR2 and SMR3 in R section during monitoring and its locations: (a) SMR2 sensor; and (b) SMR3 sensor.
Figs. 21(a and b) shows the behavior of pore water pressure measured by the TC2 sensor until slope failure. This figure indicates that pore pressure was a negative value (−2 to −5 kPa) and was constant immediately before slope failure and that it sharply increased at failure; thereafter it decreased. This indicated that monitoring of pore water pressure using a tensiometer might be one method to predict surface failure for unsaturated slopes, similar to the change in water content. Details of data on other monitoring devices are described by Matsumura (2014).
Fig. 21. Changes in pore water pressure for TC2 sensor in C section during monitoring: (a) from September 18 to November 15, (2 months); and (b) from October 17 to 18 (1 day).
Compared with model test results, the failed shape and the development of saturation degree indicated a similar tendency to that of embankments subjected to freezing and thawing. For instance, the scale (model/prototype) of length was 1/λ [=0.14(=0.7 m/5.0 m) in height] in this study. Based on the slip line depth of 51 mm that was obtained for the case of w0 = 43% with the freezing and thawing (Fig. 13), the corresponding depth converted to this in the prototype (the large-scale embankment) was 365 mm. Therefore, the model test result explained the field data if the slip line was induced at the depth of 0.37 m in the field. In addition, the development of saturation was approximately the same compared with the corresponding depths [the sensor of sm3 at the depth of 25 mm (Fig. 11(b))] and the sensor of SMR2 or SMR3 at the depth of 0.2 m). However, transducers for pore water pressure could not be installed around the slip line under freeze–thaw action due to the capacity of sensors; however, it was presumed that the behavior of pore water pressure had a similar tendency as the large-scale embankment, although the difference in infiltration process of supply water was highlighted. However, it is recognized that the evaluation of the mechanical behavior of a large-scale embankment from the results of 1 g small-scale model tests is difficult in practice. In particular, because the corresponding points between a large-scale and a small-scale embankment were shallower zones in this study, the confining pressure effect on the mechanical behavior might be ignored. Further discussions are required in this area of research.
Based on the failure situation, the analysis was performed using the simplified Janbu method (Ugai 1987). Figs. 22(a–c) show a numerical model of the large-scale embankment at failure and models of representative distributions of water content on September 18 and October 17, 2013. In this study, the slope stability was evaluated by taking changes in water content (unit weight) into account, because the suction effect was very small when the water content was >w = 40% (Sr = 56%). The boundaries of both sides, the back side, and bottom were fixed and were undrained conditions.
Fig. 22. Numerical models of the large-scale embankment at failure and distribution of water content: (a) a numerical model; (b) a model on September 18 (L section); and (c) a model on October 17 (L section).
Fig. 23 shows a variation in the safety factor (Fs) on the slip line with elapsed days. Numerical parameters adopted in this study are shown in this figure. It was found that the safety factor decreased daily and that it was <Fs = 1 on September 28, 2013. By considering the slope failure data, the embankment might have destabilized on September 28, 2013. The analytical results correspond well with the mechanical behavior at failure. Water content at the initial state (at the construction) (w0) normalized by that at the failure (wf, w0/wf ) was approximately 80% for this case. If approximately 20% of the increase in water content could be monitored for this case, stability evaluation should be performed.
Fig. 23. Changes in safety factor with elapsed days using the simplified Janbu method.
Soil samples were collected from the zones on the slip line to investigate water content at failure. The water content on the slip line area was 54% on average. Based on the previous results for initial water content at construction and failure, the validity of Eq. (5) was confirmed (Fig. 15). It is noted that the field data of w = 54% (see star symbols) was approximately the failure line of freeze–thaw action for Komaoka volcanic soil. The result obtained from the field data might be expressed as an underestimated value, because there is a difference in depth between the slip line estimated by the behavior of soil moisture meters (=0.2–0.4 m) and the frozen depth (=0.2 m) estimated by thermocouple sensors as aforementioned; however, it was confirmed that the estimated value was in the range from no freeze–thaw action to freeze–thaw action in Eq. (5). Therefore, Eq. (5) adequately explains the field data for the embankment constructed from volcanic coarse-grained soil.
Therefore, the risk evaluation of embankments could be enabled by the proposed Eqs. (4) and (5) based on the changes in water content if the threshold of slope failure could be prescribed. For example, if the line of water content (wf) potentially causing slope failure was estimated as shown in Fig. 15 and the depth of the frozen area (including frost heaving), the slope stability could be evaluated based on the change in water content measured by instruments, such as soil moisture meters.

Conclusions

Based on the field monitoring of a large-scale embankment, the model tests, and the numerical analysis conducted in this study, the following conclusions were obtained.
1.
Compared with model and field test results, the failed shape and the development of saturation degree in the field indicated a similar tendency with that of model embankments subjected to freeze–thaw action. This implied that simulation by model testing was enabled for cohesionless soils, such as Komaoka volcanic soils if infiltration speed and rainfall intensity in the model were consistent with those in the field.
2.
The depths of the frozen layer (Df) and slip line (Ds) measured were approximately the same in the model or large-scale embankments. Therefore, areas of embankments affected by freezing and thawing were already in plastic equilibrium. In addition, the effect of freezing and thawing for the first cycle of failure mechanism was significant for a series of our research. From the results, if an influenced zone of freezing and thawing was assessed, surface slope failure could be evaluated.
3.
Risk evaluation of embankments could be enabled by the proposed method based on the changes in water content if the threshold of slope failure was prescribed. For example, if the line of water content (wf) that potentially caused slope failure was estimated (as shown in Fig. 15) and the depth of the frozen area (including frost heaving), slope stability could be evaluated based on the change in water content measured by instruments, such as soil moisture meters.

Acknowledgments

The authors wish to express their sincere gratitude to Ms. A. Kudo (Hokkaido University) and Mr. H. Nakagawa and Mr. K. Fukutsu (Muroran Institute of Technology) who conducted a major part of field monitoring, experiments and stability analysis, and Dr. S. Yokohama (Hokkaido University). This study was undertaken with the financial supports of KAKENHI [Grant-in-Aid for Science Research (A) No. 23241056, (C) No. 24560597 and (B) 20H02404], Japan Society for the Promotion Science.

References

Cascini, L., S. Cusmo, M. Pastor, and G. Sorbino. 2010. “Modeling of rainfall-induced shallow landslides of the flow-type.” J. Geotech. Geoenviron. Eng. 136 (1): 85–98. https://doi.org/10.1061/(ASCE)GT.1943-5606.0000182.
Comegna, L., E. Damiano, R. Greco, A. Guida, L. Olivares, and L. Picarelli. 2016. “Field hydrological monitoring of a sloping shallow pyroclastic deposit.” Can. Geotech. J. 53 (3): 1115–1137. https://doi.org/10.1139/cgj-2015-0344.
Flynn, D., D. Kurz, M. Alfaro, J. Graham, and U. L. Arenson. 2016. “Forecasting ground temperatures under a highway embankment on degrading permafrost.” J. Cold Reg. Eng. 30 (3): 04016002. https://doi.org/10.1061/(ASCE)CR.1943-5495.0000106.
Harries, C., and A. G. Lewkowicz. 2000. “An analysis of the stability of thawing slopes, Ellesmere Island, Nunavut, Canada.” Can. Geotech. J. 37 (2): 449–462. https://doi.org/10.1139/t99-118.
Horton, E. R. 1939. “Analysis of runoff-plat experiments with varying infiltration-capacity.” Trans. Am. Geophys. Union 20 (4): 693–711. https://doi.org/10.1029/TR020i004p00693.
JGS (Japanese Geotechnical Society). 2009a. Test method for soil compaction using a rammer. JGS 0711-2009. Tokyo: JGS.
JGS (Japanese Geotechnical Society). 2009b. Test method for water retentivity of soils. JGS 0150-2009. Tokyo: JGS.
Kawamura, S., S. Kawajiri, W. Hirose, and T. Watanabe. 2019. “Slope failures/landslides over a wide area in the 2018 Hokkaido Eastern Iburi earthquake.” Soils Found. 59 (6): 2376–2395. https://doi.org/10.1016/j.sandf.2019.08.009.
Kawamura, S., and S. Miura. 2013. “Rainfall-induced failures of volcanic slopes subjected to freezing and thawing.” Soils Found. 53 (3): 443–461. https://doi.org/10.1016/j.sandf.2013.04.006.
Kawamura, S., and S. Miura. 2014a. “Failure of volcanic slopes in cold regions and its prediction.” Procedia Earth Planet. Sci. 9: 143–152. https://doi.org/10.1016/j.proeps.2014.06.010.
Kawamura, S., and S. Miura. 2014b. “Stability of volcanic slopes in cold regions.” J. Geogr. Geol. 6 (3): 34–54. https://doi.org/10.5539/jgg.v6n3p34.
Kawamura, S., S. Miura, M. H. Dao, and R. Yamada. 2016. “Rainfall-induced failure of volcanic embankments subjected to cyclic loadings in cold regions.” In Geo-China 2016: Advances in Numerical and Experimental Analysis of Transportation Geomaterials and Geosystems for Sustainable Infrastructure, Geotechnical Special Publication 257, edited by R. Bulut, X. Yu, and S.-R. Yang, 116–123. Reston, VA: ASCE.
Kawamura, S., S. Miura, and S. Matsumura. 2015. “Stability evaluation of full-scale embankment constructed by volcanic soil in cold regions.” Jpn. Geotech. Soc. Spec. Publ. 2 (26): 971–976. https://doi.org/10.3208/jgssp.JPN-063.
Kawamura, S., S. Miura, S. Yokohama, A. Kudo, and N. Kaiya. 2013. “Field monitoring of embankment constructed by volcanic soil and its evaluation.” In Geo-Congress 2013: Stability and Performance of Slopes and Embankments III, Geotechnical Special Publication 231, edited by C. Meehan, D. Pradel, M. A. Pando, and J. F. Labuz, 373–382. Reston, VA: ASCE.
Kim, K. Y., and R. S. Lee. 2010. “Field infiltration characteristics of natural rainfall in compacted roadside slopes.” J. Geotech. Geoenviron. Eng. 136 (1): 248–252. https://doi.org/10.1061/(ASCE)GT.1943-5606.0000160.
Kokusho, T. 2014. Fundamentals of earthquake geo-dynamics. [In Japanese.] Tokyo: Kajima Institute Publishing Co. Ltd.
Matsumura, S. 2014. “Laboratory and in-situ studies on mechanical properties of volcanic soil embankment in cold region.” Doctoral dissertation, Graduate School of Engineering, Hokkaido Univ.
Matsumura, S., S. Miura, S. Yokohama, and S. Kawamura. 2015. “Cyclic deformation-strength evaluation of compacted volcanic soil subjected to freeze-thaw sequence.” Soils Found. 55 (1): 86–98. https://doi.org/10.1016/j.sandf.2014.12.007.
Miura, S., K. Yagi, and T. Asonuma. 2003. “Deformation-strength evaluation of crushable volcanic soils by laboratory and in-situ testing.” Soils Found. 43 (4): 47–57. https://doi.org/10.3208/sandf.43.4_47.
Rahimi, A., H. Rahardjo, and L. Eng-Choon. 2011. “Effect of antecedent rainfall patterns on rainfall-induced slope failure.” J. Geotech. Geoenviron. Eng. 137 (5): 483–491. https://doi.org/10.1061/(ASCE)GT.1943-5606.0000451.
Rocha, M. 1957. “The possibility of solving soil mechanics problems by the use of models.” In Vol. 1 of Proc., 4th Int. Conf. on Soil Mechanics and Foundation Engineering, 183–188. London, UK: Butterworths Scientific Publications.
Ugai, K. 1987. “Three-dimensional slope stability analysis by simplified Janbu method.” [In Japanese.] Landslides 24 (3): 8–14. https://doi.org/10.3313/jls1964.24.3_8.
van Genuchten, M. T. 1980. “A closed-form equation for predicting the hydraulic conductivity of unsaturated soils.” Soil Sci. Soc. Am. J. 44 (5): 892–898. https://doi.org/10.2136/sssaj1980.03615995004400050002x.
White, D. G., W. A. Take, and M. D. Bolton. 2003. “Soil deformation measurement using particle image velocimetry (PIV) and photogrammetry.” Géotechnique 53 (7): 619–631. https://doi.org/10.1680/geot.2003.53.7.619.
Xu, G., Z. Yang, U. Dutta, L. Tang, and E. Maex. 2011. “Seasonally frozen soil effects on the seismic site response.” J. Cold Reg. Eng. 25 (2): 53–70. https://doi.org/10.1061/(ASCE)CR.1943-5495.0000022.
Yu, W., F. Han, X. Yi, W. Liu, and D. Hu. 2016. “Cut-slope icing prevention: Case study of the seasonal frozen area of Western China.” J. Cold Reg. Eng. 30 (3): 05016001. https://doi.org/10.1061/(ASCE)CR.1943-5495.0000105.
Zhang, Y., and L. Hulsey. 2015. “Temperature and freeze-thaw effects on dynamic properties of fine-grained soils.” J. Cold Reg. Eng. 29 (2): 04014012. https://doi.org/10.1061/(ASCE)CR.1943-5495.0000077.
Zwissler, B., T. Oommen, and S. Vitton. 2016. “Method to quantify freeze-thaw effects on temperate climate soils: Calvert Cliffs.” J. Cold Reg. Eng. 30 (3): 06016002. https://doi.org/10.1061/(ASCE)CR.1943-5495.0000103.

Information & Authors

Information

Published In

Go to Journal of Cold Regions Engineering
Journal of Cold Regions Engineering
Volume 35Issue 1March 2021

History

Received: Nov 16, 2016
Accepted: Aug 12, 2020
Published online: Nov 10, 2020
Published in print: Mar 1, 2021
Discussion open until: Apr 10, 2021

Authors

Affiliations

Shima Kawamura, A.M.ASCE [email protected]
Dr. Eng., Professor, Graduate School of Engineering, Muroran Institute of Technology, 27-1 Mizumoto-cho, Muroran 050-8585, Japan (corresponding author). Email: [email protected]
Seiichi Miura [email protected]
Dr. Eng., Emeritus Professor, Hokkaido Univ., Kita 13 Nishi 8, Kita-ku, Sapporo 060-8264, Japan. Email: [email protected]
Satoshi Matsumura [email protected]
Dr. Eng., Senior Researcher, Dept. of Geotechnical Engineering, Port and Airport Research Institute, 3-1-1 Nagase, Yokosuka 239-0826, Japan. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share