Open access
Technical Papers
Oct 7, 2014

Design of Stable Concave Slopes for Reduced Sediment Delivery

Publication: Journal of Geotechnical and Geoenvironmental Engineering
Volume 141, Issue 2

Abstract

Constructed slopes are traditionally given a planar form. However, natural slopes are more likely to be concave in cross section. In addition, laboratory and computational studies have demonstrated that concave slopes yield less sediment than planar slopes. With current autoguided construction equipment, it is now possible to construct slopes with concave profiles and a more natural appearance, yet a simple method to describe such concave slopes for a given level of mechanical stability does not exist. This article begins with an examination of concave shapes satisfying a desired degree of stability and compares results with those from the FEM and limit-equilibrium method of analysis. An erosion model is used to demonstrate that the concave slopes proposed here yield 15–40% less sediment than planar slopes with the same factor of safety. Finally, a sensitivity analysis suggests that reasonable construction deviations do not compromise the stability of typical concave slopes.

Introduction

Constructed slopes designed with traditional planar cross sections are encountered in most land development, including highway cut and fill sections, constructed embankments, and reclaimed mine lands. However, planar landscape profiles are seldom encountered in nature. Curvilinear slopes with concave shapes usually arise as the result of evolutionary processes in fluvial systems and hillslopes (Leopold and Langbein 1962; Miyamoto et al. 2005; Rodríguez-Iturbe et al. 1992; Twidale 2007; Yang and Song 1979). Landforming approaches such as the geomorphic reclamation of mine lands (Toy and Chuse 2005) include the construction of concave shapes in both the transverse (cross-slope) and longitudinal (downslope) directions to create natural self-sustainable ecosystems (Martín-Duque et al. 2010) with improved erosion resistance (Schor and Gray 2007). Nevertheless, not all concave shapes are mechanically stable, requiring a rational definition of those concave slopes that provide the desired degree of stability expressed in terms of a desired factor of safety (FS). For example, Howard et al. (2011) point out the risk associated with the growing practice of shaping slopes to reflect natural regional landforms without appropriate stability and erosion analyses and without accounting for the limited precision of the construction equipment employed to build concave profiles. Hence, the objectives of this work are to (1) describe concave shapes that provide a desired degree of stability (a desired FS) for given soil properties, (2) provide a quantitative measure of the difference in soil loss (erosion) between concave and planar slopes that satisfy the same degree of mechanical stability, and (3) investigate the precision to which concave forms can be constructed and how this affects the desired slope stability. To accomplish these objectives, this study focuses on (1) materials that contain some fine-grained (silt and clay) particles and thus at least a small measure of cohesion and (2) slopes with concavity in the longitudinal direction, which have been shown to deliver less sediment than transversely convex, transversely concave, longitudinally convex, and planar slopes (Rieke-Zapp and Nearing 2005).

Background

Mechanical Stability of Concave Profiles

Although a number of studies have shown that concave slopes produce less sediment than planar slopes, few studies have linked the benefits of concave slopes to mass stability. Utili and Nova (2007) demonstrated, by means of the upper-bound method, that concave log-spiral slopes result in higher FS values than planar slopes of the same height and width, as well as planar slopes of the same soil mass. Based on the slip-line field theory and the associated characteristic equations, Sokolovskiĭ (1960) found a longitudinal concave critical slope contour (FS=1) for plane-strain conditions. Sokolovskiĭ’s solution is the result of a numeric boundary value problem, which for a medium with unit weight γ>0, cohesion c, and friction angle ϕ>0 has no analytical counterpart. Recently, Jeldes et al. (2013) proposed an approximate analytical expression to obtain coordinates of the critical contour for any desired combination of ϕ, c, γ, and slope height (Hs). This expression brings a system of differential equations into a single algebraic expression, which highly simplifies the manipulation and solution as
x={0,hcry0A[σy(B1)(cscϕ1)+HB(cscϕ+1)],y>0
(1)
where
A=cosϕ2γ(1sinϕ)
(2)
B=ln[σyH(1sinϕ1+sinϕ)+1]=ln(σyHKa+1)
(3)
σy=γy
(4)
H=ccotϕ
(5)
hcr=2ccosϕγ(1sinϕ)
(6)
where H = tensile strength of the soil; Ka=(1sinϕ)/(1+sinϕ) is the Rankine active coefficient of earth pressure; σy = geostatic vertical stress; and hcr is the height of the tension zone. The equation describes a slope contour in the quadrant with the x-axis positive to the right and the y-axis positive downward (Fig. 1), and with hcr lying above the x-axis from the coordinates (0, 0) to (0, hcr). Note that this hcr tension zone does not contribute to resistance, but only to destabilization; this is an artifact used to obtain a complete mathematical solution of the problem. Whereas this assumption influences the way the characteristics would develop inside the soil medium, it was shown via the limit-equilibrium method (LEM) and FEM analyses that the outcomes are not significantly affected by this assumption (Jeldes et al. 2013). In practice, the sharp bluff at the top of the tension zone may be removed or rounded (as indicated by the dashed line in Fig. 1), adding conservatism to the slope design. A limiting slope height (hL) was also found, defining where the failure mechanism changes. Slopes of heights Hs<hL indicate a toe-failure mechanism with FS values slightly greater than 1, whereas those with Hs>hL indicate a face-failure mechanism with a steady FS1. A linear correlation between hL and c/γ was approximated by
hL=[(15.8)c/γ]+0.3
(7)
which can be used as a preliminary estimate of the most likely failure mechanism for a concave slope (Jeldes et al. 2013).
Fig. 1. Illustration of the reference frame and the components of the concave slope formulation; in practice, the sharp bluff at the top of the slope may be rounded, as indicated by the dashed line

Concave Slopes and Soil Erosion

Experimental and computational studies have shown that slopes with longitudinally concave shapes yield less sediment than slopes with planar or other curvilinear shapes. Here, the focus is on the difference in soil loss between longitudinally concave (referred to henceforth as concave) and planar slopes, which are the two shapes of interest in this study. Because erosion rates increase with slope length and steepness, for a concave slope, the increased erosion due to the increased length is partially counteracted by the decreased erosion due to decreasing slope steepness. In examining this, Young and Mutchler (1969) used field-scale plots on a silt soil and measured 10% less sediment loss on concave slopes with the same average inclination as planar ones. D’Souza and Morgan (1976) used laboratory scale models with a sandy soil, and measured 20–25% less soil loss on concave slopes with the same average inclination as planar slopes. Rieke-Zapp and Nearing (2005) reported 75% less soil loss for concave slopes with similar surface area as planar slopes using experiments conducted on laboratory models with silt soil. Reductions in erosion for concave slopes can also be predicted via computational models. Meyer and Kramer (1969), using the universal soil loss equation (USLE), reported less soil loss for concave slopes of the same height and average steepness as planar slopes, and the difference in soil loss increased with horizontal slope distance. Williams and Nicks (1988), employing the chemical, runoff, and erosion from agricultural management systems (CREAMS) field-scale model, computed approximately 50% lower soil loss for a concave slope of the same average inclination as a planar slope; grass cover and filter strips did not alter the ratio of soil losses between concave and planar slopes. Hancock et al. (2003), using the landform evolution model SIBERIA, computed up to 80% less soil loss for concave slopes of the same horizontal length and average inclination as planar slopes and showed that sediment loss decreases as the slope becomes more concave. Lee et al. (2004) simulated moderately intense storm conditions via the Hillslope Erosion Model and found that concave slopes yielded 17, 22, 24, and 28% less sediment loss than planar slopes when sandy clay, clay, silt, and silty clay soils were used in the simulation. Priyashantha et al. (2009), also using SIBERIA, computed approximately 50% reduction in soil loss between planar and concave slopes with the same overall inclinations and again found that the reduction occurs regardless of vegetation and climate type.
Whereas the aforementioned erosional studies reported significantly less soil loss for concave slopes, most compared the difference in soil loss between planar and concave slopes with similar average steepness, and none of them linked the concave geometry to mechanical stability considerations. Accordingly, this study begins with the examination of concave shapes satisfying a desired degree of stability and then investigates their ability to reduce sediment yield. This is accomplished by comparing the soil loss on concave and planar slopes with the same FS. Finally, this study investigates the impact of construction precision on the concave slope stability.

Methods

Mechanical Stability

In this section, concave slopes for a given design FS for long-term stability are defined and the undrained shear strength required such that short-term stability is unlikely to govern is determined. This approach is conservative in that it neglects any additional shear strength that may be present due to partial saturation and resulting soil suction.

Stable Concave Slopes for ϕ>0 (Long-Term Stability) with Preselected FS

Whereas Eqs. (1)(6) generate concave slopes at limiting equilibrium, a design factor of safety FSD>1 is desired for design and construction. The expression for the concave slope can be extended to any desired FS if a strength reduction factor (Griffiths and Lane 1999; Spencer 1967) equal to FSD is used to reduce the strength parameters as follows:
ϕ=arctan(tanϕFSD)
(8)
c=cFSD
(9)
where ϕ and c = reduced friction angle and cohesion employed in the design to achieve the desired FS. To obtain a concave slope satisfying a preselected FSD, replace ϕ and c in Eqs. (1)(6) with ϕ and c and determine the x,y-coordinates of the concave slope. The resulting concave shape will satisfy a FSFSD under the original ϕ and c strength conditions.
The proposed methodology was compared with results obtained via LEM and FEM analyses for soils with ϕ=20,30,and40° and c/γ ranging from 0.2 to 5 m, where the range c=540kN/m2 covers most of the cohesion values reported by Mesri and Abdel-Ghaffar (1993), and the range γ=1023kN/m3 covers most materials from clays to coarse granular soils [Naval Facilities Engineering Command (NAVFAC) 1986]. Each combination of ϕ and c/γ describing a unique concave slope was analyzed for heights Hs ranging from 2 to 100 m. The FEM analyses were conducted using Phase2 7.0, and the LEM analyses using Slide 6.0. In the FEM analyses, an elastic perfectly plastic model with a Mohr-Coulomb yield criterion and a nonassociated flow rule (zero-dilatancy angle) was employed because it has been shown to provide reliable FS values (Griffiths and Lane 1999). Also, nominal values for Young’s modulus (E=2×104kPa) and Poisson’s ratio (υ=0.3) were used, because they do not influence the calculation of the FS (Cheng et al. 2007; Griffiths and Lane 1999). In the LEM analyses, the Morgenstern-Price, the Spencer, and the simplified Bishop methods were used with over 30,000 critical surfaces analyzed for each set of variables.

Stability Check for Short-Term (ϕ=0) Loading Conditions

Once the long-term (ϕ>0) stable concave shapes are defined, it is necessary to investigate under what conditions short-term failure may occur, e.g., failure during the construction period where the stability is controlled by the undrained shear strength (Su), which is the soil shear strength for ϕ=0 conditions. This is specifically important for soils with a significant amount of fines, where the generated excess positive pore-water pressure may not quickly dissipate. The question then becomes the following: what would be the minimum value of undrained shear strength (Sumin) such that the concave slope defined for long-term stability is also stable for short-term conditions? This problem is approximated by making use of the critical slope concept. A critical contour implies that slopes less steep than this contour are safe, and steeper ones are unsafe, under the same soil properties and slope height. Thus, any critical contour defined for long-term stability with a steepness less than or equal to the critical contour defined for short-term (ϕ=0) loading conditions will be stable in the short term for a given height y. Although the long-term critical contour can be obtained via Eqs. (1)(6), the critical slope contour for the ϕ=0 case can be obtained as a closed-form solution of the characteristic equations (Sokolovskiĭ 1965)
xu(y)=2Suγln[cos(γy2Su)],y0
(10)
where xu = horizontal projection of the slope contour. In the tension zone (hcruy0), xu=0 and hcru=2Su/γ (Fig. 2). This concept is illustrated in Fig. 2, where the long-term critical contour is steeper than the short-term Critical Contours 1, 2, and 3 that have undrained shear strengths Su1, Su2, and Su3, respectively, and therefore it is unsafe in the short term. The opposite yet desirable case is shown for Contour 4, where Su4 is sufficiently large that the global stability is governed by the long-term condition. The value of Sumin to satisfy this condition is found at the intersection of the long- and short-term contours. To find Sumin (1) hcruhcr must be enforced and (2) the transcendental equation xu(y)=xd(y+Δhcr) must be solved numerically, where Δhcr=hcruhcr, Δhcr>0. It should be noted that Sumin is a conservative estimate of the necessary undrained shear strength and can be taken as an upper limiting value of the exact solution.
Fig. 2. Long-term (ϕ>0) critical contour versus short-term (ϕ=0) critical contours defined by different Su values

Soil Erosion and Sediment Yield

Soil loss for concave and planar slopes with the same FS values was investigated using the widely recognized revised universal soil loss equation (RUSLE2) (USDA 2008)
A=RKLSCP
(11)
where the predicted soil loss A (Mg/ha/year) is directly proportional to the rainfall erosivity R (MJmm/hahyear), quantifying the rainfall’s erosive potential; the soil erodibility K (Mghah/haMJmm) defining the soil’s susceptibility to that erosivity; the topographic factor LS (dimensionless) representing slope length and steepness effects; the surface cover factor C (dimensionless); and the conservation practices factor P (dimensionless).
Through RUSLE2 the pairs of planar and concave slopes [Eqs. (1)(6)] that had the same FS under the same values of ϕ, c, γ, and Hs (Fig. 3) were investigated. For soil-loss calculations, the concave slopes were idealized by discretizing the surface into 10–20 linear segments. Because RUSLE2 determines soil loss over the horizontal projection of the slope, for slope segments steeper than 45° (which are not defined in RUSLE2), the soil loss was calculated assuming a planar slope of 45° over the total horizontal length occupied by the steeper part of the contour. This is possible because (1) the runoff is accumulated based on the horizontal projection of the slope rather than over the inclined slope surface length; and (2) in most cases, the amount of accumulated turbulent energy at the upper portion will be minor, because sheet erosion is occurring. Three different soil textures were employed for the analyses: silt loam (high erodibility), clay (moderate erodibility), and sandy loam (moderately low erodibility). A range of possible K values for each soil texture (Haan et al. 1994; USDA 2008) was used (Table 1). All the simulations were conducted for two locations with very different climates: Dakota County, Minnesota (R=2,180MJmm/hahyear) representing dry weather conditions, and Monroe County, Florida (R=10,500MJmm/hahyear) representing high rainfall conditions. For simplicity, a bare surface (C=1) and no conservation practices (P=1) were modeled. A representative value of friction angle was assumed for each soil texture (Budhu 2011; Hough 1957). Values of c/γ ranging from 0.2 to 3 m were employed for the clay soils and from 0.2 to 2 m for the silt and sandy loam soils. Table 1 summarizes the erosion and mechanical properties of the investigated soils.
Fig. 3. Illustration of concave and equivalent planar slopes (same FS and slope height)
Table 1. Soil Composition, Classification, Erodibility, and Assumed Internal Friction Angle of Investigated Soils
 Composition (USDA 2008) (%)   
 ClaySiltSand   
RUSLE soil texture (USDA 2008)<0.002mm<0.050.002mm<20.05mmEquivalent USCS classificationRange of erodibility (K)aAssumed ϕ (degrees)
Silt loam206020ML0.037–0.057b25
Clay452827CL0.032–0.04220
Sandy loam102565SM0.026–0.037b35
a
All K values in SI units of Mghah/haMJmm.
b
Upper bound reported for cases of low permeability and low organic matter content.

Construction and Sensitivity

The widespread use of high-accuracy autoguidance GPS-based construction equipment for grading and earthwork operations suggests that the construction of more complex, nonplanar slope shapes can become more commonplace. The effects of construction precision on the slope stability was investigated by considering the worst-case scenario: the vertical component of the contour is constructed deeper than designed, resulting in a steeper slope. Although the vertical accuracy of three-dimensional grade-control systems is commonly within 30 mm (Trimble 2013), an accuracy of T=200mm was investigated, which may be also achieved by conventional equipment. Because the horizontal accuracy of GPS receivers is usually up to three times greater than the vertical (Berber et al. 2012), the horizontal errors are expected to stay in the millimeter scale, with no significant effects on stability. The effect of construction accuracy on the stability was investigated by (1) lowering by 200 mm the vertical component of various critical concave contours (FS=1) for ϕ=20,30,and40° and c/γ=0.2,1,and5m [Eqs. (1)(6)]; and (2) conducting stability analyses on them via the simplified Bishop’s method.

Results and Discussion

Concave Slopes with Preselected FS Values for Long-Term (ϕ>0) Conditions

Results from LEM analyses are shown in Fig. 4 for the case ϕ=30°, where the computed FS values are plotted against the dimensionless value Hsγ/c. Each point represents a stability analysis on a concave slope of height Hs, cohesion c, and unit weight γ, and each ϕ reflects a different value of FSD. Note that Hsγ/c is the inverse of Taylor’s stability number N (Taylor 1948). A change in failure mechanism is observed to occur at similar values of 1/N regardless of ϕ. The change in failure mode observed on the critical concave slopes (Jeldes et al. 2013) is not altered by the use of a strength-reduction factor, and Eq. (7) can still be used to estimate the most probable mode of failure with c/γ instead of c/γ. Concave slopes with Hs>hL show steady computed values of FSFSD, whereas those with Hs<hL showed FS values higher (more conservative) than the Hs>hL case, with values increasing as Hsγ/c decreases. Similar results were obtained for the ϕ=20and40° cases.
Fig. 4. Computed FS versus Hsγ/c for ϕ=30° and FSD=1.00,1.25,1.50,and2.00; lines are results from simplified Bishop’s method, triangles are results from Morgenstern-Price’s method, and squares are results from Spencer’s method
It is important to note that the methodology proposed here is based on the slip-line field theory, which requires a value of c0 for a valid mathematical solution to exist. Most soils encountered in practice contain sufficient fines to provide at least a small amount of long-term cohesion, such as that used in the example problem. Although some design codes may not permit the use of cohesion for long-term analyses, there are many opportunities not governed by these codes, such as mine reclamation, where a small level of long-term cohesion would allow the design of concave slopes. For those cases where cohesion must be neglected, concave slopes with controlled degree of stability can still be depicted by (1) enforcing the upper portion of the slope to be planar and inclined at ϕ=tan1(tanϕ/FSD) degrees from the horizontal, (2) locating the critical slip surface using any sound LEM or FEM analysis, and (3) monotonically decreasing the inclination of the slope contour below the critical slip, such that the overall FS remains equal to FSD. The rate at which the slope contour decreases can be chosen such that erosion resistance is optimized. Naturally, this approach is limited to slopes with Hs sufficiently large so that face failure is the governing mechanism for the c=0 case, and it may not provide the same erosional benefits of continuous concave slopes that can be depicted for a small c0.
It must also be emphasized that Eqs. (1)(6) constitute an approximate solution in analytical form, so even when Hs>hL there may not be a perfect match between FSD and the actual FS. As seen in Fig. 4, the best match occurs when ϕ20°. Concave slopes with ϕ>20° yield FS values higher than the design FSD (4% higher when ϕ=30°and6% higher when ϕ=40°), which is conservative for design. On the other hand, slopes with ϕ<20° will have FS<FSD, which is not conservative. This is shown in Fig. 4 for the case ϕ=16.1°, where the resulting FS=1.92 is 4% less than the desired FSD=2. Nevertheless, the deviations from the target FSD obtained through the methodology do not compromise its applicability for design, considering that neither the geometry nor the shear strength parameters are usually determined within ±6% accuracy (Duncan 1996). Furthermore, because the FS values calculated by LEMs that satisfy all conditions of static equilibrium can vary as much as 12% (Duncan 1996), an accuracy of ±6% is certainly acceptable. Once the desired shape for the design FS has been obtained, verifications can be conducted employing commercial slope stability software. In the same way, the effects of transient groundwater flow and/or external surcharges on the FS can be investigated, which extends the applications of this solution to other field conditions.

Stability Check for Short-Term (ϕ=0) Conditions

The minimum undrained shear strength Sumin that the soil must possess for concave slopes to have FS>FSD in the short term can be obtained from Fig. 5 as a function of the effective strength parameters ϕ and c. Fig. 5 is a solution chart that assembles the numerical results from Eqs. (1) and (10) and is presented in terms of the dimensionless parameters yγ/c and Sumin/(cFSD) for a range of ϕ. Here, y is the slope height below the tension zone (y=Hs|hcr|, y>0). The inner chart in Fig. 5 provides higher resolution for low values of yγ/c. Note that Fig. 5 defines the conditions under which long-term stability will always govern, so the solutions are conservative. Also, this chart is based on an assumed failure mechanism, where the critical surface exists from or above the toe. Consequently, this solution provides a first check for the short-term stability of concave slopes and does not replace the necessity of conducting stability analyses once the undrained shear strength of the soil has been determined.
Fig. 5. Solution chart for short-term stability check

Soil Loss from the Mechanically Stable Concave Slopes

Soil loss results obtained from RUSLE2 for the silt loam are shown in Fig. 6. All the analyses were conducted on concave and planar slopes with FS=FSD=1 (ϕ=ϕ and c=c), because they provide generic results that can be extended to any FS. Figs. 6(a and b) compare the predicted total soil loss for concave and planar slopes of a silt loam with c/γ=1 under wet weather conditions in Monroe County, Florida, and dry conditions in Dakota County, Minnesota. Planar slopes yielded more sediment than equally stable concave slopes for all K and Hs values investigated, regardless of climate. Soil loss increased with K and Hs in both profiles, but the upper surfaces (soil loss from planar slopes) remain constantly above the lower surfaces (soil loss from concave slopes). Fig. 6(c) compares, for the two weather conditions, the variation in ApAc (Ap = soil loss from planar slopes, Ac = soil loss from concave slopes), or the erosion savings incurred using equally stable concave slopes. This illustrates that concave slopes can reduce erosion to a much larger extent in climates with high rainfall erosivity.
Fig. 6. Erosion analyses for a silt loam soil with ϕ=25°: (a) total soil loss for Monroe County, Florida; (b) total soil loss for Dakota County, Minnesota; (c) difference in soil loss for the Dakota (dry) and Monroe (wet) counties; (d) Ac/Ap versus Hsγ/c; units: A (Mg/ha/year), ApAc (Mg/ha/year), R (MJmm/hahyear), K (Mghah/haMJmm), and Hs (m)
Generalized results for the silt loam soil are shown in Fig. 6(d) in terms of the ratio Ac/Ap and the stability number Hsγ/c. Because the same RUSLE2 R and K were used in both planar and concave slopes, they cancel out when the results are presented in terms of Ac/Ap and the only factor affecting differences in the soil loss is the shape of the slope profile. The use of Hsγ/c enables us to represent the results of different slopes with the same degree of stability for a given value of ϕ. For the range of Hsγ/c values investigated, Ac/Ap ranges from 0.85 to 0.60 [Fig. 6(d)], indicating that concave slopes yield 15–40% less sediment than their planar counterparts. Similar results are shown in Figs. 7 and 8 in terms of ApAc and the Ac/Ap ratio for ϕ=20° (clay soil) and ϕ=35° (sandy loam soil), respectively. The computed Ac/Ap values show soil-loss reduction of the same order and suggest that the concave slopes proposed in this article will reduce sediment delivery by 15–40% regardless of soil erodibility and climate. This finding is extendable to slopes having natural or artificial surface covers (C<1), where the erosion is reduced approximately proportionally for both concave and planar slopes.
Fig. 7. Erosion analyses for the clay soil with ϕ=20°: (a) difference in soil loss for the Dakota (dry) and Monroe (wet) counties; (b) Ac/Ap versus Hsγ/c; units: ApAc (Mg/ha/year), R (MJmm/hahyear), K (Mghah/haMJmm), and Hs (m)
Fig. 8. Erosion analyses for the sandy loam soil with ϕ=35°: (a) difference in soil loss for the Dakota (dry) and Monroe (wet) counties; (b) Ac/Ap versus Hsγ/c; units: ApAc (Mg/ha/year), R (MJmm/hahyear), K (Mghah/haMJmm), and Hs (m)
In Table 2 the relative reduction in soil loss [1(Ac/Ap)] obtained here is compared with values reported in the literature. Although a wide range of erosion reductions were found there, it should be noted that the concave and planar slopes in those studies likely do not satisfy the equal degree of mechanical stability (FS) purposely enforced here. Similar reductions were obtained among the studies except for those reported by Meyer and Kramer (1969), Hancock et al. (2003), and Rieke-Zapp and Nearing (2005), for which the reductions were significantly higher. Meyer and Kramer (1969) and Rieke-Zapp and Nearing (2005) observed that an important amount of sediment deposition occurred at the lowest part of the concave contours, which may explain the higher overall reduction in sediment yield for those studies. In contrast, sediment deposition was not indicated in the RUSLE2 prediction obtained here. Each segment of the concave slopes proposed in this article is the steepest possible to satisfy mass equilibrium requirements, and sedimentation on the profile will likely not occur, especially at large values of ϕ.
Table 2. Reduction in Soil Loss [1(Ac/Ap)] from Concave Contours Reported in the Literature
ReferenceMethodDetails1(Ac/Ap) (%)
Young and Mutchler (1969)ExperimentalField plots10
Meyer and Kramer (1969)ComputationalUSLE0–85
D’Souza and Morgan (1976)ExperimentalLaboratory20–25
Williams and Nicks (1988)ComputationalCREAMS50–55
Hancock et al. (2003)ComputationalSIBERIA50–80
Lee et al. (2004)ComputationalHillslope erosion17–28
Rieke-Zapp and Nearing (2005)ExperimentalLaboratory75
Priyashantha et al. (2009)ComputationalSIBERIA50
This studyComputationalRUSLE215–40

Sensitivity to Construction Accuracy

Fig. 9 shows results obtained from the construction sensitivity analyses for concave slopes with FS=FSD=1 (ϕ=ϕ and c=c) for a range of ϕ and Hsγ/c values, with results expressed as the percent decrease in FS due to the construction error. Slopes with HshL were excluded from the analyses, because it was previously shown that their performing FS values are higher than required (conservative), and they did not present stability variations below equilibrium (FS=1). The results demonstrate that the FS values are not significantly influenced by improper construction within the applied 200 mm of vertical accuracy. The largest difference was found for slopes with c/γ=0.2m, where 5–5.5% lower FS values were obtained. Considering that the analyses were conducted with a vertical accuracy only 1/6 of that which can be achieved with typical autoguidance equipment, it was concluded that the stability of concave slopes is not significantly compromised by the precision of GPS-guided equipment. In addition, practicality suggests that during construction, the top vertical section will likely not be built as depicted here, but will likely be rounded back to a convex shape with a safe average inclination chosen by the designer; this would create a more complex slope with a more natural looking profile. Note that the vertical section corresponds to the maximum surface loading allowed for stability, and any reduction in this loading by shaping this area back results in a reduced surcharge loading and added conservatism. This zone is assumed to have no contribution to the stability or resistance.
Fig. 9. Decrease in FS due to low precision construction (200 mm vertical) of the concave slope

Illustrative Examples

The design of concave slopes for long-term mechanical and erosion stability is illustrated for two hypothetical soils in Monroe County, Florida, indicating high-erosivity conditions. The FEM analyses were conducted to verify the mechanical stability for each case. The short-term stability was checked, followed by erosion analyses for concave and planar slopes with equal FS. The two selected soils were sand (SM or SC) and a fine-grained soil (ML or CL), with the assumed mechanical properties and design parameters for each case as shown in Table 3. A FSD=1.5 was required for each slope.
Table 3. Geometry, Soil Properties, and Probable Failure Mechanisms of Illustrative Examples
      Step 1Step 2
Hypothetical soilFSDHs (m)ϕ (degrees)c (kPa)γ (kN/m3)ϕ (degrees)c (kPa)c/γ (m)hL (m)ConditionPredicted failure
Sand (e.g., SM or SC)1.51535151925.0210.000.538.6hL<HsFace failure FSFSD
Fine-grained (e.g., CL or ML)1.53025351717.2723.331.3722hL<HsFace failure FSFSD

Finding the Concave Profile for Long-Term Stability

Table 3 summarizes the following initial steps: (1) obtain ϕ and c with Eqs. (8) and (9) and (2) estimate the failure mechanism via hL [Eq. (7)] with c instead of c. The x-coordinates of the concave profile were then obtained by introducing ϕ and c in Eqs. (1)(6) for different values of y. For the sandy slope, 8.6m=hL<15m and the concave slope will have a face failure mechanism with a performing FSFSD=1.5. Verification of this condition was provided via FEM analysis (Fig. 10), which yielded FS=1.51. The equivalent planar slope with FS=1.5 is depicted with dashed lines in Fig. 10. Similarly, the fine-grained slope presents a face failure (22m=hL<30m). In this case, because ϕ<20°, a small error was introduced and the FEM analysis indicated a FS=1.45.
Fig. 10. Example case for sandy soil: ϕ=35°, c=15kN/m2, γ=19kN/m3, and Hs=15m; required FS=1.5; shear strains showed for SRF=1.52 to emphasize failure mode
To investigate the sensitivity of the mechanical stability on the value of the selected cohesion, the stability of the slope created in Fig. 10 was analyzed by the FEM for values of cohesion less than that used to obtain the concave profile; values of (2/3)c, (1/2)c, and (1/3)c (10, 7.5, and 5 kPa, respectively) were used in the analysis. The results suggest that the FS would decrease to 1.25, 1.10, and 0.89, respectively. These examples illustrate the role that the small value of the long-term cohesion has on the computed stability, and the importance of determining it very carefully. However, it should be kept in mind that a concave slope with the design FSD could have been obtained for each of these values of c. Each unique combination of ϕ and c/γ will result in a unique concave shape with FSFSD. This implies that as ϕ, c, or γ change, the overall shape of the profile will change, but the expected FS and the expected range of erosion reduction will remain unchanged.
Note that the horizontal extent of the concave and planar slopes in Fig. 10 is similar, but there is less material involved in the concave slope. This has no effect on the balance of cut-and-fill volumes, because the analytical expression for the concave slope could be incorporated into the calculations. However, in mine reclamation, where a shortage of material is common, the construction of concave slopes could be an advantage and provide the benefit of more natural looking slopes and lower sediment loads.

Checking the Short-Term Stability of the Concave Slope

A conservative verification can be made using Fig. 5. For the sandy slope (ϕ=35° and ϕ=25°), y=Hs|hcr|=13.4m, and yγ/c=25.4. From Fig. 5, Sumin/(cFSD)=5.5, and thus the long-term stability of the sandy slope will govern as long as Sumin83kPa. Similarly, for the fine-grained soil y=26.3, yγ/c=19.2, and long-term stability will govern provided that Sumin150kPa.

Erosion Analyses

Total soil loss was computed for low and high erodibility values for each soil texture (Table 4) for the Monroe County, Florida, weather conditions. The results (Table 4) reveal that the sandy concave slope yields on average 24% less sediment than its planar counterpart. Similarly, the fine-grained concave slope yields on average 35% less sediment than a planar slope with the same FSD.
Table 4. Soil Parameters and Results from Erosion Analyses of Illustrative Examples
      1(Ac/Ap)
Soilϕ (degrees)1/N=Hsγ/cR (MJ·mm/ha·h·year)K (Mg·ha·h/ha·MJ·mm)Ac (Mg/ha/year)%Average
Sand25.0228.510,5000.0265572224
0.03774825
Fine-grained17.2721.810,5000.0371,3483435
0.0572,34737

Conclusions

Concave slopes not only resemble natural contours but also have superior erosion resistance. The design of concave slopes requires (1) the definition of concave shapes that provide a desired FS, (2) a quantitative measure of the erosion/sediment delivery reduction, and (3) determination of possible loss of mechanical stability due to improper construction. Using a well-established theoretical framework, a design method with considerations for both mechanical and erosional stability has been proposed in this article. Specifically
An approximate expression was described to provide the coordinates for concave slopes that satisfy a desired FS (long-term conditions) for any combination of ϕ, c, γ, and Hs. The methodology was shown to be conservative for ϕ>20° but provided slightly nonconservative values of design FS for cases with ϕ<20°. The errors introduced by the methodology are less than ±6% for the majority of the cases, which is smaller than the accuracy to which the strength parameters are typically determined.
A simplified graphical solution was developed to estimate the required undrained shear strength Sumin such that the design’s long-term (ϕ>0) stability is assured. The value of Sumin for concave slopes is obtained as a function of the effective strength parameters.
Results from RUSLE2 analyses indicate that the concave slopes proposed here yield 15–40% less sediment than planar slopes of equal FS, regardless of soil erodibility and weather conditions. Although the analyses were conducted on slopes with FSD=1, the findings are valid for slopes with other FS values, as shown in the illustrative example.
Results from the sensitivity analyses reveal that the stability of concave slopes is not significantly influenced by errors in the constructed profile of as great as 200 mm of vertical deviation. Therefore, the stability is not compromised when typical high-accuracy GPS construction equipment is employed for concave slope construction.
Although planar slopes may be simpler and less expensive to construct than concave slopes, the differences will diminish as the use of autoguided GPS-controlled construction equipment increases. The cost differential may be partially offset as well by a reduction in erosion and sediment control efforts, and the intrinsic value obtained from more natural-appearing landforms. The reduction of sediment delivery from constructed slopes is an important consideration in limiting environmental impact during land reclamation and site development. The proposed method suggests a rational design procedure, which makes concave slopes an attractive ecotechnical alternative for more sustainable and natural-appearing earthwork construction.

References

Berber, M., Ustun, A., and Yetkin, M. (2012). “Comparison of accuracy of GPS techniques.” Measurement, 45(7), 1742–1746.
Budhu, M. (2011). Soil mechanics and foundations, Wiley, New York.
Cheng, Y. M., Lansivaara, T., and Wei, W. B. (2007). “Two-dimensional slope stability analysis by limit equilibrium and strength reduction methods.” Comput. Geotech., 34(3), 137–150.
D’Souza, V. P. C., and Morgan, R. P. C. (1976). “A laboratory study of the effect of slope steepness and curvature on soil erosion.” J. Agric. Eng. Res., 21(1), 21–31.
Duncan, J. M. (1996). “State of the art: Limit equilibrium and finite-element analysis of slopes.” J. Geotech. Engrg., 577–596.
Griffiths, D. V., and Lane, P. A. (1999). “Slope stability analysis by finite elements.” Géotechnique, 49(3), 387–403.
Haan, C. T., Barfield, B. J., and Hayes, J. C. (1994). Design hydrology and sedimentology for small catchments, Academic Press, San Diego.
Hancock, G. R., Loch, R. J., and Willgoose, G. R. (2003). “The design of post-mining landscapes using geomorphic principles.” Earth Surf. Process. Landf., 28(10), 1097–1110.
Hough, B. K. (1957). Basic soils engineering, Ronald Press, New York.
Howard, E. J., Loch, R. J., and Vacher, C. A. (2011). “Evolution of landform design concepts.” Trans. Inst. Min. Metall., Sect. A, 120(2), 112–117.
Jeldes, I. A., Vence, N. E., and Drumm, E. C. (2013). “Approximate solution to the Sokolovskiĭ concave slope at limiting equilibrium.” Int. J. Geomech., 04014049.
Lee, H., Koh, H., and Al’Rabia’Ah, H. A. (2004). “Predicting soil loss from logging in Malaysia.” Proc., Int. Conf. of GIS and Remote Sensing in Hydrology, Water Resources and Environment (ICGRSHWE), International Association of Hydrological Sciences (IAHS), Wallingford, U.K., 308–315.
Leopold, L. B., and Langbein, W. B. (1962). “The concept of entropy in landscape evolution.” Geological Survey Professional Paper 500-A, U.S. Government Printing Office, Washington, DC.
Martín-Duque, J. F., Sanz, M. A., Bodoque, J. M., Lucía, A., and Martín-Moreno, C. (2010). “Restoring earth surface processes through landform design: A 13-year monitoring of a geomorphic reclamation model for quarries on slopes.” Earth Surf. Process. Landf., 35(5), 531–548.
Mesri, G., and Abdel-Ghaffar, M. E. M. (1993). “Cohesion intercept in effective stress-stability analysis.” J. Geotech. Engrg., 1229–1249.
Meyer, L. D., and Kramer, L. A. (1969). “Erosion equations predict land slope development.” Agric. Eng., 50(9), 522–523.
Miyamoto, H., Baker, V., and Lorenz, R. (2005). “Entropy and the shaping of the landscape by water.” Non-equilibrium thermodynamics and the production of entropy: Life, earth, and beyond, Springer, Berlin, 135–146.
Naval Facilities Engineering Command (NAVFAC). (1986). Design manual 7.01: Soil mechanics, Alexandria, VA.
Phase2 7.0 [Computer software]. Toronto, Rocscience.
Priyashantha, S., Ayres, B., O’Kane, M., and Fawcett, M. (2009). “Assessment of concave and linear hillslopes for post-mining landscapes.” Proc., 8th Int. Conf. on Acid Rock Drainage (ICARD) and Securing the Future: Mining, Metals & the Environment in a Sustainable Society, Swedish Association of Mines, Minerals and Metal Producers, Stockholm, Sweden, 1164–1175.
Rieke-Zapp, D. H., and Nearing, M. A. (2005). “Slope shape effects on erosion.” Soil Sci. Soc. Am. J., 69(5), 1463–1471.
Rodríguez-Iturbe, I., Rinaldo, A., Rigon, R., Bras, R. L., Marani, A., and Ijjász-Vásquez, E. (1992). “Energy dissipation, runoff production, and the three-dimensional structure of river basins.” Water Resour. Res., 28(4), 1095–1103.
Schor, H. J., and Gray, D. H. (2007). Landforming: An environmental approach to hillside development, mine reclamation and watershed restoration, Wiley, Hoboken, NJ.
Slide 6.0 [Computer software]. Toronto, Rocscience.
Sokolovskiĭ, V. V. (1960). Statics of soil media, Butterworths, London.
Sokolovskiĭ, V. V. (1965). Statics of granular media, Pergamon Press, New York.
Spencer, E. (1967). “A method of analysis of the stability of embankments assuming parallel inter-slice forces.” Géotechnique, 17(1), 11–26.
Taylor, D. W. (1948). Fundamentals of soil mechanics, Wiley, New York.
Toy, T. J., and Chuse, W. R. (2005). “Topographic reconstruction: A geomorphic approach.” Ecol. Eng., 24(1–2), 29–35.
Trimble. (2013). “Trimble introduces new low-power compact card GPS receiver for high-precision applications.” 〈http://www.trimble.com/news/release.aspx?id=012203a〉 (Feb. 11, 2013).
Twidale, C. R. (2007). “Backwearing of slopes: The development of an idea.” Cuatenario Geomorfologia, 21(1–2), 135–146.
USDA. (2008). Science documentation: Revised universal soil loss equation version 2 (RUSLE2), Agricultural Research Service, Washington, DC.
Utili, S., and Nova, R. (2007). “On the optimal profile of a slope.” Soils Found., 47(4), 717–729.
Williams, R. D., and Nicks, A. D. (1988). “Using CREAMS to simulate filter strip effectiveness in erosion control.” J. Soil Water Conserv., 43(1), 108–112.
Yang, C. T., and Song, C. C. S. (1979). “Theory of minimum rate of energy dissipation.” J. Hydr. Div., 105(7), 769–784.
Young, R. A., and Mutchler, C. K. (1969). “Soil movement on irregular slopes.” Water Resour. Res., 5(5), 1084–1089.

Information & Authors

Information

Published In

Go to Journal of Geotechnical and Geoenvironmental Engineering
Journal of Geotechnical and Geoenvironmental Engineering
Volume 141Issue 2February 2015

History

Received: Jun 28, 2013
Accepted: Sep 3, 2014
Published online: Oct 7, 2014
Published in print: Feb 1, 2015

Authors

Affiliations

Isaac A. Jeldes, S.M.ASCE [email protected]
Postdoctoral Research Associate, Dept. of Biosystems Engineering and Soil Science, Univ. of Tennessee, Knoxville, TN 37996; formerly, Graduate Research Assistant, Dept. of Civil and Environmental Engineering, Univ. of Tennessee, Knoxville, TN 37996. E-mail: [email protected]
Eric C. Drumm, M.ASCE [email protected]
Professor, Dept. of Biosystems Engineering and Soil Science, Univ. of Tennessee, Knoxville, TN 37996 (corresponding author). E-mail: [email protected]
Daniel C. Yoder, M.ASCE [email protected]
Professor, Dept. of Biosystems Engineering and Soil Science, Univ. of Tennessee, Knoxville, TN 37996. E-mail: [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