You are on page 1of 12

Geoderma 237–238 (2015) 137–148

Contents lists available at ScienceDirect

Geoderma
journal homepage: www.elsevier.com/locate/geoderma

In situ surface shear strength as affected by soil characteristics and land


use in calcareous soils of central Iran
S. Havaee, M.R. Mosaddeghi ⁎, S. Ayoubi
Department of Soil Science, College of Agriculture, Isfahan University of Technology, Isfahan 84156-83111, Iran

a r t i c l e i n f o a b s t r a c t

Article history: Soil surface shear strength is an important input parameter in the process-based soil erosion models, but its direct
Received 28 December 2013 measurement at the watershed scale is difficult, time-consuming and costly. This study was conducted to predict
Received in revised form 19 August 2014 in situ shear strength of the soil surface using multiple-linear regression (MLR). The land use impact on the sur-
Accepted 22 August 2014
face shear strength was examined as well. A direct shear box was constructed to measure in situ shear strength
Available online xxxx
(cohesion, c and angle of internal friction, φ) of the soil surface in the Semirom region located in Isfahan province,
Keywords:
central Iran. The shear device consisted of a circular shear box (with 10 cm internal diameter and height of 1 cm),
Soil cohesion an S-shaped load cell for measuring horizontal (shear) stress and an electric motor for applying the shear stress.
Angle of internal friction The normal stress acting on the failure surface was adjusted by adding weights on the shear box. Soil surface
In situ direct shear box shear strength was determined using the shear box at 100 locations under three land use systems of grassland,
Multiple-linear regression irrigated farming and dryland farming. Soil particle size distribution (clay, silt, sand and fine clay), organic matter
Pedotransfer functions content (OM), carbonate content, bulk density and gravel content were determined as predictors of the surface
Soil spatial prediction functions shear strength. Normalized difference vegetation index (NDVI) was also calculated using satellite images. The
MLR was employed to predict the shear strength (i.e. c and φ) using two groups of input variables: i) easily-
available soil properties (pedotransfer functions, PTFs) and ii) easily-available soil properties and NDVI (soil spa-
tial prediction functions, SSPFs). A strong negative correlation (r = −0.72, p b 0.001) was found between c and φ
in the studied area. Positive correlation (r = 0.41, p b 0.001) was obtained between c and fine clay content. The c
was negatively correlated (r = −0.31, p b 0.01; r = −0.37, p b 0.001) with sand and gravel contents, respective-
ly. A significant positive correlation (r = 0.47, p b 0.001) was observed between φ and gravel content, indicating
the roughness effect of coarse particles on frictional shear strength. The results also showed that NDVI is an im-
portant factor indirectly explaining the variability of both c and φ in the studied soils. The land use effect on the
soil properties was investigated using the LSD0.05 test. The means of φ somehow follow the variation of gravel
content among the land uses; the highest means of clay and fine clay contents and c, and the lowest means of
sand and gravel contents and φ were observed in dryland farming. The c has the same trend as did clay content;
irrigated farming with the highest mean of sand content and lowest mean of clay content had the lowest mean of
c. The c means were significantly different between irrigated farming and dryland farming. Prediction models of
in situ shear strength derived using both soil properties and NDVI as predictors (SSPFs) were more accurate than
those derived using only soil properties (PTFs).
© 2014 Elsevier B.V. All rights reserved.

1. Introduction order to control accelerated erosion. Factors affecting the rate of erosion
may be considered under three headings: energy, protection and soil
Accelerated soil erosion is a serious threat to environmental sustain- erodibility. The energy group includes the potential ability of rainfall,
ability (Abbaszadeh Afshar et al., 2010). In recent years, in addition to rainsplash, runoff and wind as detaching agents. The protection group
creating financial losses such as filling reservoirs of dams, diminished focuses on factors relating to the plant cover. Soil erodibility defines
soil fertility and productivity, accelerated erosion has become a serious the soil resistance to detachment/transport and depends on soil texture,
threat to human health due to its significant role in polluting surface and structural stability, shear strength, infiltrability and organic matter
subsurface water resources and air (Morgan, 2005). Therefore, it is nec- (OM) and chemical content (Morgan, 2005; Morgan and Nearing,
essary to understand the important factors and take urgent steps in 2011; Torri and Borselli, 2011). Therefore, due to dependence of soil
erodibility on many factors and properties (Morgan and Nearing,
2011; Torri and Borselli, 2011; Zachar, 1982), it is difficult to generalize
⁎ Corresponding author. soil erosion findings on a specific soil type or soils of a region to other
E-mail address: mosaddeghi@cc.iut.ac.ir (M.R. Mosaddeghi). soils/regions.

http://dx.doi.org/10.1016/j.geoderma.2014.08.016
0016-7061/© 2014 Elsevier B.V. All rights reserved.
138 S. Havaee et al. / Geoderma 237–238 (2015) 137–148

Several studies showed that shear strength of the surface soil is an Knapen et al. (2007b) analyzed the WEPP dataset and found a trend of
appropriate dynamic index for evaluating soil erodibility (Foster et al., higher critical flow shear stresses for clayey soils. They combined the con-
1995; Franti et al., 1985; Higuchi et al., 2013; Knapen et al., 2007a,b; centrated flow erodibility and critical flow shear stress data to present a
Léonard and Richard, 2004; Luk and Hamilton, 1986; Torri and texture-based erodibility ranking. The results indicated that silt loam
Poesen, 2014; Torri et al., 2013). Soil shear strength is the maximum and clayey soils have high and low erodibilities, respectively.
shear stress a soil resists before shear failure happens. It is quantified Plant roots could reinforce soils; mechanical properties of the root–
by the two-component Mohr–Coulomb equation (Johnson et al., 1987; soil system are regulated by a combination of soil strength, single root
Koolen and Kuipers, 1983): strength, and the interface strength between soil and roots (Comino
et al, 2010). Waldron (1977) and Wu et al. (1979) considered the
τ ¼ c þ σ n tanφ ð1Þ root contribution to soil shear strength as a cohesion term added to
the Mohr–Coulomb equation (Eq. (1)). Stokes et al. (2009) reported
where τ is the shear strength, c is the cohesive shear strength or cohe- that the presence of plant roots physically reinforce the shear zone.
sion due to intrinsic bonds between soil particles/aggregates, σn is the Moreover, root exudates, as chemical stabilizing agents, could affect
normal stress acting on the failure surface and φ is the angle of internal soil structure and erodibility (Angers and Caron, 1998; Morgan, 2005).
friction. Frictional shear strength (σn tan φ) is due to internal friction be- Several laboratory and field methods are used to measure soil shear
tween soil particles/aggregates and depends on the normal stress acting strength (Bowles, 1978; Bradford et al., 1992; Brunori et al., 1989;
on the failure surface. Koolen and Kuipers, 1983; Léonard and Richard, 2004; Zimbone et al.,
Soil shear strength (τ) is considered as a key property to characterize 1996). Torsional shear apparatus is used to measure c and φ in the field.
soil detachability by raindrop impact (Al-Durrah and Bradford, 1982a; However, shear stress and soil failure are not uniformly distributed over
Bradford et al., 1992; Cruse and Larson, 1977; Nearing and Bradford, the shear surface under the shear ring (Johnson et al., 1987; Koolen and
1985; Torri et al., 1987a,b) and soil resistance to concentrated flow ero- Kuipers, 1983). Although Zimbone et al. (1996) suggested torvane and
sion (Foster et al., 1995; Knapen et al., 2007a,b). Rainsplash detachment hand vane tester as most appropriate and simple devices for quick mea-
processes are linked to shear strength of the surface soil (Brunori et al., surement of soil cohesion, in situ measurement of shear strength using
1989; Watson and Lafflen, 1986). Detachment is due to a combination of small-size shear vane and torvane might not be reliable and a larger mea-
compression and shear under raindrop impact (Al-Durrah and Bradford, suring volume is preferred (Léonard and Richard, 2004). The shear vane
1982a,b). Al-Durrah and Bradford (1982a) observed stronger linear re- test underestimated the shear strength of frictional low-strength soils
gression between rainsplash weight and ratio of rainfall kinetic energy like loams and silt loams (Bradford and Grossman, 1982). Torvane
and undrained soil shear strength than any other soil property. It is re- might underestimate the resistance of sealed soils to raindrop splash be-
ported that detachment decreases in a nonlinear manner with increment cause insertion of the vane blades into soil breaks the crust and measures
of shear strength (Bradford et al., 1992; Cruse and Larson, 1977; Torri lower soil strength. Moreover, torvane measurement in frictional sandy
et al., 1987b). Léonard and Richard (2004) found that saturated shear soils is highly susceptible to uncontrolled vertical force exerted on the
strength could be considered as the best soil property to predict critical vane while shearing (Bradford et al., 1992). Fall-cone penetrometer was
shear stress (i.e. minimum shear stress of flowing water required to used as a simple and rapid test for in situ measurement of near-surface
move soil particles) and runoff erosion. Critical shear stress, a measure shear strength in soil erosion studies (e.g. Al-Durrah and Bradford,
of soil resistance to water erosion, is an important parameter governing 1982a; Bradford and Grossman, 1982; Bradford et al., 1992; Nearing
soil detachment by runoff in several erosion models. Rauws and Govers and Bradford, 1985). However, two major limitations of this test are: 1)
(1988) noticed that the soil conditions at which rill flow becomes erosive dependency of empirical calibration constant, needed for translating
are governed by shear strength of the surface soil. depth of cone penetration to shear strength, on cone angle and soil
There are evidences for the importance of φ in soil detachment/ type/texture (Bradford et al., 1992; Towner, 1973), and 2) uncontrolled
erosion processes as well. Al-Durrah and Bradford (1982b) concluded variable depth of cone penetration which ultimately affects the mea-
that rainsplash weight and angle are influenced by c and φ, and by soil sured soil layer (Bradford and Grossman, 1982; Bradford et al., 1992).
deformability. Nearing and Bradford (1985) found that shear strength Collis-George et al. (1993) introduced a quick/easy resin-impregnated
(measured using a fall-cone penetrometer) overestimated overall soil re- plate method to measure the surface shear strength. However, the shear
sistance against splashing. A correction term including φ was suggested surface is not easily detectable and is wavy-like at the edges of the
to develop a unique linear relation between soil detachment and ratio shear plate. Shear device proposed by Zhang et al. (2001) for measuring
of waterdrop kinetic energy and fall-cone shear strength. Thus, φ should the surface shear strength is interesting and promising but the measured
also be considered in soil strength measurements/predictions for erosion c and φ are not intrinsic (internal) soil strength parameters but depend on
researches. the sandpaper–soil interface properties.
Most frequently soil physical properties that could affect shear Direct measurement of soil shear strength parameters (c and φ) at
strength of the surface soil are particle size distribution (Knapen et al., large scale is difficult, time-consuming and costly. Researchers have
2007b; Shainberg et al., 1994), water content/matric potential (Bradford used indirect methods such as pedotransfer functions (PTFs) to predict
and Grossman, 1982; Cruse and Larson, 1977; Knapen et al., 2007b), ag- difficult-to-measure soil properties (e.g. hydraulic properties) using
gregation (Baumgartl and Horn, 1991), stone size (Léonard and Richard, easily-available properties as predictors (Bouma, 1989; Wösten et al.,
2004), network of plant roots and vegetation cover (Franti et al., 1999; 2001). In a first attempt, Horn and Fleige (2003) provided class PTFs
Knapen et al., 2007b; Torri et al., 2013) and tillage systems and time for c and φ in terms of soil texture, structure and matric suction in
(Knapen et al., 2007a). Gilley et al. (1993) reported that soil shear Germany. Goktepea et al. (2008) estimated shear strength of plastic
strength could be related to texture, OM and bulk density. Soil texture clays using PTFs that were derived by statistical and neural network ap-
(mean particle size or clay content) plays a key role in soil structure proaches. An artificial neural network (ANN) framework was employed
and/or erodibility, and is usually used as primary indicator of soil resis- by Habibagahi and Bamdad (2003) to predict mechanical behavior of
tance to erosion (Knapen et al., 2007b; Line and Meyer, 1989). Based on unsaturated soils. Hirata et al. (1990) employed multiple regression
WEPP dataset, Knapen et al. (2007b) observed linear relation between analysis between the mechanical and physical properties of cohesive
soil detachment capacity and flow shear stress which differed among soils. Khalilmoghadam et al. (2009) and Besalatpour et al. (2012) devel-
the soil textural classes. Poesen (1992) obtained a relation between soil oped the ANN-based and ANFIS-based PTFs to predict c, measured by a
erodibility factor and geometric mean diameter of primary particles and shear vane, using easily-available data in Zagros region of Iran.
found that medium (silt loam) soils are most susceptible to erosion High rate of land use change from pasture to dryland farming is re-
with coarse- and fine-textured soils having lower erodibility factors. ported in Iran (Abbaszadeh Afshar et al., 2010). Previous reports on
S. Havaee et al. / Geoderma 237–238 (2015) 137–148 139

land use impacts focused mostly on soil physical and chemical proper- 31°02′56″ to 31°32′58″ E latitudes (Fig. 1). The average rainfall and tem-
ties (Ayoubi et al., 2012) and soil hydraulic properties (Kelishadi et al., perature in the region are 350 mm and 10.6 °C, respectively. The site se-
2014), but there is little knowledge about the effects of land use lected is under three land uses including dryland farming, grassland and
(change) on the surface shear strength especially in calcareous soils. irrigated farming. The drylands in the studied site are mainly used for
There are few studies that have reported on soil cohesion measurement cereals such as wheat and barley production and irrigated farming is
and prediction at a large scale (i.e. watershed scale). Khalilmoghadam used to produce a variety of crops and orchards. Soil erosion rate is
et al. (2009) measured soil c using a shear vane and compared its esti- reported about 10 Mg per hectare per year in the region (Group of
mation using ANNs and MLR models in Zagros region. They reported Experts, 1998) which is greater than its average in the world.
significant lower c in degraded rangelands compared to rangelands The distribution of soil sampling points in the study area (45 in
due to destruction of soil aggregates and reduced OM by tillage and cul- grassland, 31 in irrigated farming and 24 in dryland farming, in total
tivation. Knapen et al. (2007a) also reported lower c values in conven- 100 locations) is shown in Fig. 1. Generally, the soils of the studied
tional tillage compared to conservational tillage because area are classified as Typic Calcixerepts according to Soil Taxonomy
conventionally-tilled soil is slightly consolidated after tillage and (Soil Survey Staff, 2006). The studied area is located on different parent
would be more erodible when compared with consolidated and materials including limestone, marly limestone, and quaternary deposi-
residue-protected soil under conservational tillage. It is crucial to inves- tions which are all enriched by carbonates. Soil sampling (0–5 cm) was
tigate the effect of land use (change) on shear strength of the surface soil done in a random manner at 100 locations; the sampling locations were
towards better understanding the impact of land use on soil erodibility well scattered among the land uses and are representative of the study
and erosion. To the best of authors' knowledge, no documentation is area (Fig. 1).
found reporting in situ measurement and prediction of both c and φ in
large scale. Hence, the objectives of this study were to: i) investigate
the impact of land use on in situ shear strength of the surface soil (c 2.2. Soil Shear Strength Measurement
and φ), ii) derive prediction equations for c and φ using MLR, and iii) de-
termine factors that are most effective in controlling the c and φ in cal- In situ shear strength of the saturated surface soil was measured at the
careous soils of Semirom region, central Iran. 100 locations in 2011, August and September. The surface soil was care-
fully saturated before the tests. However, the surface soil might not be
fully saturated (but satiated) because of the effect of entrapped/dissolved
2. Materials and Methods air although the matric potential was essentially zero. The saturated con-
dition was chosen for soil strength measurement because saturated soil
2.1. Description of the Study Area strength is frequently used as a measure of soil erosion resistance in
several erosion models such as WEPP (Foster et al., 1995), EUROSEM,
The study site was located in the Semirom region, Isfahan, central KINEROS and LISEM (cited in Knapen et al., 2007b). Moreover, Léonard
Iran between 51°27′07″ to 51°37′54″ N longitudes and between and Richard (2004) found that saturated shear strength could be

Fig. 1. Location of the study area in central Iran and the distribution of soil sampling points.
140 S. Havaee et al. / Geoderma 237–238 (2015) 137–148

considered as the best soil property to predict critical shear stress and Normal stress acting on the failure surface was adjusted by adding
runoff erosion. weights 2, 5 and 10 kg (σn of 25, 63 and 125 hPa, respectively), on the
In this study, a new direct shear box was designed and constructed shear box (Fig. 2b). Three shear tests with the mentioned σn values
to measure in situ shear strength of the surface soil. The device can were performed in each location as close as possible to each other.
determine the shear failure envelope (i.e. both c and φ) in the field Therefore, in total 300 shear tests were done. Low values of σn were
(Fig. 2a). The device consists of a frame, stand, shear box, weights, chosen to represent small overburden load on the surface soil in accord
S-shaped load cell and electric motor (Fig. 2a, b). The heavy frame is with Zhang et al. (2001). The shear force (stress) is applied using an
inserted by stand into the surrounding soil to resist against vibration electric motor working by a portable battery at a fixed displacement
and the shear force applied to the soil surface. The shear box is a circular rate of 4 mm min−1 (Fig. 2b). This loading rate was quick enough to en-
metal box with 10 cm internal diameter (i.e. area of 78.5 cm2) and sure the undrained shear test that did not allow pore water drainage.
height of 1 cm. The shear box is inserted into the surface soil and the Undrained shear test is usually used in soil erosion studies because
surrounding soil is carefully excavated (Fig. 2c, d) to ensure that the fail- both soil splash and detachment are fast and do not allow the pore
ure only happens under the shear box at a depth of 1 cm. The upper sur- water drainage (Al-Durrah and Bradford, 1982a; Knapen et al., 2007a,
faces of the frame and the shear box are checked for the horizontal b; Léonard and Richard, 2004). The shear force was measured using
position by a leveler. an S-shaped load cell (DACELL:UU-K50 model, ± 0.05 N), and moni-
We did not treat and/or level the soil surface in order to preserve the tored using a data logger during a test (Fig. 2a, b). The shear force vs.
surface intact as much as possible. A point with smooth soil surface was time data could be saved and transferred to PC after a test. The un-
chosen because the surface roughness would result in uneven distribu- drained c (intercept) and φ were calculated by regressing the three τ
tion of the normal stress on the failure surface. The free gravels on the vs. σn pairs (see Eq. (1)).
surface (not embedded in the soil) were collected because they would
not contribute to the soil strength and would cause uneven distribution 2.3. Soil Sampling and Analyses
of the normal stress on the failure surface. The soils were not leveled be-
cause most of the selected positions were not steepy (slope 1–2%). For Disturbed and undisturbed surface soil samples were collected from
the rest, we chose points with naturally level surface on the slopes. the locations close to where the shear tests were conducted. The sam-
The measurements were done on the unvegetated spots but close to ples were transported to the laboratory for soil analyses. Gravel content
the plants. In some places, roots, root residues and debris were present (N2 mm) was determined using the volumetric method. Organic matter
in the measured soil volume. We did not clip the vegetation because it content and calcium carbonate content (CaCO3) were measured using the
was not specifically the aim of this study to explore the effect of root wet-oxidation (Nelson, 1982) and back-titration methods (Nelson and
on shear strength; it was not possible to measure the shear strength of Sommers, 1986), respectively. Soil texture and particle size distribution
soil including the large and strong roots of apple and walnut orchards were determined by sieving, sedimentation and centrifuge methods
in the irrigated farming using the constructed shear box. However, the after pre-treatments. Pre-treatments included removal of organic matter
litter input and residue would certainly affect the shear strength of the as a cementing agent by hydrogen peroxide and dispersion of the suspen-
soil close to the plants. sion using sodium hexametaphosphate (Calgon®). Sand (0.05–2 mm)

Motor position

S-shaped load
cell
Weight

Stand

(a) (b)

σn
Shear box τ

Failure surface Soil

(c) (d)
Fig. 2. In situ direct shear box used in this study: (a) schematic of the device, (b) a picture showing as to how shear strength of the surface soil is measured, (c) 3D schematic of the soil
shearing during a test, and (d) 2D schematic of the soil shearing during a test.
S. Havaee et al. / Geoderma 237–238 (2015) 137–148 141

content was determined by sieving and sedimentation method was used easily-available soil properties and NDVI (in SSPFs). Easily-available soil
for measuring silt (0.002–0.05 mm) and clay (b 0.002 mm) contents (Gee properties for simple and MLR analyses included particle size distribution
and Bauder, 1986). Fine clay (b0.0002 mm) was determined using the (sand, silt, and clay percent or dg and σg), organic matter content (OM),
centrifuge method. Moreover, geometric mean diameter (dg) and geo- calcium carbonate content (CaCO3) and gravel content. Input variables
metric standard deviation (σg) were calculated as follows (Shirazi and (x) and transformed variables (ln(x), x2, x−1, x0.5 and exp(x)) were used
Boersma, 1984): as predictors in the MLR analyses. Stepwise method of the SAS package
(SAS Institute Inc., 1999) was used to develop the MLR models.
Xn
The accuracy of the derived PTFs and SSPFs was evaluated by the co-
dg ¼ expðaÞ and a ¼ 0:01 f i lnMi ð2Þ
i¼1
efficient of determination (R2) and root mean square errors (RMSE) cal-
culated as follows:

!0:5 N 
X 2
Xn
2 2 Y i −Y^ i
σ g ¼ expðbÞ and b ¼ 0:01 f i ln M i −a ð3Þ
2 i¼1
i¼1 R ¼ 1− 2
ð7Þ
X
N  
Y i −Y i
where the constant of 0.01 converts mass percent to mass ratio, n is the i¼1
number of particle fractions (sand, silt, clay, …), and fi is the mass percent-
age of particles with arithmetic mean diameter Mi. The dg and σg were
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
used because they describe log-normal particle size distribution. u
uX N  2
Moreover, dg and σg might be better predictors of soil shear strength u Y i −Y^ i
u
(c and φ) than fractions' percents since some researchers (e.g. Poesen, t i¼1
RMSE ¼ ð8Þ
1992; Torri et al., 1997) successfully related the soil erodibility factor N
to the dg.
Soil bulk density (ρb, Mg m−3) was determined using undisturbed where Yi and Ŷi are, respectively, observed and predicted values of c or
soil samples which were taken by stainless steel cylinders from the sur- φ. TheY i is the mean of the observed values of c or φ and N is the number
face layer. The soil cores were dried at 105 °C for 48 h, and then weighed. of measurements (i.e. 100).
The relative bulk density (ρb-rel) was calculated using the following equa- The Akaike Information Criterion (AIC) was also used to compare the
tion (Håkansson and Lipiec, 2000) to characterize the degree of compact- efficacy of the PTFs and SSPFs as follows (Akaike, 1974):
ness of soils irrespective of soil texture:  
SSE
AIC ¼ 2K þ N ln ð9Þ
ρb N
ρb‐rel ¼ ð4Þ
ρb‐ref
where K is the number of regression coefficients in the PTFs or SSPFs, N
where ρb-ref is the reference bulk density (Mg m−3) and was calculated by is the number of measurements and SSE is the sum of squares of errors.
the equation proposed by Jones (1983) and used by Dexter (2004) for the Because N is relatively small in our study, a version of Eq. (9) was used
lower limit of critical ρb for root growth: to calculate the corrected AIC (AICc):

ρb‐ref ¼ 1:882–0:0083 clay ð5Þ 2K ðK þ 1Þ


AICc ¼ AIC þ : ð10Þ
N−K þ 1
where clay is expressed in percent.
The absolute AICc values are not interpretable since they depend on
2.4. Normalized Difference Vegetation Index (NDVI) the units of model variables and sample size. However, the difference
(Δ) between AICc values of two models is independent of the mentioned
Normalized difference vegetation index (NDVI) was also calculated factors. It can be used to measure the degree of support for one model rel-
using satellite images. The NDVI for the study region was derived ative to the other (best) model (i.e. Δ = AICc(model) − AICc(best model)) as
using 6 main band ETM+ landsat at a spatial resolution of 30 m by follows: Δ ≤ 2, substantial support; 2 ≤ Δ ≤ 4, less support; 4 ≤ Δ ≤ 7,
30 m (Indian Space Applications Centre, 2007). The NDVI was calculated considerably less support; and Δ N 10, no support. The model with a
as the ratio of the red (RED) and near infrared (NIR) bands of a sensor lower value of AICc is the better one (Burnham and Anderson, 2004).
system:
3. Results and Discussion
NIR−RED
NDVI ¼ : ð6Þ
NIR þ RED 3.1. Examples of Shear Tests Using the In Situ Direct Shear Box

Results of in situ shear tests showed that two main types of soil be-
2.5. Statistical Analyses and Soil Shear Strength Prediction havior under shear might be recognized: brittle failure and ductile fail-
ure. Examples of shear stress–displacement curves are shown in Fig. 3
Descriptive statistics including the mean, standard deviation (SD), for a sandy clay loam soil under grasslands and a clay soil under dryland
minimum, maximum, coefficient of variation (CV), kurtosis and skew- farming. It was observed that at low normal stress (e.g. 25 hPa) and in
ness, and Pearson linear correlations among various parameters and compacted state, the soils usually sheared in a brittle manner indicating
properties were calculated using the SAS package (SAS Institute Inc., quick and sudden failure with minimal shear displacement before fail-
1999). The normality of data was evaluated by the Kolmogorov– ure. The peak strength followed by residual strength is easily distinguish-
Smirnov test. The land use effect on soil properties including the surface able in this type of soil behavior (see Fig. 3a). The shear behavior shifted
shear strength (c and φ) was investigated using the LSD0.05 test. to the ductile (compressive) failure with increasing the normal stress
Simple- and multiple-linear regression (MLR) methods were used to (i.e. 63 and 125 hPa) and for the loose soils. Ductile soil behavior is indi-
predict soil shear strength (i.e. c and φ) using two groups of input vari- cated by gradual failure with large shear displacement before failure
ables entered the pedotransfer functions (PTFs) and soil spatial predic- (see Fig. 3a, b). Since it was difficult to determine the end-point of the
tion functions (SSPFs): i) easily-available soil properties (in PTFs) and ii) test and consequently the shear strength for the ductile behavior, the
142 S. Havaee et al. / Geoderma 237–238 (2015) 137–148

70 et al. (2009) also reported that soil cohesion (c) increased with incre-
ment of clay content. Aggregation due to the presence of clay particles
60 (Baumgartl and Horn, 1991; Lebert and Horn, 1991) and microstructure
(Hatibu and Hettiaratchi, 1993) might play a role, too.
50
Shear stress (hPa)

3.2. Descriptive Statistics of the Measured Soil Properties and NDVI


40
Descriptive statistics of the measured soil properties and NDVI
30
are given in Table 1. The clay content varied in a wide range (i.e.
20.2–62.4 kg 100 kg−1) with a mean of 36.5 kg 100 kg−1 in the stud-
20
25 hPa ied area. The dominant soil textural classes were clay, clay loam, sandy
clay loam and loam. The c and φ had broad range, indicating very low to
10 63 hPa
high saturated shear strength in the studied soils, which were useful for
125 hPa
deriving reliable PTFs and SSPFs. The saturated c and φ values and their
0
0 5 10 15 means, minimums and maximums (Fig. 5 and Table 1) are comparable
Shear displacement (mm) with those reported in the literature (e.g. Cruse and Larson, 1977;
Léonard and Richard, 2004; Nearing and Bradford, 1985). Cruse and
(a) Larson (1977) reported effective c and φ values of a silt loam soil to be
in the ranges 6–11 hPa and 21–35° depending on ρb. Nearing and
50 Bradford (1985) measured the saturated c and φ of four soils using a
tri-axial test in the ranges (0–17.5 hPa and 14.8–18.6°) which were
comparable with our results. Baumgartl and Horn (1991) and Lebert
40
and Horn (1991) measured greater c and φ values for unsaturated
Shear stress (hPa)

well-structured soils using the laboratory shear box in the σn range of


30 0–4000 hPa. Using the σn range 0–500 hPa where the aggregates
were not crushed, the bulk c and φ for a well-structured clay soil at
matric suction 30 hPa were about 70 hPa and 38°, respectively
20
25 hPa

10 63 hPa 70
125 hPa
60
τ = 40.2 + 0.0411σn
Shear strength (hPa)

0 2
50 R =1
0.0 2.5 5.0 7.5 10.0 12.5 15.0
Shear displacement (mm) φ
40
(b) 30

Fig. 3. Shear stress–displacement curves using three normal stresses (i.e. 25, 63 and 20 c
125 hPa) for: (a) a sandy clay loam soil under grassland, and (b) a clay soil under dryland
farming. Arrows indicate the values of soil shear strength. 10

0
0 25 50 75 100 125
shear stress at relative shear displacement of 15% (i.e. 15 mm) was con-
sidered as soil shear strength (see arrows in Fig. 3a, b) in accord with Normal stress (hPa)
Bowles (1978) and Zhang et al. (2001). Frictional soil behavior is real-
ized by increasing the shear strength (or peak strength) with increasing
(a)
the normal stress (Fig. 3a, b).
70
It was observed that sandy soils mainly sheared in brittle manner,
while clayey soils behaved like a soft/ductile material as revealed by 60
comparing Fig. 3a with Fig. 3b. As mentioned before, brittle failure was
Shear strength (hPa)

τ = 17.9 + 0.3525 σn
typically observed in compacted or coarse-textured soils at low normal 50 2
R = 0.99
stresses. Hatibu and Hettiaratchi (1993) also stated that a range of me-
chanical behavior might be expected in agricultural soils depending on 40
φ
soil physical properties such as texture, water content and microstruc-
30
ture. They reported that with increasing clay content, the soil behavior
approaches towards ductile, but with increasing the sand content the 20
brittle behavior would be expected. However, a range of intermediate
behaviors might be observed depending on the soil water content and 10 c
microstructure.
The Mohr–Coulomb shear failure envelopes are shown in Fig. 4 for 0
0 25 50 75 100 125
two soils as examples. It was generally found that shear failure enve-
lopes for fine-textured soils had greater c and smaller slope (or φ) Normal stress (hPa)
(Fig. 4a) compared to those for coarse-textured soils, which had rela- (b)
tively smaller c and larger slope (or φ) (Fig. 4b). These results are in
agreement with those reported in the literature, e.g. Koolen and Fig. 4. Examples of Mohr–Coulomb shear failure envelopes for: (a) a clay soil, and (b) a
Kuipers (1983) and Hatibu and Hettiaratchi (1993). Khalilmoghadam sandy clay loam soil.
S. Havaee et al. / Geoderma 237–238 (2015) 137–148 143

Table 1
Statistical description of the measured soil properties and NDVI which were used as inputs and outputs in the regression analyses.

Variables Mean SD Minimum Maximum CV Skewness Kurtosis

c (hPa) 25.30 6.50 8.40 41.00 0.25 −0.11 −0.50


φ (degree) 12.50 5.80 0.70 30.10 0.46 0.21 −0.95
Clay (kg 100 kg−1) 36.50 11.0 20.20 62.40 0.30 0.60 −0.50
Sand (kg 100 kg−1) 30.30 9.70 14.60 48.10 0.32 0.21 −0.95
Silt (kg 100 kg−1) 33.20 6.90 17.40 49.30 0.21 0.24 −0.30
OM (kg 100 kg−1) 1.90 1.30 0.20 4.80 0.68 0.63 −0.64
CaCO3 (kg 100 kg−1) 26.00 13.00 0.98 62.50 0.50 0.30 −0.26
Fine clay (kg 100 kg−1) 18.30 6.30 4.70 29.70 0.34 −0.01 −0.92
OM/clay (by mass) 0.050 0.040 0.002 0.180 0.80 0.85 0.07
Gravel (%v/v) 23.40 9.10 9.00 48.00 0.38 0.73 −0.09
ρb (Mg m−3) 1.28 0.13 0.98 1.90 0.10 0.91 0.36
ρb-rel 0.78 0.10 0.60 1.16 0.12 0.14 −0.02
NDVI 0.05 0.01 −0.13 0.31 0.20 −0.38 0.71

(Baumgartl and Horn, 1991). Horn and Fleige (2003) also provided class compactness of the studied soils varied from very loose to compacted
PTFs for c and φ in various soil textures and structures at matric suctions state (Table 1). Overall, the studied soils had a wide range in shear
of 60 and 330 hPa. Their tabulated c and φ values could not be compared strength parameters, soil properties, ρb and ρb-rel which were useful
with our data because they measured unsaturated shear strength. for deriving the reliable PTFs and SSPFs between the shear strength
Léonard and Richard (2004) analyzed the saturated c values from 10 and soil properties. Skewness and kurtosis values indicate that the dis-
different studies and found that in most cases, the c ranged from 0 to tribution of soil properties and parameters and NDVI is normal in the
150 hPa with few values exceeded 200 hPa. The results of Zhang et al. region.
(2001) indicated low adhesion values (i.e. 0–3.2 hPa) and high angle
of friction values (i.e. 42–52°) but these coefficients are not intrinsic
soil strength parameters and basically could not be compared with 3.3. Correlations Between Shear Strength Parameters and Soil Properties
our data.
In order to define variability into three classes, we followed the The correlation coefficients between the shear strength parameters
system suggested by Wilding (1985). In general, the soil properties (c and φ) and the soil properties are presented in Table 2. A strong neg-
having CV N 0.35 (highly variable) were in the order: OM/clay ative correlation was obtained between c and φ (r = −0.72, p b 0.001)
(CV = 0.80) N OM (CV = 0.68) N CaCO 3 (CV = 0.50) N φ (CV = in the studied region. The linear regression between c and φ is also
0.46) N gravel (CV = 0.38) in the region. The variables having a CV shown in Fig. 5. This relationship is reasonable indicating that any factor
between 0.35 and 0.15 (classified as moderately variable) were fine that increased the frictional shear strength of the natural soils would be
clay, clay, sand, c, silt, and NDVI. The remaining variables had low expected to reduce the cohesive shear strength. Hemmat et al. (2010)
variability (CV b 0.15). The OM had a high CV (i.e. 68.4%); it varied studied the effect of long-term incorporation of organic manures on
in the range 0.2–4.8 kg 100 kg−1, respectively, for dryland farming shear strength parameters (c and φ) of a calcareous soil in central Iran,
and irrigated farming (particularly under apple and walnut or- and similarly reported a significant negative correlation (r = − 0.73,
chards). High values of OM in irrigated farming are attributed to or- p b 0.001) between c and φ.
ganic manure application and high amount of crop residues and tree A significant positive correlation (r = 0.41, p b 0.001) was obtained
litter returned to the soil. Minimum, maximum and SD of soil CaCO3 between c and fine clay content (Fig. 6a and Table 2) indicating that
were 0.98, 62.5 and 13 kg 100 kg−1 , respectively. Gravel content fine-textured soils have relatively higher cohesive shear strength.
ranged from 9 to 48%v/v in the region (Table 1). This finding is reasonable because the fine particles strengthen
The ρb ranged from 0.98 to 1.90 Mg m−3, presumably due to the inter-particles' bonds and promote aggregation due to their colloi-
wide range of texture and OM, and different management practices in dal size, high surface area and a large number of chemical bonds
the land uses. The ρb-rel range (0.60–1.16) indicates that the degree of (Baumgartl and Horn, 1991; Bohn et al, 2001; Horn and Fleige, 2003;
Lebert and Horn, 1991). Significant positive correlation (r = 0.47,
50 p b 0.001) was observed between φ and gravel content (Fig. 6b and

c = 36.57 - 0.949 ϕ
40 2 Table 2
R = 0.512, p <0.001
Correlation coefficients between soil shear strength parameters (c and φ) and soil proper-
ties in the studied area†.
30 Variable Unit c (hPa) φ (°)
c (hPa)

c hPa 1
φ ° −0.72*** 1.00
20 Clay kg 100 kg−1 0.28** −0.28**
Sand kg 100 kg−1 −0.31** 0.24*
Silt kg 100 kg−1 −0.01 0.12
10 OM kg 100 kg−1 −0.10 0.04
CaCO3 kg 100 kg−1 0.00 0.04
Gravel kg 100 kg−1 −0.37*** 0.47***
Fine clay kg 100 kg−1 0.41*** −0.40***
0 OM/clay by mass −0.18 0.10
0 5 10 15 20 25 30 ρb Mg m−3 0.02 0.02
ρb-rel – 0.16 −0.14
NDVI – −0.23* 0.34***

Fig. 5. Relationship between soil cohesion (c, hPa) and angle of internal friction (φ, °) in *, ** and *** stand for significant correlations at 95%, 99% and 99.9% probability levels,
the studied region. respectively.
144 S. Havaee et al. / Geoderma 237–238 (2015) 137–148

50 correlation (r = 0.40, p b 0.001) between OM and NDVI supports the


c = 15.79 + 0.483Fine clay indirect effect of vegetation on soil mechanical properties. This result in-
2
R = 0.170, p <0.001 dicates an indirect protective effect of vegetation on frictional shear
40
strength of the surface soil presumably due to the strengthening effect
of plant root networks, root residues and debris in the soil. Positive ef-
30 fect of vegetation on soil mechanical resistance against water erosion
c (hPa)

has been reported previously (Normaniza and Barakban, 2006; Pollen,


2007; Roering et al., 2003; Schiechtl, 1991; Schmidt et al., 2001; Torri
20
et al., 2013; Wu and Watson, 1998). Comino et al. (2010) reported
that with increasing vegetation cover, soil shear strength increased
10 because of the interface strength between soil and roots. Torri et al.
(2013) observed that an increase in root density would considerably
increase the soil shear strength. The derived PTFs and SSPFs in
0 Table 4 reveal the nonlinear relations between c or φ and inputs
0 5 10 15 20 25 30
such as NDVI.
Fine clay content (kg 100kg-1) The correlations between c or φ and ρb, ρb-rel and chemical proper-
(a) ties (OM and carbonate contents) were not significant (Table 2). Per-
haps, soil texture and its influence on structure were the main factors
35 affecting c and φ in the region, which mask the impact of degree of com-
ϕ = 5.556 + 0.304 Gravel pactness, OM and carbonate contents. The OM effect on soil properties
30 R ² = 0.224, p <0.001 might be distinguishable when other factors such as soil texture are
not changed; for instance, the OM content varies due to different man-
25 agement systems in a farm with similar soil texture (Dexter, 2004).
Moreover, nonlinear/complex relations between the c or φ and OM
20
(see PTFs and SSPFs in Table 4) could not be inferred from simple corre-
15 lation analysis. Complex effects of OM on soil mechanical properties
have been reported in the literature (Arthur et al., 2012a,b, 2013;
10 Chakraborty et al., 2014; Défossez et al., 2014; Mosaddeghi et al.,
2000; Soane, 1990). The impact of OM on soil mechanical strength is
5 variable depending on soil type, water content and porosity, organic
matter type (plant residue, manure) and decomposition state (fresh
0
0 10 20 30 40 50 crop residue, humified organic matter) (Arthur et al., 2013; Soane,
1990). Zhang and Hartge (1990) also stated that mechanisms involved
Gravel content (kg 100kg-1)
in interaction between mineral components and OM with respect to
(b) soil mechanical properties are complex. Some authors reported a posi-
tive relationship between soil cohesion and OM content (Davies, 1985;
Fig. 6. Soil cohesion (c, hPa) as a function of fine clay content (a), and the angle of internal Hartge, 1975; Zhang and Hartge, 1990). On the contrary, some re-
friction (φ, °) as a function of gravel content (b) in the studied region. searchers found, in cases such as incorporated peat, highly humified
OM in compacted soils, that there was a negative relationship between
c and OM content (Ekwue, 1990; Ohu et al., 1985; Zhang and Hartge,
Table 2), indicating the roughness effect of coarse particles on frictional 1990). Hemmat et al. (2010) studied the effect of long-term incorpora-
shear strength. tion of three organic manures on shear strength parameters (c and φ)
Similar to fine clay, a significant (but weaker) positive correlation of a calcareous soil in central Iran. They found that the c value decreased
(r = 0.28, p b 0.01) was obtained between c and clay content. Morgan when high rate of organic manures (i.e. 100 Mg ha−1) was added to the
(2005) stated that soil shear strength increases with an increment in soil. They concluded that at high soil OM, a large portion of the mineral
clay or OM contents. Khalilmoghadam et al. (2009) and Besalatpour surfaces are coated by organic compounds that may reduce the number
et al. (2012) also reported that cohesion (c) of the soil surface (deter- of mineral to mineral contacts. Therefore, the effect of OM on soil me-
mined using a shear vane) was affected by clay content. The c negatively chanical behavior is very different and depends on various factors such
correlated (r = −0.31, p b 0.01; r = −0.37, p b 0.001) with sand and as amount and type of organic materials (Mosaddeghi et al., 2000;
gravel contents (Table 2). These findings indicated that coarse particles Soane, 1990). Arthur et al. (2012a) using soils with clay contents ranging
would reduce soil cohesion by separating finer soil particles and reducing from 4.5 to 50 kg 100 kg−1 found no significant effect of OM on soil com-
soil aggregation. Baumgartl and Horn (1991) and Lebert and Horn (1991) paction behavior. Arthur et al. (2013) observed that soil response to
observed that aggregation could significantly increase the shear strength compressive stresses was not directly controlled by OM, rather the indi-
of aggregates and bulk soil. Moreover, a negative correlation was ob- rect effects due to OM-induced changes in soil water content and bulk
served between φ and fine clay content (r = −0.39, p b 0.001) due to density were reported. Chakraborty et al. (2014) reported higher com-
the fact that fine particles could reduce the internal (inter-particles) fric- pressibility and porosity for the soil which had greatest OM due to
tion of the soil. Sand and clay were significantly correlated with φ (r = long-term application of farmyard manure. However, Arthur et al.
0.24, p b 0.05 and r = −0.28, p b 0.01, respectively, Table 2). Therefore, (2012b) found that soil resistance and resilience to mechanical stresses
coarser particles (i.e. gravel) decreased the soil cohesion and finer parti- were significantly affected by OM among differently-managed soils;
cles (i.e. fine clays) decreased frictional strength greatly in the soils of high OM soils were less susceptible to compressive stresses and had
the region. Wang et al. (2010) reported that the type of clay minerals greater recovery than low OM soils. Although a wide range of OM
could affect internal friction angle. Horn et al. (2005) indicated that the (0.2–4.8 kg 100 kg−1) was recorded in the region (Table 1), different
shear strength of arable soils depends on the shape and size distribution types of OM inputs to the soils such as animal manures, crop residues
of sand particles. and tree litter could affect the soil mechanical behavior differently. In
Positive correlation was found between φ and NDVI but the c was other words, not only the quantity of OM could determine soil behavior
negatively correlated with NDVI in the study area (Table 2). Significant but also the OM quality might be decisive in terms of soil mechanical
S. Havaee et al. / Geoderma 237–238 (2015) 137–148 145

drylands and irrigated farmlands on different geomorphic surfaces.


Low clay content in irrigated farming could partially contribute to
its lower c (Fig. 7). The highest mean of fine clay content in dryland
farming might also be associated with minimum clay leaching
under low precipitation.
Maximum and minimum means of OM content were obtained for ir-
rigated and dryland farming, respectively (Fig. 7). In dryland farming,
cereals were conventionally cultivated with little addition of organic
materials to the soils because most of the residues are consumed by
livestock. Therefore, dryland farming had the highest mean of clay con-
tent and the lowest mean of OM content, which resulted in the highest
mean of c (Fig. 7). Clay loam and sandy clay loam are the dominant tex-
tural classes in irrigated farming. This land use is particularly under
apple and walnut orchards and receives high amounts of OM input
Fig. 7. Mean comparisons of soil cohesion (c), angle of internal friction (φ), clay, sand, fine with least disturbance. Therefore, the highest mean of OM content and
clay, gravel, OM and relative bulk density (ρb-rel) among the land uses. Values with differ- the lowest means of clay content and c were observed in irrigated farm-
ent letters in each group are significantly different (LSD, p b 0.05). ing (Fig. 7). It was expected that c should have increased with increasing
OM content. However, it is not only the OM content which varied
among the land uses, but also the soil texture was different. Moreover,
behavior and the relationships between shear strength parameters and the quality of OM might be different among different land uses, which
OM content (Soane, 1990). was not evaluated in this study. The main part of the OM inputs into
the soils under irrigated farming was the low-quality apple/walnut
3.4. Effects of Land Use on Soil Physical and Chemical Properties and Shear tree residues and litter. Such OM would not effectively participate in
Strength Parameters the cohesion between particles and even in the soil frictional strength
(φ). Ohu et al. (1985) examined three texturally-different soils mixed
The means of φ, c and soil properties were compared among the land with 3, 10 and 17%w/w peat moss, and found that the vane shear
uses using LSD test (Fig. 7). Fig. 7 shows the statistically significant dif- strength of the compacted soils increased with increasing compressive
ference (p b 0.05) between the means of φ in dryland farming and other force and decreased with increasing rate of peat moss addition. Zhang
land uses. The variation of φ means somehow follows the variation of and Hartge (1995) found that c of a silt loam soil was not influenced
gravel content among the land uses; dryland farming with the lowest by OM additions at high initial matric potential (−60 and −300 hPa),
gravel content had lowest mean of φ. There was no significant difference but was decreased at low initial matric potential (−700 hPa). General-
between the means of φ or gravel content in grassland and irrigated ly, they concluded that soil c and φ reduced with addition of highly-
farming although it is greater in grassland. Field observations indicated humified OM.
that density of grass roots was high in the surface soil of grassland. How- The differences between averages of ρb-rel in grassland, irrigated farm-
ever, root density in the surface soil was lower in the irrigated farming ing and dryland farming are statistically significant (Fig. 7). The average
presumably due to soil manipulation and tillage practices. Accordingly ρb-rel and ρb were in the order: dryland farming N grassland N irrigated
Tengbeh (1993), Stokes et al. (2009) and Torri et al. (2013) reported farming. However, Kelishadi et al. (2014) obtained lower OM and higher
that an increase in root density would considerably increase the soil ρb-rel in swelling soils under pasture as compared to cultivated soils due to
shear strength. Fan and Su (2008) also reported greater soil shear uncontrolled overgrazing and its-induced soil compaction in Koohrang
strength in the presence of plant roots. Moreover, root exudates region of central Zagros. They observed that soil hydraulic properties
could stabilize soil structure and increase the soil strength (Angers and are majorly influenced by soil structure and management practices rather
Caron, 1998; Morgan, 2005). Waldron (1977) and Wu et al. (1979) con- than by intrinsic soil properties like texture.
sidered the contribution of plant roots to the soil shear strength by adding
a term to the Mohr–Coulomb equation. 3.5. PTFs and SSPFs for Predicting Soil Shear Strength Parameters
The lowest mean of φ was observed in dryland farming wit/Similar
to fine clay, a significant the highest mean of fine clay content, while Two groups of available inputs for multiple-linear regression (MLR)
the highest mean of φ was obtained in grassland with the lowest analyses were used (Table 3) to predict soil shear strength parameters
mean of fine clay content (Fig. 7). These findings confirm that fine par- (see Table 4). The inputs for model group 1 (i.e. PTFs) were based on
ticles reduce the internal (inter-particles) friction of the soil. Results soil properties (particle size distribution, OM, CaCO3, ρb and ρb-rel).
in Fig. 7 show that c had statistically significant difference only between While the inputs for model group 2 (i.e. SSPFs) were those for model
irrigated farming and dryland farming. Nevertheless, the c has the same group 1 (i.e. soil properties) and NDVI. These models were derived for
trend as clay content: dryland farming had the highest mean of c and two PSD scenarios: (i) clay, silt and sand percents, and (ii) dg and σg
grassland and irrigated farming had the intermediate and lowest (Table 4).
means of c, respectively (Fig. 7). Khalilmoghadam et al. (2009) reported Developed SSPFs with higher R2 and smaller RMSE and AICc
significant lower c in degraded rangelands compared to rangelands in values (Table 4), were identified more powerful to predict c and φ
Zagros region due to destruction of soil structure and reduced OM con-
tent by tillage and cultivation. However, they suggested that land degra-
dation might increase the c in steep areas because fine particles would
be removed due to erosion and the proportion of coarse particles in-
Table 3
creased. In low lands, crust formation could increase c in the degraded
Inputs used in multiple-linear regression analyses for developing pedotransfer functions
rangelands. (PTFs) and soil spatial prediction functions (SSPFs).
The highest and lowest means of clay content were observed in
Group Model Inputs
dryland farming and irrigated farming, respectively, perhaps due to
clay leaching as well as translocation of subsoil coarse materials by 1 PTFs Soil properties including particle size distribution, organic matter,
moldboard plow inversion into the surface soil in the irrigated farm- calcium carbonate, bulk density and relative bulk density
2 SSPFs Soil properties + NDVI
ing. This difference might be partly attributed to the existence of
146 S. Havaee et al. / Geoderma 237–238 (2015) 137–148

Table 4
Derived PTFs and SSPFs using multiple-linear regression methods for predicting c and φ in the study area†.

Model Parameter Developed models R2 RMSE AICc

Scenario (i)
PTF c c = 23.23 − 0.26 gravel + 1.94 fine clay0.5 − 1.85 × 10−20 exp(sand) 0.37*** 6.04 −523.03
PTF φ φ = 287.17 + 15.71 OM/clay0.5 + 1.19 × 10−20 exp(sand) − 1.44 × 10−20 exp(gravel) − 66.77 ln(clay) − 0.54*** 4.20 −594.58
429.4 OM/clay + 325.5 silt−1 + 0.5 gravel − 2232 sand−1
SSPF c c = 26.69 − 30.43 NDVI − 1.75 exp(sand) + 107.94 NDVI2 + 0.24 fine clay − 0.24 gravel 0.43*** 5.93 −528.62
SSPF φ φ = 427.6 − 56.8 ln(clay) + 180.16 NDVI + 1.009 exp(sand) − 0.08 exp(OM) − 5.6 CaCO−13 + 1.91.5 silt−1 − 0.59*** 3.86 −602.65
2107.8 clay−1 − 1.83 fine clay0.5 − 151.9 exp(NDVI) + 0.21 gravel

Scenario (ii)
PTF c c = 14.41 − 0.28 gravel + 6.05 ln(fine clay) + 0.001 CaCO3 − 1401.03 dg 0.35*** 6.28 −517.91
PTF φ φ = −2.34 + 239.63 OM/clay2 + 1.97 fine clay0.5 + 0.58 gravel − 0.02 σg − 1.83 OM0.5 0.45*** 4.38 −586.81
SSPF c c = 15.52 − 1404 d2g + 113.3 NDVI2 − 0.26 gravel + 0.001 CaCO23 + 5.52 ln(fine clay) − 32.29 NDVI 0.41*** 6.04 −522.74
SSPF φ φ = 235.1 + 197.01 NDVI − 52.2 exp(OM/clay) − 1.9 fine clay0.5 − 164.7 exp(NDVI) + 0.23 gravel − 1.64 OM−1 0.51*** 4.11 −593.73

*, ** and *** stand for significance at 95%, 99% and 99.9% probability levels, respectively.

in the region. Results showed that SSPFs could predict the shear 4. Conclusions
strength parameters using both scenarios better than PTFs. The dif-
ferences between AICc of two models (Δ) can be used to measure 1) In situ shear strength of the soil surface (c and φ) was measured, using
the degree of support for one model relative to the other (best) model a shear box, in calcareous soils of central Iran. A strong negative corre-
(i.e. Δ = AICc(model) − AICc(best model)) as follows: Δ ≤ 2, substantial sup- lation was found between c and φ in the studied area. Shear strength
port; 2 ≤ Δ ≤ 4, less support; 4 ≤ Δ ≤ 7, considerably less support; and parameters (c and φ) strongly depended on soil particle size distribu-
Δ N 10, no support (Burnham and Anderson, 2004). The PTFs had con- tion and gravel content, as revealed by correlation analysis.
siderably less support (4 ≤ Δ ≤ 7) when compared to the SSPFs (i.e. 2) Positive correlation was obtained between c and fine clay content.
best models). Generally scenario (i) led to prediction equations with However, c was negatively correlated with sand and gravel contents.
greater R2 and smaller RMSE and AICc values (Table 4). For example, it Significant positive correlation between φ and gravel content indi-
was found that the SSPFs using scenario (i) are more accurate than cates the roughness effect of coarse particles on frictional shear
using scenario (ii) (Δ = 6 and 11 for c and φ, respectively). Overall, strength. This finding reveals the importance of surface soil gravel
the two scenarios showed approximately similar evaluation criteria in erosion processes.
(R2, RMSE and AICc), but better performance of models was obtained 3) Land use affected the shear strength of the surface soil as well. The
using scenario (i). Generally, better prediction models were devel- highest mean clay and fine clay content and c, and the lowest mean
oped for φ than for c in the region (Table 4). sand and gravel contents and φ were observed in dryland farming.
Gravel, fine clay, sand and NDVI were entered into the model using The irrigated farming with the highest mean sand content and lowest
scenario (i) for predicting c (Table 4). Fine clay and gravel had the addi- mean clay content had the lowest mean c. However, grassland and ir-
tive and reductive effects, respectively, on c as discussed in previous sec- rigated farming did not have significant differences for soil c and φ.
tions. In addition, sand reduced soil cohesion (c). In addition to gravel, 4) It was found that the c and φ are mainly controlled by soil particle
fine clay and dg, CaCO3 entered the PTF and SSPF using scenario (ii) for size distribution and gravel content in the study region. Moreover,
c. The σg negatively affects the φ as revealed in the PTF using scenario the NDVI is an important factor indirectly explaining the variability
(ii) (Table 4). Julien (1995) showed that the soil shear strength was in- of both c and φ in the region. It is suggested to examine the effect
fluenced by the size and shape of soil particles and cohesion decreased of aggregation and structural stability on the c and φ in the region.
with increasing particle size. Fine clay particles due to their colloidal 5) Prediction models of shear strength derived using soil properties and
nature and high surface area, could enhance chemical bonding and in- NDVI as predictors (i.e. SSPFs) were more accurate than those de-
crease c. Similarly, Oliveira (1997) found that the shear strength param- rived using soil properties only (i.e. PTFs). Developed prediction
eters are strongly influenced by the colloidal clay particles and as a models might be used in soil erodibility researches in the region
consequence by the soil microstructure. and/or could be included in the process-based soil erosion models.
It was found that NDVI is an important parameter for predicting shear Feasibility of these applications should be examined in the future.
strength of the surface soil in the studied region. Results (Table 4)
showed that with increasing NDVI, c initially decreased, and then it in- Acknowledgments
creased at higher NDVI values (N0.15) in a nonlinear trend. Gravel,
clay, sand, fine clay, silt, carbonate, and OM contents and NDVI were in- We would like to thank the Isfahan University of Technology for the
cluded in the SSPF model for φ (Table 4). In this model, the effects of financial support of the study. Special appreciation is extended to
gravel and sand were positive. The NDVI affected the φ in a nonlinear Mohsen Morshedizad for drawing Fig. 2a, c.
manner: initially increased, and then decreased at high NDVI values
(N0.15). The results show that OM content had a negative effect on φ. Appendix A. Supplementary Data
Ohu et al. (1985) reported the vane shear strength of the compacted
soils decreased with increasing amount of incorporated OM. Zhang and Supplementary data associated with this article can be found in the
Hartge (1995) found that upon adding high amount of OM to the soil, c online version, at http://dx.doi.org/10.1016/j.geoderma.2014.08.016.
and φ were reduced due to the impact of effective stress at three initial These data include Google map of the most important areas described
matric potential (− 60, − 300 and − 700 hPa). Défossez et al. (2014) in this article.
found that an increase in OM could weaken soil because of higher soil
water content and lower bulk density. Ekwue (1990) showed that References
adding peat to the soil reduced the c and φ. Moreover, clay and fine
Abbaszadeh Afshar, F., Ayoubi, S., Jalalin, A., 2010. Soil redistribution rate and its relation-
clay contents had negative effects on φ in the developed SSPFs ship with soil organic carbon and total nitrogen using 137Cs technique in a cultivated
(Table 4) as previously discussed. complex hillslope in western Iran. J. Environ. Radioact. 101, 606–614.
S. Havaee et al. / Geoderma 237–238 (2015) 137–148 147

Akaike, H., 1974. A new look at the statistical model identification. IEEE Trans. Autom. Hartge, K.H., 1975. Organic matter contribution to stability of soil structure. Soil Condi-
Control 19, 716–723. tioners. SSSA Special Publication No. 7, Madison, WI.
Al-Durrah, M.M., Bradford, J.M., 1982a. Parameters for describing soil detachment due to Hatibu, N., Hettiaratchi, D.R.P., 1993. The transition from ductile flow to brittle failure in
single water drop impact. Soil Sci. Soc. Am. J. 46, 836–840. unsaturated soils. J. Agric. Eng. Res. 54, 319–328.
Al-Durrah, M.M., Bradford, J.M., 1982b. The mechanism of raindrop splash on soil sur- Hemmat, A., Aghilinategh, N., Sadeghi, M., 2010. Shear strength of repacked remoulded
faces. Soil Sci. Soc. Am. J. 46, 1086–1090. samples of a calcareous soil as affected by long-term incorporation of three organic
Angers, D.A., Caron, J., 1998. Plant-induced changes in soil structure: processes and feed- manures in central Iran. Biosyst. Eng. 107, 251–261.
backs. Biogeochemistry 42, 55–72. Higuchi, K., Chigira, M., Lee, D.H., 2013. High rates of erosion and rapid weathering in a
Arthur, E., Moldrup, P., Holmstrup, M., Schjønning, P., Winding, A., Mayer, P., de Jonge, L. Plio–Pleistocene mudstone badland, Taiwan. Catena 106, 68–82.
W., 2012a. Soil microbial and physical properties and their relations along a steep Hirata, S., Yao, S., Nishida, K., 1990. Multiple regression analysis between the mechanical
copper gradient. Agric. Ecosyst. Environ. 159, 9–18. and physical properties of cohesive soils. Soils Found. 30, 91–108.
Arthur, E., Schjønning, P., Moldrup, P., de Jonge, L.W., 2012b. Soil resistance and resilience Horn, R., Fleige, H., 2003. A method for assessing the impact of load on mechanical stabil-
to mechanical stresses for three differently managed sandy loam soils. Geoderma ity and on physical properties of soils. Soil Tillage Res. 73, 89–99.
173–174, 50–60. Horn, R., Fleige, H., Richter, F.H., 2005. Prediction of mechanical strength of arable soils
Arthur, E., Schjønning, P., Moldrup, P., Tuller, M., de Jonge, L.W., 2013. Density and perme- and its effects on physical properties at various map scales. Soil Tillage Res. 82, 47–56.
ability of a loess soil: long-term organic matter effect and the response to compres- Indian Space Applications Centre (ISRO) Ahmedabad-380 015, 2007. Data Products Soft-
sive stress. Geoderma 193–194, 236–245. ware Division. Signal and Image Processing Group, SAC/ISRO, Ahmedabad-380 015,
Ayoubi, S., Ahmadi, M., Abdi, M.R., Abbaszadeh, Afshar F., 2012. Relationships of 137Cs in- (Government of India).
ventory with magnetic measures of calcareous soils of hilly region in Iran. J. Environ. Johnson, C.E., Grisso, R.D., Nichols, T.A., Bailey, A.C., 1987. Shear measurement for agricul-
Radioact. 112, 45–51. tural soils: a review. Trans. ASAE 30, 935–938.
Baumgartl, T., Horn, R., 1991. Effect of aggregate stability on soil compaction. Soil Tillage Jones, C.A., 1983. Effect of soil texture on critical bulk densities for root growth. Soil Sci.
Res. 19, 203–213. Soc. Am. J. 47, 1208–1211.
Besalatpour, A., Hajabbasi, M.A., Ayoubi, S., Afyuni, M., Jalalian, A., Schulin, R., 2012. Soil Julien, P., 1995. Erosion and Sedimentation. Cambridge University Press, UK.
shear strength prediction using intelligent systems: artificial neural networks and Kelishadi, H., Mosaddeghi, M.R., Hajabbasi, M.A., Ayoubi, S., 2014. Near-saturated soil hy-
an adaptive neuro-fuzzy inference system. Soil Sci. Plant Nutr. 58, 149–160. draulic properties as influenced by land use management systems in Koohrang re-
Bohn, H.L., McNeal, B.L., O'Connor, G.A., 2001. Soil Chemistry. John Wiley & Sons, Inc., New gion of central Zagros, Iran. Geoderma 213, 426–434.
York. Khalilmoghadam, B., Afyuni, M., Abbaspour, K.C., Jalalian, A., Dehghani, A.A., Schulin, R.,
Bouma, J., 1989. Using soil survey data for quantitative land evaluation. Adv. Soil Sci. 9, 2009. Estimation of surface shear strength in Zagros region of Iran — a comparison
177–213. of artificial neural networks and multiple-linear regression models. Geoderma 153,
Bowles, J.E., 1978. Engineering Properties of Soils and Their Measurement, Second edition. 29–36.
McGraw-Hill Book Company, New York, (213 pp.). Knapen, A., Poesen, J., De Baets, S., 2007a. Seasonal variations in soil erosion resistance
Bradford, J.M., Grossman, R.B., 1982. In-situ measurement of near-surface soil strength by during concentrated flow for a loess-derived soil under two contrasting tillage prac-
the fall-cone device. Soil Sci. Soc. Am. J. 46, 685–688. tices. Soil Tillage Res. 94, 425–440.
Bradford, J.M., Truman, C.C., Huang, C., 1992. Comparisons of three measures of resistance Knapen, A., Poesen, J., Govers, G., Gyssels, G., Nachtergaele, J., 2007b. Resistance of soils to
of soil surface seals to raindrop splash. Soil Technol. 5, 47–56. concentrated flow erosion: a review. Earth-Sci. Rev. 80, 75–109.
Brunori, F., Penzo, M.C., Torri, D., 1989. Soil shear strength: its measurement and soil Koolen, A.J., Kuipers, H., 1983. Agricultural soil mechanics. Advanced Series in Agricultural
detachability. Catena 16, 59–71. Sciences. vol. 13. Springer-Verlag, Berlin, (241 pp.).
Burnham, K.P., Anderson, D.R., 2004. Multimodel inference: understanding AIC and BIC in Lebert, M., Horn, R., 1991. A method to predict the mechanical strength of agricultural
model selection. Sociol. Methods Res. 33, 261–304. soils. Soil Tillage Res. 19, 275–286.
Chakraborty, D., Watts, C.W., Powlson, D.S., Macdonald, A.J., Ashton, R.W., White, R.P., Léonard, J., Richard, G., 2004. Estimation of runoff critical shear stress for soil erosion from
Whalley, W.R., 2014. Triaxial testing to determine the effect of soil type and organic soil shear strength. Catena 57, 233–249.
carbon content on soil consolidation and shear deformation characteristics. Soil Sci. Line, D.E., Meyer, L.D., 1989. Evaluating interill and rill erodibilities for soils of different
Soc. Am. J. 78, 1192–1200. textures. Trans. ASAE 32 (6), 1995–1999.
Collis-George, N., Philippa, E., Tolmie, E., Moahansyah, H., 1993. Preliminary report on a Luk, S.H., Hamilton, H., 1986. Experimental effects of antecedent moisture and soil
new method for determining the shear strength of a soil surface: the resin plate strength on rainwash erosion of two Luvisols, Ontario. Geoderma 37, 29–43.
method. Aust. J. Soil Res. 31, 539–548. Morgan, R.P.C., 2005. Soil Erosion and Conservation, 3rd ed. Blackwell Publishing Ltd., UK,
Comino, E., Marengo, P., Rolli, V., 2010. Root reinforcement effect of different grass (304 pages).
species: a comparison between experimental and models results. Soil Tillage Res. Morgan, R.P.C., Nearing, M.A. (Eds.), 2011. Handbook of Erosion Modelling. Wiley-
110, 60–68. Blackwell.
Cruse, R.M., Larson, W.E., 1977. Effect of soil shear strength on soil detachment due to Mosaddeghi, M.R., Hajabbasi, M.A., Hemmat, A., Afyuni, M., 2000. Soil compactibility as af-
raindrop impact. Soil Sci. Soc. Am. J. 41, 777–781. fected by soil moisture content and farmyard manure in central Iran. Soil Tillage Res.
Davies, P., 1985. Influence of organic matter content, moisture status and time after 55, 87–97.
reworking on soil shear strength. Soil Sci. 36, 299–306. Nearing, M.A., Bradford, J.M., 1985. Single waterdrop splash detachment and mechanical
Défossez, P., Richard, G., Keller, T., Adamiade, V., Govind, A., Mary, B., 2014. Modelling the properties of soils. Soil Sci. Soc. Am. J. 49, 547–552.
impact of declining soil organic carbon on soil compaction: application to a cultivated Nelson, R.E., 1982. Carbonate and gypsum. In: Page, A.L. (Ed.), Methods of Soil Analysis:
Eutric Cambisol with massive straw exportation for energy production in Northern Part I: Agronomy Handbook No 9. American Society of Agronomy and Soil Science So-
France. Soil Tillage Res. 141, 44–54. ciety of America, Madison, WI, pp. 181–195.
Dexter, A.R., 2004. Soil physical quality. Part I. Theory, effects of soil texture, density, and Nelson, D.W., Sommers, L.P., 1986. Total carbon, organic carbon and organic matter. In:
organic matter, and effects on root growth. Geoderma 120, 201–214. Page, A.L. (Ed.), Methods of Soil Analysis: Part 2: Agronomy Handbook No 9.
Ekwue, E.J., 1990. Organic matter effects on soil shear strength properties. Soil Tillage Res. American Society of Agronomy and Soil Science Society of America, Madison, WI,
16, 289–297. pp. 539–579.
Fan, C.C., Su, C.F., 2008. Role of roots in shear strength of root-reinforced soils and with Normaniza, O., Barakban, S.S., 2006. Parameters to predict slope stability — soil water and
high moisture content. Ecol. Eng. 33, 157–166. root profiles. Ecol. Eng. 28, 90–95.
Foster, G.R., Flanagan, D.C., Nearing, M.A., Lane, L.J., Risse, L.M., Finkner, S.C., 1995. Water Ohu, J.O., Raghavan, G.S.V., Mckyes, E., Mehuys, G., 1985. The shear strength of compacted
Erosion Prediction Project (WEPP). Technical documentation. NSERL Report No. 10. soils with varying organic matter contents. Trans. ASAE 29, 0351–0355.
National Soil Erosion Research Laboratory. USDA-ARS, West Lafayette, IN 47907- Oliveira, R., 1997. Understanding adhesion: a means for preventing fouling. Exp. Therm.
1196. Fluid Sci. 14, 316–322.
Franti, T.G., Laflen, J.M., Watson, D.A., 1985. Soil Erodibility and Critical Shear Under Con- Poesen, J., 1992. Mechanisms of overland flow generation and sediment production on
centrated Flow. ASAE Summer Meet. The Ohio State University, Columbus, OH, USA, loamy and sandy soils with and without rock fragments. In: Parsons, A.J.,
(256 pp.). Abrahams, A.D. (Eds.), Overland Flow Hydraulics and Erosion Mechanics. UCL Press,
Franti, T.G., Laflen, J.M., Watson, D.A., 1999. Predicting soil detachment from high dis- London, UK, pp. 275–305.
charge concentrated flow. Trans. ASAE 42, 329–335. Pollen, N., 2007. Temporal and spatial variability in root reinforcement of streambanks:
Gee, G.W., Bauder, J.W., 1986. Particle size analysis. In: Klute, A. (Ed.), Methods of Soil accounting for soil shear strength and moisture. Catena 69, 197–205.
Analysis: Part 1 Agronomy Handbook No 9. American Society of Agronomy and Soil Rauws, G., Govers, G., 1988. Hydraulic and soil mechanical aspects of till generation on ag-
Science Society of America, Madison, WI, pp. 383–411. ricultural soils. J. Soil Sci. 39, 111–124.
Gilley, J.E., Elliot, W.J., Laflen, J.M., Simanton, J.R., 1993. Critical shear stress and critical Roering, J.J., Schmidt, K.M., Stock, J.D., Dietrich, W.E., Montgomery, D.R., 2003. Shallow
flow rates for initiation of rilling. J. Hydrol. 142, 251–271. landsliding, root reinforcement, and the spatial distribution of trees in the Oregon
Goktepea, A.B., Altunb, S., Altintasc, G., Tan, O., 2008. Shear strength estimation of plastic Coast Range. Can. Geotech. J. 40, 237–253.
clays with statistical and neural approaches. Build. Environ. 43, 849–860. SAS Institute Inc, 1999. SAS/STAT User's Guide. Ver. 8.0. SAS Institute Inc., Cary, NC.
Group of Experts, 1998. Range Land Management Project of Semirom District. The Bureau Schiechtl, H.M., 1991. Ground Bioengineering Systems, Ground Bioengineering Tech-
of Range land Conservation. Land Resources Bureau of Isfahan Province, (64 pp. (In niques for Slope Protection and Erosion Control. Blackwell Science Ltd., UK.
Farsi)). Schmidt, K.M., Roering, J.J., Stock, J.D., Dietrich, W.E., Montgomery, D.R., Schaub, T., 2001.
Habibagahi, G., Bamdad, A., 2003. A neural network framework for mechanical behavior The variability of root cohesion as an influence on shallow susceptibility in the
of unsaturated soils. Can. Geotech. J. 40, 684–693. Oregon Coast Range. Can. Geotech. J. 28, 995–1024.
Håkansson, I., Lipiec, J., 2000. A review of the usefulness of relative bulk density values in Shainberg, I., Laflen, J.M., Bradford, J.M., Norton, L.D., 1994. Hydraulic flow and water qual-
studies of soil structure and compaction. Soil Tillage Res. 53, 71–85. ity characteristics in rill erosion. Soil Sci. Soc. Am. J. 58, 1007–1012.
148 S. Havaee et al. / Geoderma 237–238 (2015) 137–148

Shirazi, M.A., Boersma, L., 1984. A unifying quantitative analysis of soil texture. Soil Sci. Wang, G., Suemine, A., William, H.S., 2010. Shear-rate-dependent strength control on the
Soc. Am. J. 48, 142–147. dynamics of rainfall-triggered landslides, Tokushima Prefecture, Japan. Earth Surf.
Soane, B.D., 1990. The role of organic matter in soil compactibility: a review of some prac- Process. Landf. 35, 407–416.
tical aspects. Soil Tillage Res. 16, 179–201. Watson, D.A., Lafflen, J.M., 1986. Soil strength, slope and rainfall effects on intertill erosion.
Soil Survey Staff, 2006. Keys to Soil Taxonomy. U.S. Department of Agriculture, Natural Trans. ASAE 29, 98–102.
Resources Conservation Service. Wilding, L.P., 1985. Spatial variability: its documentation, accommodation and implica-
Stokes, A., Atger, C., Bengough, A.G., Fourcaud, T., Sidle, R.C., 2009. Desirable plant root traits tion to soil surveys. In: Nielsen, D.R., Bouma, J. (Eds.), Soil Spatial Variability. Pudoc,
for protecting natural and engineered slopes against landslides. Plant Soil 324, 1–30. Wageningen, The Netherlands, pp. 166–194.
Tengbeh, G.T., 1993. The effect of grass roots on shear strength variations with moisture Wösten, J.H.M., Pachepsky, Y.A., Rawls, W.J., 2001. Pedotransfer functions: bridging the
content. Soil Technol. 6, 287–295. gap between available basic soil data and missing soil hydraulic characteristics. J.
Torri, D., Borselli, L., 2011. Water erosion, In: Huang, P.M., Li, Y., Sumner, M.E. (Eds.), 2nd Hydrol. 251, 123–150.
ed. Handbook of Soil Science. vol. 2. CRC Press, pp. 22.1–22.19. Wu, T.H., Watson, A., 1998. In situ shear tests of soil blocks with roots. Can. Geotech. J. 35,
Torri, D., Poesen, J., 2014. A review of topographic threshold conditions for gully head de- 579–590.
velopment in different environments. Earth Sci. Rev. 130, 73–85. Wu, T.H., Mckinnell, W.P., Swanston, D.N., 1979. Strength of tree roots and landslides on
Torri, D., Sfalanga, M., Chisci, G., 1987a. Threshold conditions for incipient rilling. Catena 8, Prince Of Wales Island, Alaska. Can. Geotech. J. 16, 19–33.
97–105. Zachar, D., 1982. Soil Erosion. Elsevier/North-Holland, Inc., 52 Vanderbilt Avenue New
Torri, D., Sfalanga, M., Del Sette, M., 1987b. Splash detachment: runoff depth and soil co- York, New York.
hesion. Catena 14, 149–155. Zhang, H.Q., Hartge, K.H., 1990. Cohesion in unsaturated sandy soil and the influence of
Torri, D., Poesen, J., Borselli, L., 1997. Predictability and uncertainty of the soil erodibility organic matter. Soil Technol. 3, 311–326.
factor using a global dataset. Catena 31, 1–22. Zhang, H.Q., Hartge, K.H., 1995. Mechanical properties of soils influenced by the incorpo-
Torri, D., Santi, E., Marignani, M., Rossi, M., Borselli, L., Maccherini, S., 2013. The recurring ration of organic matter. Adv. in Soil Sci 85, 93–108.
cycles of biancana badlands: erosion, vegetation and human impact. Catena 106, Zhang, B., Zhao, Q.G., Horn, R., Baumgartl, T., 2001. Shear strength of surface soil as affect-
22–30. ed by soil bulk density and soil water content. Soil Tillage Res. 59, 97–106.
Towner, G.D., 1973. An examination of the fall-cone method for the determination of Zimbone, S.M., Vickers, A., Morgan, R.P.C., Vella, P., 1996. Field investigations of dif-
some strength properties of remolded agricultural soils. J. Soil Sci. 24, 470–479. ferent techniques for measuring surface soil shear strength. Soil Technol. 9,
Waldron, L.J., 1977. The shear resistance of root-permeated homogeneous and stratified 101–111.
soil. Soil Sci. Soc. Am. J. 41, 843–849.

You might also like