You are on page 1of 23

International Soil and Water Conservation Research 7 (2019) 203e225

Contents lists available at ScienceDirect

International Soil and Water Conservation Research


journal homepage: www.elsevier.com/locate/iswcr

Review Article

Using the USLE: Chances, challenges and limitations of soil erosion


modelling
Christine Alewell a, *, Pasquale Borrelli a, Katrin Meusburger b, Panos Panagos c
a
Environmental Geosciences, University of Basel, Switzerland
b
Swiss Federal Institute for Forest, Snow and Landscape Research, Birmensdorf, Switzerland
c
European Commission, Joint Research Center, Ispra, Italy

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

Article history: To give soils and soil degradation, which are among the most crucial threats to ecosystem stability, social
Received 10 January 2019 and political visibility, small and large scale modelling and mapping of soil erosion is inevitable. The most
Received in revised form widely used approaches during an 80year history of erosion modelling are Universal Soil Loss Equation
20 May 2019
(USLE)-type based algorithms which have been applied in 109 countries. Addressing soil erosion by
Accepted 29 May 2019
Available online 18 June 2019
water (excluding gully erosion and land sliding), we start this review with a statistical evaluation of
nearly 2,000 publications). We discuss model developments which use USLE-type equations as basis or
side modules, but we also address recent development of the single USLE parameters (R, K, LS, C, P).
Keywords:
Universal soil loss equation
Importance, aim and limitations of model validation as well as a comparison of USLE-type models with
RUSLE other erosion assessment tools are discussed. Model comparisons demonstrate that the application of
CSLE process-based physical models (e.g., WEPP or PESERA) does not necessarily result in lower uncertainties
Soil redistribution compared to more simple structured empirical models such as USLE-type algorithms. We identified four
Soil degradation key areas for future research: (i) overcoming the principally different nature of modelled (gross) versus
Water erosion measured (net) erosion rates, in coupling on-site erosion risk to runoff patterns, and depositional regime,
Model (ii) using the recent increase in spatial resolution of remote sensing data to develop process based
Review
models for large scale applications, (iii) strengthen and extend measurement and monitoring programs
to build up validation data sets, and (iv) rigorous uncertainty assessment and the application of objective
evaluation criteria to soil erosion modelling.
© 2019 International Research and Training Center on Erosion and Sedimentation and China Water and
Power Press. Production and Hosting by Elsevier B.V. This is an open access article under the CC BY-NC-
ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Background the early nineties, it was already estimated that 56% of the global
land being degraded and showed light to severe forms of soil
In a world of climate and land use change, fertile soils are one of erosion by water (Oldeman, 1992, pp. 16e36). Since water erosion is
the most essential resources to sustain humankind. In a recent strongly exacerbated by the conversion of natural vegetation to
review paper in Science (Amundson et al., 2015) soil is discussed agricultural land, with nearly 40% of Earth's land currently utilized
comprehensively as THE essential resource for human security for agricultural production (Foley, 2017), accelerated forms of soil
(including climate and food security) in the 21st century with the erosion became a widespread phenomenon representing a major
main threat to soils being soil erosion by wind and water ever since challenge to achieving the United Nations Sustainable Develop-
humankind had started with agriculture. ment Goals (Keesstra et al., 2016).
To date, most of the world's soils are only in fair, poor, or very The fact that soil erosion is a threat to one of the most essential
poor condition as was stressed by the latest publications of the resources of humankind is as old as Cain and Abel (Panagos et al.,
United Nations, i.e. Status of the World's Soil Resources (FAO, 2015) 2016c). In modern ages, Russian soil scientists were probably the
where soil erosion was identified as one of the major soil threats. In first to be successful in directing attention towards soil erosion
which resulted in governmental mitigation programs at the end of
the 19th century (Dokuchaev, 1892). Today, tools to not only map
the actual status of soil erosion, but also to test the influence of
* Corresponding author.
E-mail address: christine.alewell@unibas.ch (C. Alewell).
mitigation strategies as well as management or conservations

https://doi.org/10.1016/j.iswcr.2019.05.004
2095-6339/© 2019 International Research and Training Center on Erosion and Sedimentation and China Water and Power Press. Production and Hosting by Elsevier B.V. This
is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
204 C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225

practices and last but not least future (climate change) scenarios are Soil Erosion Model (LISEM, (De Roo, Wesseling, & Ritsema, 1996),
urgently needed (Tan, Leung, Li, & Tesfa, 2018). Without the pos- the European Soil Erosion Model (EUROSEM; (Morgan et al., 1998)
sibility to map and visualize soil erosion on large scales, soil erosion and Pan European Soil Erosion Risk Assessment (PESERA, Kirkby
will be out of the focus of all major environmental and agricultural et al. (2008)) resulted in 220 hits (125, 32, 34 and 29, respec-
policies such as the European Union's Common Agricultural Policy tively) of the same study period.
(CAP), the United Nations Sustainable Development Goals (SDGs), One of the main reasons why USLE type modelling is so widely
the United Nations Convention to Combat Desertification (UNCCD) used throughout the world (see also Figs. 1e3 and section 2.1) is
and the Intergovernmental Science-Policy Platform on Biodiversity certainly its high degree of flexibility and data accessibility, a
and Ecosystem Services (IPBES). Moreover, the lack of large scale parsimonious parametrization, extensive scientific literature and
erosion rate data sets also creates knowledge gaps in climate comparability of results allowing to adapt the model to nearly every
change and carbon mitigation scenarios, hydrology and flood pre- kind of condition and region of the world. Nevertheless, the USLE
diction as well as earth science system modelling. approach is an empirical modelling approach with significant
Modelling and prediction of soil erosion by water has a long limitations which already have been addressed in the very first
history with first studies published in international journals more publications as there is no simulation of soil deposition (e.g.,
than seven decades ago using north American data sets (e.g. sedimentation) and that in most cases not enough measured data
(Bennett, 1939; Smith, 1941; Zingg, 1940b)). Many mathematical exist to rigorously determine the single factors for all needed sit-
models categorized as empirical, conceptual, physically-based or uations and scenarios (Wischmeier and Smith, 1965, 1978, p. 58).
process-oriented are available to estimate soil erosion at different The question remains, whether or not research and development of
spatial and temporal scales (De Vente & Poesen, 2005; Jetten, the last 5 decades (1965 e today) was able to overcome these main
Govers, & Hessel, 2003; Karydas, Panagos, & Gitas, 2014; King & limitations, or, if not, at least improved them to such an extent that
Delpont, 1993; Merritt, Letcher, & Jakeman, 2003; Morgan, 2005, is justifiable to apply the model algorithm to large scales and if so,
pp. 365e376; Toy T. J, 2002; Vrieling, 2006; Zhang, O'Neill, & Lacey, under which conditions and resolution.
1996a). Most erosion models contain a mix of modules from each of The aim of this review is a thorough and rigorous evaluation of
these categories. The research community has been contempora- USLE-type modelling to come to some conclusions on the sense and
neously working for both, improving the applicability of complex non-sense of soil erosion modelling with this model concept
process-oriented models such as the Water Erosion Prediction depending on specific circumstances and conditions. We do not
Project (WEPP) (Boardman, 2006; Morgan & Nearing, 2011) or the aim to provide a user guide or handbook of USLE-type modelling
European Soil Erosion Model (EUROSEM; Morgan et al. (1998)) as but ask the interested reader to depend in this respect on the
well as updating the existing empirical approaches such as the wealth of literature starting from the original publications
Universal Soil Loss Equation (USLE) (Bagarello et al., 2008, 2015; (Wischmeier and Smith, 1965, 1978, p. 58) to reviews on its history
Bagarello & Ferro, 2004; Di Stefano, Ferro, & Pampalone, 2017; and past use (e.g. (Laflen & Flanagan, 2013; Laflen & Moldenhauer,
Kinnell, 2014) which not only remains attractive from a practical 2003)), on special applications and further development (e.g.
point of view (Cao, Zhang, Dai, & Liang, 2015; Gessesse, Bewket, & consideration of runoff and event based modelling (Kinnell, 2010))
Bra€uning, 2015) but also gives estimates on large spatial scales (see to general evaluations of model concepts and usability (Merritt
below, (Diodato, Borrelli, Fiener, Bellocchi, & Romano, 2017; et al., 2003; Nearing, 2004; Quinton, 2013). We also do not claim
Doetterl, Van Oost, & Six, 2012; Van Oost et al., 2007; Yang, Kanae, or even make an attempt to review all existing studies using USLE-
Oki, Koike, & Musiake, 2003a)). type models which is clearly beyond the scope of a manuscript in
At present, USLE and the Revised USLE (RUSLE) are by far the ESR but have included a mere numerical geographical meta-
most widely applied soil erosion prediction models globally and analysis to embrace the wealth of literature published on USLE-
according to (Risse, Nearing, Laflen, & Nicks, 1993) “USLE has been type models quantitatively (see section 2.1).
used throughout the world for a variety of purposes and under
many different conditions simply because it seems to meet the
2. Historical aspects and evolution of USLE-type modelling
need better than any other tool available”. A Web of Science query
(http://apps.webofknowledge.com/) for the period 2007e2017
The origin of USLE-type models was in the US to provide a
resulted in 1149 hits corresponding to the keywords “Universal Soil
management decision support tool and was based upon thousands
Loss Equation”, “USLE”, “Revised Universal Soil Loss Equation” or
of controlled studies on field plots and small watersheds since 1930
“RUSLE”, with average citations per item of 7.8 (as of August 2017).
(Wischmeier & Smith, 1965). As with all empirical methods the
A query with the Science Direct tool for the last 40 years resulted in
model concept is not based on process description and simulation
1556 studies using USLE or RUSLE with an average citation rate of
but rather on understanding a process, capturing the confounding
cited publications of 18.8 (studies with no citations were not
measureable parameters and delineating a mathematical algorithm
considered in this average rate, see meta-analysis in section 2.1).
out of the relationship between these parameters and the
The most well-known models based on USLE/RUSLE technology are
measured output (in this case measured eroded sediments).
the Soil and Water Assessment Tool (SWAT) (Arnold, Srinivasan,
As such, the USLE was defined as:
Muttiah, & Williams, 1998), the AGricultural Non-Point Source
Pollution Model (AGNPS) (Cronshey & Theurer, 1998; Young, A ¼ R,K,L,S,C,P (1)
Onstad, Bosch, & Anderson, 1989), and the Water and Tillage
Erosion and Sediment Model (Watem/Sedem) (Van Rompaey, where: A (Mg ha1 yr1) is the annual average soil erosion, R (MJ
Verstraeten, Van Oost, Govers, & Poesen, 2001) and the Chinese mm h1 ha1 yr1) is the rainfall-runoff erosivity factor, K (Mg h
Soil Loss Equation (CSLE, Liu, Zhang, and Xie (2002)) which total- MJ1 mm1) is the soil erodibility factor, L (dimensionless) is the
ized 1385, 311, 67 and 35 hits, respectively, in the same period (Web slope length factor, S (dimensionless) is the slope steepness factor,
of Science query 2007e2017). Modelling approaches independent C (dimensionless) is the land cover and management factor, P
from the USLE such as the Water Erosion Prediction Model (WEPP, (dimensionless) is the soil conservation or prevention practices
Laflen, Elliot, Flanagan, Meyer, and Nearing (1997)), the Limburg factor. The impact of the factors R capturing the energy and amount
C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225 205

Fig. 1. Geographic distribution per country of published papers using USLE or RUSLE modelling during the last 40 years (1977 to July 2017).

sediment yield.
Even though Wischmeier and Smith (1965) are generally cited
as the origin of the equation, the development of the modelling
concept actually started much earlier. The soil-loss equation was
the result of 20 years of development starting from the so called
slope-practice method (Zingg, 1940b) on which Smith (1941) added
crop and conservation-practice factors. Eventually soil erodibility,
management factors (Browning, Parish, & Glass, 1947) and the
rainfall factor (Musgrave, 1947) were added. The resulting “Mus-
grave equation” was published as a graphical solution in 1952
(Lloyd & Eley, 1952) but was further improved throughout the
1950's with an improved rainfall-erosion index, a method of eval-
uating cropping-management effects on the basis of local climatic
conditions, a quantitative soil-erodibility factor and a method of
accounting for effects of interrelations of such variables as pro-
Fig. 2. Number of studies per continent and percentage of total publication number ductivity level, crop sequence and residue management (USDA,
(1977 to July 2017).
1961; Wischmeier, 1959; Wischmeier, 1960). Wischmeier and
Smith (1965) with an update by Wischmeier and Smith (1978, p.
58) finally put figures, tables and equations for all five single fac-
of precipitation, K accounting for the soil parameters determining
tors of the USLE together in a guide book explaining application and
erosion potential, C describing the vegetation cover and manage-
use in detail. This guidebook was published by the US National
ment as well as P delineating human management intervention is
Runoff and Soil Loss Data Center, in cooperation between the
directly related to our process understanding of soil erosion. Un-
Agricultural Research Service and Purdue University, and it resulted
derstanding erosional processes and driving factors, it might seem
from statistical analysis of more than 10,000 plot-years of runoff
surprising that neither runoff nor infiltrating was included in the
and soil loss data carried out in plots having a length less than or
algorithm. However, from a pure statistical perspective the rainfall
equal to 122 m and a slope ranging from 3 to 18%. For defining the
erosivity was simply the best predictor for the measured erosion
mathematical structure of the USLE a reference condition, named
output and attempts to include runoff even reduced the quality of
as unit plot, was used, which built on the set-up of field experi-
the assessment (Wischmeier, 1966; Wischmeier & Smith, 1965).
ments which were commonly used. The unit plot was defined as a
From a soil scientific perspective infiltration is not a helpful
22.1 m long plot, with a 9% slope, maintained in a continuous
parameter for such a modelling endeavour, because it is very prone
regularly tilled fallow condition with up-and-down hill tillage. The
to measurement errors, extremely variable in soils and we will
unit plot was used to compare soil loss data collected on plots that
hardly ever be able to capture infiltration at larger scales. The
had different slopes, lengths, cropping and management and con-
implementation of the L and S are meant to capture the impact of
servation practices (Wischmeier and Smith, 1965, 1978, p. 58).
runoff energy which is influenced by a mix of processes and pa-
The main motivation of the early studies using the USLE were to
rameters. However, it has to be noted that in all these endeavours
quantify soil erosion rates and their single contributing factors in
modelling targeted at on site soil erosion risk, not at catchment
206 C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225

Fig. 3. Trend of published papers (USLE/RUSLE). The period 2013e2017 covers until July 2017, the projection for the full 5 year period would approximate 600 papers (striped box).

comparison to soil-loss tolerance values and assess possible com- Johnson, Savabi, & Loomis, 1984). Johnson et al. (1984) suggested
binations of cropping systems and management plans for mitiga- that a more accurate assessment of the cover and management
tion (Schwertmann, Vogl, & Kainz, 1987, pp. 1e64; Wischmeier & conditions is needed for applications of the USLE on rangelands.
Smith, 1965). Weltz, Kidwell, and Fox (1998) stated that ‘USLE is a lumped
Since its first definition, USLE-type modelling was developed empirical model that does not separate factors that influence soil
further to meet the multiple needs and conditions of modelled erosion, such as plant growth, decomposition, infiltration, runoff,
systems. Renard, Foster, Weesies, McCool, and Yoder (1997) stated soil detachment, or soil transport’. The increasing criticisms with
that the USLE for more than four decades has proven that this regard to the limitations of the USLE were met by an increasing
technology is valuable as a conservation-planning guide in the US, interest in both the US action agencies and the soil science research
providing farmers and conservation planners with a tool to esti- community to update the USLE thus creating a substantial need for
mate rates of soil erosion for different cropping systems and land a revision of the USLE which were made possible by the fast im-
managements. Even though conceptualized and calibrated for provements in computation capacity.
agricultural areas, the USLE has been soon adjusted to extend its Accordingly, the USLE was upgraded to the Revised Universal
applicability to undisturbed land. In the early 1970's, as a result of a Soil Loss Equation (RUSLE) during the 1990s (Renard et al., 1991,
meeting between the Natural Resources Conservation Service 1997, pp. 1e404). The basic structure of the RUSLE retains the
(NRCS) and the Forest Service of the USDA, a first extension of the multiplicative form of the USLE, but it also has process-based
USLE was made by developing a sub-factor to compute the cover- auxiliary components such as calculating time-variable soil erod-
management factors (C-factor) for woodland, permanent pasture ibility, plant growth, residue management, residue decomposition
and rangeland (Spaeth, Pierson, Weltz, & Blackburn, 2003; and soil surface roughness as a function of physical and biological
Wischmeier, 1975, pp. 118e124). The proposed method to estimate processes. RUSLE also has updated values for erosivity (R), new
the USLE C-factor used relationships to extrapolate three major relationships for the topographical components (L and S factors)
sub-factors, canopy, ground cover and below soil effects (Renard & which include ratios of rill and inter-rill erosion, consideration of
Foster, 1985), since for these areas an extensive data base of the C seasonality of the K factor and additional P factors for rangelands
factor was not available. Later, Dissmeyer and Foster (1980) pro- and subsurface drainage, among other improvements (please see
posed further additions and modifications to extend the number of section 3 for description of advancement regarding the single fac-
sub-factors operating in woodland, including (i) amount of bare soil tors of RUSLE).
(or conversely ground cover), (ii) canopy, (iii) soil reconsolidation, RUSLE combines the advantage of being based on the same
(iv) high organic content, (v) fine roots, (vi) residual binding effect, extensive database as was the USLE with some process-based
(vii) onsite storage, (viii) steps and (ix) contour tillage (agrofor- computations for time-varying environmental effects on the
estry). As reported by Spaeth et al. (2003) early applications of the erosional system (Nearing, 2004). However, it still has the limita-
model in rangelands showed little agreement between estimated tions in model structure that allows only for limited interactions
and measured soil loss in both catchment (Simanton, Roger, and inter-relationships between the basic multiplicative factors
Herbert, & Renard, 1980) and plot experiments (Hart, 1984; (Nearing, 2004). As USLE-type models were designed to predict
C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225 207

long-term average annual soil loss, they have been successful to of the RUSLE published by the Research Branch Agriculture and
predict event soil losses reasonably well at some geographic loca- Agri-Food Canada (Wall, Coote, Pringle, & Shelton, 2002, pp.
tions (Kinnell, 2010), but often fail to predict event erosion, which is 1e117)); Australia (Ferro & Porto, 1999); Africa (Haileslassie, Priess,
highly influenced by the fact that the USLE and its revisions (RUSLE) Veldkamp, Teketay, & Lesschen, 2005; Lufafa, Tenywa, Isabirye,
do not consider runoff explicitly. Majaliwa, & Woomer, 2003) and Asia (Lee & Heo, 2011;
RUSLE2 was developed to scientifically enhance the USLE/RUSLE Meusburger, Mabit, Park, Sandor, & Alewell, 2013; Xu, Miao, & Liao,
equations and offer an improved tool to guide and assist erosion- 2008). In China, the USLE was adapted in substituting the C and the
control planning (USDA, 2008). Although the fundamental empir- P factor with three factors considering the biology, engineering and
ical equation scheme of the previous equations was retained, tillage practices (Liu et al., 2002). For the application in Europe
RUSLE2 uses both empirical and process-based equations that (Schwertmann et al., 1987, pp. 1e64) was the first to rigorously
allow it to extend significantly beyond the original USLE structure. evaluate the applicability of the USLE to the different climatic
In fact, RUSLE2 can be defined as a hybrid soil erosion prediction conditions and land use management systems of Bavaria (Southern
(estimation) approach since it is a combination of the empirical, Germany). Today, USLE-type modelling is being used in Europe
index-based USLE and process-based equations for the detach- from the northern Scandinavia to the Southern Mediterranean or
ment, transport, and deposition of soil particles (USDA, 2008). The eastern countries like Turkey and Hungary (e.g., (Bagarello & Ferro,
RUSLE2 scheme allows to compute net erosion or deposition (mass/ 2004; Ferro & Porto, 1999; Ozsoy, Aksoy, Dirim, & Tumsavas, 2012;
area) for each segment in which the overland flow path is divided, Podmanicky et al., 2011; Porto, 2016; Sivertun & Prange, 2003)).
sediment load (mass/unit flow width) at the end of each segment Eventually USLE-type modelling was increased to continental
and at the end of the overland flow path, and sediment charac- (Europe: Panagos et al. (2015d); China: Yue et al. (2016); Australia
teristics at the detachment point and in the sediment load at the Teng et al. (2016)) and finally to global scale (Borrelli et al., 2017b;
end of each segment (USDA, 2008). Among other structural im- Diodato et al., 2017; Doetterl et al., 2012; Ito, 2007; Van Oost et al.,
provements (USDA, 2018), RUSLE2 works on a daily time step and 2007; Yang, Kanae, Oki, Koike, & Musiake, 2003b).
introduced the concept of erosivity density which resulted in sig- A database of studies was developed within the frame of this
nificant improvements in the calculations and mapping of rainfall review using the Science Direct tool searching for the keywords
erosivity (Nearing et al., 2017). Validation of the parameters of the USLE and RUSLE (from 1977 to July 2017) with specific focus on title,
original RUSLE2 was achieved with the same 10,000 plot-years data year of publication, study area, journal, paper type and keywords.
as the original USLE. Additional 451 studies were not included as they did not propose a
As USLE has been frequently criticized of its empirical nature, specific study area or addressed theoretical issues of the model. The
Ferro (2010) demonstrated that the original structure of USLE can author keywords of the articles have been processed by a word
be theoretically obtained applying the dimensional analysis and the cloud application to produce the most common issues addressed by
self-similarity theory (Barenblatt, 1979, 1987) using the same soil authors. In this process, we excluded self-explained keywords
erosion representative variables and the reference condition (USLE, RUSLE, soil, erosion) and geographical locations. Besides the
adopted by Wischmeier and Smith (1965). In other words, using the most common keywords (GIS, model, factor, map, risk) the highest
factor scheme and the reference condition adopted by Wischmeier frequency was reached by keywords such as climate, geo-statistics,
and Smith (1965), Ferro (2010) challenges the criticism of the policy, crop, terraces, 137Cs, planning and nutrient (see visualization
empirical origin in claiming that “USLE is the subsequent logical in the graphical abstract).
structure with respect to the variables used to simulate the physical The USLE/RUSLE model has been extensively used during the
soil erosion process”. However, with this statement, we always last 40 years in 109 countries (Fig. 1).
need to keep in mind, that USLE edriven modelling targets at the The largest number of publications with the application of USLE/
physical soil erosion process and will thus not be capable of RUSLE model has been found in the United States of America (274
simulating catchment sediment transport or yields. papers), China (218 papers), Brazil (88), Italy (87), India (67), Spain
Today, USLE-type modelling has been further advanced to meet (66) Australia (50) and Turkey (43). In an analysis per continent, 519
numerous special requirements and specific needs. E.g., Bagarello, papers (33% of total) have study sites in 32 countries of Asia (Fig. 2).
Di Stefano, Ferro, and Pampalone (2017) as well as Larson, In Europe (24% of the total number of studies) USLE/RUSLE is used
Lindstrom, and Schumacher (1997) adapted USLE-type models for in 31 countries and in Northern America and the Caribbean in 13
event based soil erosion modelling. USLE-type modelling has also countries (22% of total number of studies). In Africa, 146 publica-
been used in all kind of extreme ecosystem types and for various tions (9% of total) used USLE/RUSLE for estimating soil loss by water
management scenarios, e.g. from volcanic soils in Chile with in 21 countries. The meta-analysis resulted also in a wide use of the
Mediterranean climate (Stolpe, 2005) to the possible mitigation USLE/RUSLE model in East-Southern Asia (South Korea, Malaysia,
impact of organic farming on soil erosion rates from mountainous Thailand and Indonesia), Central East Africa (Ethiopia, Kenya,
monsoonal watersheds in South Korea (Arnhold et al., 2014) or the Nigeria) and the Mediterranean (Italy, Spain, Greece, Portugal,
comparison of conventional with organic farming in northern Morocco, Tunisia, Algeria) (Fig. 1).
Bavaria (Auerswald, Kainz, & Fiener, 2003). For a discussion of ad- The 1556 papers that estimated soil loss by water erosion using
vancements in model development and extension of the model USLE/RUSLE at local/regional/national and continental scale have
concept please see section 3. been published with an average rate of 39 papers per year during
the period 1977e2017. The split into time windows of 5-years
yielded a strongly increasing trend with the highest number of
2.1. A geographical and temporal meta-analysis of publications published articles (131) in 2016 (extrapolation from July 2017 to the
addressing USLE/RUSLE models full year of 2017 is approximating 150 papers). A sharp increase in
publication rates started in the late 90ties. The highest relative
The USLE concept, originally developed for the US agricultural increase (76%) is noted in the period 2008-12 which is also ex-
systems, has been adapted by scientists all over the world and pected to continue in the current period (2013-17). Estimated
applied to datasets of different regions, countries and continents. published papers during the last 5 years will be more than 600
Examples would be Canada (e.g., the handbook for the application
208 C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225

compared to the 414 papers published in 2008-2012 (Fig. 3). implemented in new modelling concepts considering runoff,
USLE/RUSLE applications were published in a wide variety of sediment transport and/or sediment delivery. An implementation
journals. The 1556 peer reviewed papers have been published in of the USLE where the rainfall erosivity is replaced by runoff vol-
more than 500 Science Direct indexed journals with the highest ume (Modified USLE, MUSLE) resulted in a satisfying prediction of
frequency in Catena (4.2%), followed by the Journal of Soil & Water measured sediment yield already in the mid-seventies and early
Conservation (3.3%) and Land Degradation and Development eighties (Smith, Williams, Menzel, & Coleman, 1984; Williams,
(2.6%). More than 83% of the papers are classified as research arti- 1975, pp. 244e252). The latter allows the equation to be applied
cles, followed by conference papers (14%), book chapters (2%) and to individual storm events. Several variants of the MUSLE equation
review papers (1%). Finally, 1195 articles (77% of the total) were exist and MUSLE was even integrated in GIS (e.g. in ArcMUSLE;
cited with an average of 18.8 citations per article (publications with (Zhang, Degroote, Wolter, & Sugumaran, 2009). However, even
no citations were not considered in this average rate). though runoff is considered, the restriction that USLE-modelled
and observed data represent different parameters (e.g., gross and
3. Developments of USLE-type modelling with extended net erosion rates, respectively, please see box 1 and also section 5)
modelling concept are not yet overcome. Another event based derivate of USLE is
USLE-M, which also includes event runoff in R, but K-factor values
Numerous advanced developments of USLE-type modelling are adjusted accordingly by multiplying them with the ratio of the
have been implemented during the last decades (for selections total value of the EI30 index to the total value of the QREI30 index
please see Table 1). A substantial change to the USLE concept were (Kinnell & Risse, 1998).

Table 1
Selected models based on the USLE concept or incorporating USLE parameters.

Model Full name Original references USLE components Additional features compared to USLE

RUSLE Revised Universal Soil Loss (Renard et al., 1991, 1997, USLE components to process-based auxiliary components (e.g., time-variable soil
Equation pp. 1e404) simulate erosion erodibility, plant growth, residue management, residue
decomposition, soil surface roughness, updated values for
erosivity (R), new relationships for L and S factors,
additional P factors
RULSE2 Revised Universal Soil Loss USDA (2008) USLE components to Net erosion and deposition, hybrid soil erosion prediction
Equation 2 simulate erosion with combination of the empirical, index-based USLE and
process-based equations for the detachment, transport, and
deposition of soil particles, calculation on a daily basis,
biomass accounting in the C factor,
Adding three dimensional components
RUSLE3D Revised Universal Soil Loss (Desmet & Govers, 1996; USLE components but Considering the flow convergence of the upslope area thus
Equation for Complex Terrain Mitasova et al., 1996; replacing the slope length improving the impact of concentrated flow on increased
Mitasova & Mitas, 1999) with contribution of erosion
upslope area
USPED Unit Stream Power Erosion and Mitasova et al. (1996) USLE components to Considering 3-D topography and prediction erosion and
Deposition simulate erosion deposition
Adding factors to consider nutrient transport, management measures or replacement of C factors to improve biological dynamic
CSLE Chinese Soil Loss Equation Liu et al. (2002) USLE but C and P factor C factor replaced by a biological conservation measures
replaced factor (B), and engineering conservation measures factor (E)
as well as a tillage conservation measures factor (T)
PERFECT Productivity, Erosion and (Littleboy et al., 1992a, Sediment yield is simulated Water balance and runoff predictions, erosion and crop
Runoff Functions to Evaluate 1992b) using MUSLE growth and crop yield; including sequences of plantings,
Conservation Techniques harvests and stubble management during fallow
G2 Geoland 2 erosion model (Karydas & Panagos, 2016; USLE but C and P factor C factor replaced by a vegetation factor retention combining
Panagos, Christos, Cristiano, replaced. land use and fractional vegetation coverage; P factor
& Ioannis, 2014a) replaced by landscape features factor quantifying the effect
of obstacles to interrupt rainfall runoff. Model erosion at
monthly step
Considering runoff, sediment transport and delivery
MUSLE Modified Universal Soil Loss (Smith et al., 1984; USLE components to prediction of sediment yield; simulation of individual storm
Equation Williams, 1975, pp. 244 simulate erosion events
e252)
SWAT Soil and Water Assessment Tool Arnold et al. (1998) Sediment delivery based on Simulation of the hydrological water balance, then using the
MUSLE runoff to simulate sediment dynamics
WATEM/SEDEM Water and Tillage Erosion and (Van Oost et al., 2000a; Van USLE components to sediment production as well as sediment transport and
Sediment Model Rompaey et al., 2001) simulate erosion pathways to derive ultimate source strengths, consideration
of tillage erosion
SEDD Sediment Delivery Distributed Ferro and Minacapilli USLE-type modelling to Gross erosion rates are coupled to a sediment delivery tool
Model (1995) calculate gross erosion to estimate sediment transport and delivery
rates
Considering runoff and sediment dynamics as well as nutrients and/or pollutants
CREAMS Chemical Runoff and Erosion (Knisel, 1980; Silburn & USLE components to Runoff modelling, erosion and sedimentation, gully erosion;
from Agricultural Management Freebairn, 1992; Silburn & simulate erosion chemistry and target non-point source pollutants
Systems model Loch, 1989) Individual storms/events as well as long term modelling
AGNPS Agricultural Non-Point Source Young et al. (1989) USLE or RUSLE used to Runoff and nutrient modelling
Model model soil erosion
EPIC Erosion Productivity Impact Williams, Renard, and Dyke USLE used to model water in addition to water erosion, hydrology, nutrient dynamics,
Calculator (1983) erosion plant growth, soil temperature tillage and economics are
simulated with physically based modules
C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225 209

Box 1 erosion modes, namely WaTEM (Water and Tillage Erosion Model;
Definition of soil erosion in Universal Soil Loss Equation (USLE)- (Van Oost et al., 2000a) and SEDEM (Sediment Delivery model, Van
type algorithms Rompaey et al. (2001).
A similar approach is followed with the Sediment Delivery
Distributed Modell (SEDD) where USLE-type modelling is used to
Soil erosion is, strictly speaking, the detachment of soil calculate spatially distributed gross erosion rates, which are then
particles from a source site and its transport to some coupled to a sediment delivery tool to estimate net sediment
depositional sink. As such, the processes of erosion and transport and delivery on catchment scale (Ferro & Minacapilli,
deposition (often also addressed as sedimentation) are al- 1995; Ferro & Porto, 2000).
ways coupled; both terms together might be described as The above discussion clearly demonstrates that USLE e type
soil redistribution. models have long ago left the pure empirical modelling realm and
Soil redistribution encompasses soil redistribution by water process-based elements have been incorporated into some of the
and wind as well as mass transfer of soil particles (e.g. advanced model development approaches. However, the question
landslides). Soil redistribution by water can further be arises how far this route can be followed if at the same time large
differentiated into transport of single grains such as sheet scale modelling is envisaged, needing spatially discrete and inde-
or interrill erosion, rill and gully erosion. In the modelling pendent units. As such, any modelling endeavor must, at the stage
concepts of all USLE-type algorithms soil erosion is defined of planning, clearly define envisaged aims and available data, to be
as: “Soil loss refers to the amount of sediment that reaches able to choose the best suited modelling concept (see also section 6,
the end of a specified area on a hillslope that is experiencing Model Comparison).
net loss of soil by water erosion. It is expressed as a mass of
soil lost per unit area and time. Soil loss refers to net loss, 4. State of development of the USLE factors
and it does not in any way include areas of the slope that
experience net deposition over the long term.” (Nearing, As stated above, the origin of the USLE was strictly empirical
Yin, Borrelli, & Polyakov, 2017). USLE-Type modelling with the calibration of the single factors from basic erosion plot
does not address larger rills or gully erosion (linear struc- treatments (see section 2). However, this basic structure was sub-
tures with a depth > 30 cm), but is restricted to sheet/interrill stantially modified over the last decades with many process-based
and small rill erosion only (<several cm deep, generally auxiliary components in each of the five single factors accompanied
defined as structures which might be removed by by integration into high resolution Geographic Information Sys-
cultivation.). tems (GIS) and the application of state of the art geostatistical tools.
As such, the aim of this section is to present the progress of each of
the five USLE parameters from their original definition to todays
advanced and highly developed possibilities.
For a better consideration of the terrain shape and complex
topographic conditions of the upslope area, the hillslope length 4.1. R factor
factor was replaced by the upslope contributing area in various
forms (Desmet & Govers, 1996; Mitasova, Hofierka, Zlocha, & The development of the R factor was one of the major im-
Iverson, 1996), an approach which was later called RUSLE3D (e.g., provements of the USLE compared to the original Musgrave equa-
(Mitasova & Mitas, 1999; Rodriguez & Suarez, 2012)). A modified tion. After analyzing c.a. 8,000 plot-years of runoff, soil loss and
equation for computation of the LS factor in finite difference form in rainfall, Wischmeier (1959) defined a “Rainfall Erosion Index” as a
a grid cell representing a hillslope segment was derived by Desmet product of the total kinetic energy of the storm and its maximum
and Govers (1996). A simpler, continuous form of the equation for 30-min intensity (EI). Consequently, the so called iso-erodent maps
computation of the LS factor at a point on a hillslope was developed (Wischmeier & Smith, 1965) were produced from rainfall data of
by Mitasova et al. (1996). Both applications demonstrated that the about 2000 locations fairly evenly distributed over 37 States of the
upslope area-based factor improves the impact of concentrated US using 22-yr station rainfall records and computing the EI for
flow on increased erosion. Mitasova et al. (1996) also added a 3-D each storm. The R-factor was included as one of the inputs in the
dimensional enhancement to the USLE, the Unit Stream Power USLE model (Wischmeier & Smith, 1978, p. 58) using a logarithmic
Erosion and Deposition (USPED) model which enabled to predict function between Kinetic Energy (KE) and Intensity (I) plus a con-
erosion as well as deposition. Based on a comparison to 137Cs based stant value for intensities exceeding 76 mm h1. Kinnell (1981)
erosion/deposition measurements (Warren et al., 2005) concluded proposed an exponential relationship of KE-I compared to the
that USPED derived estimates were more accurate and less biased logarithmic form and this change was also confirmed by Brown and
than USLE applications. However, a comparison of USPED and Foster (1987).
RUSLE3D for a region in the Mediterranean (Bradano basin, Italy) RUSLE used the proposed exponential relationship for esti-
demonstrated for both models their feasibility to estimate the mating the unit rainfall energy (er) based on rainfall intensity (ir):
spatial distribution of soil loss, estimates for soil erosion/deposition h i
at the watershed scale and a good match with the measured silting er ¼ 0:29 1  0:72eð0:05ir Þ (2)
data (Aiello, Adamo, & Canora, 2015).
A combination of RUSLE with sediment transport and delivery Accordingly, the calculation of rainfall erosivity (EI30) of a single
processes has been achieved with the Water and Tillage Erosion event was based on the following equation:
and Sediment Model (WaTEM/SEDEM (Van Oost, Govers, &
Desmet, 2000a; Van Oost, Govers, Van Muysen, & Quine, 2000b; X
k 
Van Rompaey et al., 2001)), which describes not only soil erosion EI30 ¼ er vr (3)
I30
r¼1
risk, e.g. risk for sediment production, but also considers sediment
transport and pathways to derive ultimate source strengths. The With vr the rainfall volume (mm) during the rth time period of a
spatially distributed soil erosion and sediment delivery model rainfall event divided in k-parts. I30 is the maximum 30-min rainfall
WaTEM/SEDEM is a combined version of two empirically-based soil intensity (mm h1). Finally, the mean annual erosivity sums the
210 C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225

erosive events during a given period and is divided by the number influence erodibility by water which are those that affect the
of years. In RUSLE2 a slight modification of unit rainfall energy infiltration rate, permeability, and total water capacity, as well as
function (Yin, Xie, Liu, & Nearing, 2015) was proposed but its use those that might influence the dispersion, splashing, abrasion, and
has been mostly limited to the United States. Equation (3) is mostly transporting forces of the rainfall and runoff. The soil erodibility
used in R-factor estimations elsewhere in the world. factor K was at first an empirical value determined from 20 years of
The varying relationship between rainfall kinetic energy (KE) data on experimental plots from 23 major soil types from the US
and intensity (I) have been addressed in a recent study by Angulo- kept fallow for at least 2 years with all other factors kept constant
Martinez, Begueria, and Kysely (2016). In the absence of KE mea- (Wischmeier & Smith, 1965) and soil erodibility was regarded as
surements (e.g., measuring raindrop size and velocity with dis- the amount of soil loss per unit erosive force with K equal to A/R.
drometers) KE is usually estimated with empirical equations from Since the direct measurement of the K-value requires the estab-
measured rainfall intensity. Angulo-Martinez et al. (2016) evalu- lishment and maintenance of natural runoff plots over lengthy,
ated 14 different KE-I equations to estimate the 1 min KE and event expensive observation periods at various locations, numerous at-
total KE and compared these results with 821 observed rainfall tempts have been made to simplify the technique and to establish
events recorded by an optical disdrometer. They concluded that (i) estimators for soil erodibility calculation from readily available soil
empirical relationships performed well, when complete events property data and standard profile description (Wischmeier,
were considered but performed poorly for within-event variation Johnson, & Cross, 1971).
(1 min resolution) and (ii) to use local measured data or local ki- The nomograph considering the four most crucial soil parame-
netic energy equations. In a recent global R factor assessment 97.7% ters (particle-size, percent organic matter, soil structure and soil
of the calculated R-factors stations were based on the original permeability) was suggested to derive the K-factor (Wischmeier
equation of Brown and Foster (1987) and many of those were et al., 1971). Later, an approximation equation of the nomograph
supported by local observation and validation (Panagos et al., was developed (Wischmeier & Smith, 1978, p. 58):
2017b).
The impact of snow and snowmelt was not considered in R- K ¼ 2.77*105*M1,14*(12- a)þ 0.043 (b- 2) þ 0.033 (4- c) (4)
factor equations but Schwertmann et al. (1987, pp. 1e64) suggested
a very rough approximation in adding 1/10 of the precipitation for where M ¼ the particle-size parameter, which equals % silt (0.1-
those month, where significant soil movement due to snow melt is 0.002 mm) times the quantity (with the quantity defined as 100 - %
observed. Recent approaches subtracted the precipitation amount clay), a ¼ percent organic matter, b ¼ the soil-structure code used in
that falls below 0  C (Meusburger, Steel, Panagos, Montanarella, & soil classification, and c ¼ the profile-permeability class. However,
Alewell, 2012). The later approach considers that snowfall does Wischmeier and Smith (1978, p. 58) state that this equation did not
not exert erosive energy, but neglects the subsequent process of describe the entire nomograph and differs for soils having high silt
snow melt that may trigger soil erosion. content, low erodibility or high organic matter content.
An open point of discussion regarding equation (3) is the rain (Auerswald, Fiener, Martin, & Elhaus, 2014) developed a set of
drop size. In tropical areas such as Ethiopian highlands the algo- equations mimicking the original K factor nomograph. The de-
rithm for unit rainfall energy (equation (2)) may underestimate the viations from the original equation were tested on 19055 soils
rainfall erosivity due to large drop sizes in those areas (Nyssen et al., obtained during soil surveys throughout Germany. The set of
2005). equations proposed resulted in deviations compared to the classical
In the erosion studies applying USLE/RUSLE in late 1990s and K factor equation in more than 50% of all cases. Thus, the authors
early 2000s, many scientists have either applied simplistic multi- recommend using this set of equations, which describe the
plications for estimating erosivity (R-factor ¼ 1.3 * precipitation) or nomograph even beyond the limitations of the classical equation,
have proposed empirical erosivity equations such as the Fournier but they restrict that predictions may be far from “perfect” since
Index (FI) and its modification (MFI) by Arnoldus (1980). The lack of these equations are ignoring seasonality or interaction with
stations with sub-hourly data for long-periods and measured climate. Generally, the nomograph should only be used if there is
rainfall intensity forced scientists to develop simple empirical not a locally derived relationship or measured data that would also
equations correlating R-factor with available daily/monthly/annu- account for seasonal changes in the K-factor.
ally rainfall data (Aronica & Ferro, 1997; Bonilla & Vidal, 2011; Seasonal effects on the K-Factor due to freezing and thawing
Diodato, Knight, & Bellocchi, 2013; Loureiro & Coutinho, 2001). processes (accompanied by effects on shear strength) or on
Recently, the availability of high temporal resolution data compaction (e.g., by rainfall during winter/fall or by life stock
(hourly, sub-hourly) for long periods and the development of geo- trampling during summer), and subsequent release processes have
statistical algorithms for spatial interpolations contributed to been discussed (Renard et al., 1991; Renard & Ferreira, 1993).
regional, national and continental erosivity maps in Spain (Angulo- However, none of these seasonal processes and effects can easily be
Martinez, Lopez-Vicente, Vicente-Serrano, & Begueria, 2009), captured for inclusion in the K factor. Kinnell (2010) reviewed
Switzerland (Meusburger et al., 2012; Schmidt, Alewell, Panagos, & different approaches to assess the seasonality of the K-factor.
Meusburger, 2016), Italy (Borrelli, Diodato, & Panagos, 2016a), Ko- However, none of these approaches include the hardly measurable
rea (Risal et al., 2016), Iran (Sadeghi, Zabihi, Vafakhah, & Hazbavi, influencing interactions and effects (e.g., climate influences and
2017), Brazil (Oliveira, Wendland, & Nearing, 2013) and the Euro- seasonality of freeze-thaw, compaction by life stock trampling,
pean Union (Panagos et al., 2015a). weathering, human management activities,) simultaneously for a
A group of scientists, working with R-factor data all over the proper process-oriented modeling (Leitinger, Tasser, Newesely,
world, has collected data on rainfall erosivity from 3,625 meteo- Obojes, & Tappeiner, 2010; Pineiro, Paruelo, Oesterheld, &
rological stations in 63 countries and established the Global Rainfall Jobbagy, 2010; Vannoppen, Vanmaercke, De Baets, & Poesen,
Erosivity Database (GloReDa) developing a global erosivity map 2015). Furthermore, the divergence of seasonal K-factors to an
(Panagos et al., 2017a). annual K-factor is poorly discussed in the literature (e.g.(Wall,
Dickinson, Rudra, & Coote, 1988)). In the RUSLE2 User's Reference
4.2. K factor Guide (Foster et al., 2008) it is even stated that no statistical evi-
dence exists for an inconsistency of soil erodibility over time.
The K factor is statistically related to the soil properties that Another important parameter affecting the erodibility is stone
C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225 211

content. Originally the effect of surface stones was not included in the factor slope length (L) and slope steepness (S). The original
the K factor equation, instead (Wischmeier & Smith, 1978, p. 58) value of S was derived empirically as S ¼ (0.43 þ 0.30s þ 0.043s2)
suggested to consider the effect of surface stones in the C factor. The *6.613 with s ¼ slope gradient in percent (Smith & Wischmeier,
effect of stone cover on soil loss is scale dependent. At the meso- 1957; Wischmeier & Smith, 1965). Previous parametrizations
scale stones and rock fragments on the soil surface may have (Musgrave, 1947; Smith & Whitt, 1948; Zingg, 1940b) and many
ambivalent effects while at the micro and macroscale a decrease in posteriori derivatives (Fig. 4) of the original empirical equation
sediment yield is predominant (Poesen, Torri, & Bunte, 1994). For with different curve shapes of S against slope angle have been
the macroplot scale, Poesen et al. (1994) developed a soil erodibility developed since then (Foster, 1982, pp. 17e35; Liu, Nearing, & Risse,
reduction factor expressed as an exponential decay function based 1994; McCool, Brown, Foster, Mutchler, & Meyer, 1987; Nearing,
on a world-wide compilation of experimental field data. Imple- 1997; Renard et al., 1997, pp. 1e404). All equations allow the
menting the effect of stoniness, reductions of 20-40% of the soil assessment of the S-factor up to a slope angle of 100%, even though
erodibility were found to be particularly significant in the Medi- empirical measurements only exist up to a slope angle of 55% (Liu
terranean (Panagos, Meusburger, Ballabio, Borrelli, & Alewell, et al., 1994). According to our knowledge only one study explored
2014b). soil loss up to slope angles of 90% using rainfall experiments
Today, soil surveys together with large scale data bases such as (Tresch, 2014, p. 132).
ISRIC SoilGrids database at a 250 m spatial resolution (Hengl et al., Wischmeier and Smith (1978) defined the slope length (L) as:
2014) and remote sensing give new opportunities for mapping soil “the distance from the point of origin of the surface flow to the
erodibility in space and time (Panagos et al., 2014b; Wang, Fang, point where each slope gradient (S) decreases enough for the
Teng, & Yu, 2016). Several studies already explored the suitability beginning of deposition or when the flow comes to concentrate in a
of hyper-spectral reflectance from soil surfaces to assess the defined channel”. As such, the L factor is defined as the ratio of field
chemical and physical properties of soil (Ben-Dor & Banin, 1995; soil loss to that from a 22 m slope and the value of L may be
Luleva, van der Werff, Jetten, & van der Meer, 2011; Rossel & expressed as (l/22.1)m, where l is field slope length in meters and
Behrens, 2010; Shepherd & Walsh, 2002; Stenberg & Rossel, m is a factor that varies with slope gradient in the USLE and the
2010; Wang et al., 2016). ratio of rill to interrill erosion in the RUSLE.
It has to be noted, that the underlying pedotransfer functions to Soil loss is much less sensitive to changes in slope length than to
derive soil erodibility estimations are essentially all based on changes in slope steepness (McCool et al., 1987). Most measured
American soil-erosion databases, which raises the obvious question slope lengths are less than 120 m and generally do not exceed 300
about their applicability in other geographical locations. Conse- m (McCool, Foster, & Weesies, 1997). Nevertheless, it has to be
quently several studies aimed to determine the best suited soil stated, that little research has been conducted to assess the relation
erodibility estimation for specific scales, regions or conditions between slope length and soil loss. Moreover, the existing findings
(Hussein, Kariem, & Othman, 2007; Ro €mkens, Poesen, & Wang, are not ambiguous as with increasing slope length soil erosion was
1988; Torri, Poesen, & Borselli, 1997; Wang, Zheng, & Wang, observed to decrease (Joel, Messing, Seguel, & Casanova, 2002;
2012; Wawer & Nowocien, 2006; Zhang et al., 2004a, 2008). Kara, Şensoy, & Bolat, 2010; van de Giesen, Stomph, & de Ridder,
However, hardly any study came up with an ultimate conclusion on 2005; Xu et al., 2009; Yair & Raz-Yassif, 2004), increase (Rejman
the suitability of different erodibility estimations, because of a lack & Brodowski, 2005; Wischmeier & Smith, 1958; Zingg, 1940a) or
of measured plot data (Wang et al., 2016). As such, some new having no remarkable effect (Wischmeier & Smith, 1958).
studies, and databases incorporating measured K-values from long- The USLE was firstly developed to predict soil loss on uniform
term observations of natural runoff plots are of utmost interest. slopes and fields (Wischmeier, 1976; Wischmeier and Smith, 1965,
1978, p. 58). But already in 1974, Foster and Wischmeier (1974)
developed a method to divide an irregular slope into a number of
4.3. LS factor uniform segments and as such accounting for the effect of the
shape of the slope on soil loss opening the doors to watershed scale
In USLE models the effect of topography is usually considered by

Fig. 4. Dependency of the S factor on inclination (slope steepness) for different parametrizations.
212 C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225

USLE applications (Griffin, Beasley, Fletcher, & Foster, 1988; management measures (Wischmeier & Smith, 1965). In other
Williams & Berndt, 1977). words, it reflects the effects of biomass cover and soil-disturbing
However, these manual methods were demanding and by activities. It is expressed as the ratio of soil loss from land under
replacing plots with watersheds the slope length loses significance specified conditions to the corresponding loss from clean-tilled,
for predicting overland flow and soil loss, since in reality, surface continuous fallow (given by the product of R$K$L$S). The soil loss
water flows converge and diverge across the landscape (Desmet & ratio (SLR) is the ratio of the soil loss from the non-bare fallow
Govers, 1996). Thus, it was proposed that the unit contributing area surface and the soil loss from the bare fallow surface over a given
which is the upslope drainage area per unit of contour length period of time. Accordingly, it is inversely proportional to the soil
should substitute slope length (Desmet & Govers, 1996). The latter surface cover which intercepts raindrops and hinders surface
approach did not only account for segment length or grid cell size, runoff by slowing it down.
but also for the position of the segment or cell within the landscape As nearly all crops can be grown continuously or in rotations
and consequently the LS factor was not one-dimensional anymore with varying sequences, the soil loss ratios can vary in cropped land
(Kinnell, 2008; Zhang et al., 2013). The first physically based slope- over time as they are a function of canopy, ground cover, roughness,
length factor was developed by Moore and Burch (1986). They soil biomass and consolidation change. The C factor depends on the
concluded that the theoretical and empirical slope-length factors particular stage of vegetation and vegetation cover development at
are equivalent and thus the USLE slope-length factor is indeed a the time the rain event occurs, but also on the prior land use con-
measure of the sediment transport capacity of runoff from the ditions. For instance, plant residues can be removed or left on the
landscape. However, the USLE slope-length factor fails to fully ac- field, incorporated near the surface or ploughed under, chopped or
count for the hydrological processes that affect runoff and erosion omitted by the harvesting operation. All these different measures
(Moore & Burch, 1986). can have different effects on soil loss. Originally five crop stages
A multiple flow algorithm is implemented by MUSLE, RUSLE, were defined including rough fallow, seedling, establishment,
EPIC and G2 models (Karydas et al., 2014). Winchell, Jackson, growing and maturing crop and residue or stubble. From 10,000
Wadley, and Srinivasan (2008) further improved the method and plot years of runoff and soil-loss data assembled from 47 research
compared several variations of the GIS approach. A second algo- stations in 24 states of the US empirical tables were created taking
rithm was associated with the development of the USPED model by the crop stage, vegetation type, management measures and their
Mitasova et al. (1996). The greatest limitation of all these methods timing, and rainfall intensity EI in the specific periods into
is (i) the absence of an algorithm for predicting deposition consideration (Wischmeier & Smith, 1965). In RUSLE, the compu-
(Mitasova, Hofierka, Zlocha, & Iverson, 1997; Winchell et al., 2008) tation of the C factor follows a scheme proposed by Laflen, Foster,
and ii) the neglect of the variability of infiltration along the flow and Onstad (1985) and (Weltz, Renard, & Simanton, 1987, pp.
path. The flow path and cumulative cell length-based method (FCL) 104e111) and tables of SLR are no longer provided. Five sub-factors
proposed by Dunn and Hickey (1998) and Hickey (2000) and are multiplied together to yield soil loss ratios (prior-land-use
advanced by Van Remortel, Hamilton, and Hickey (2001) and Van (PLU), canopy-cover (CC), surface-cover (SC), surface-roughness
Remortel, Maichle, and Hickey (2004) who tried to overcome this (SR) and soil-moisture (SM)). Individual values of SLR are
disadvantage. In 2013, Zhang et al. (2013) presented an algorithm computed in time intervals within which the five sub-factors
that merges both approaches to account for flow convergence remain constant. Subsequently, the SLR values are weighted by
based on the contributing area as well as slope cutoff conditions the fraction of rainfall erosivity (EI) of the coinciding time period (i)
and thus accomplished an important step towards net erosion and combined to achieve an overall C factor value using
estimation.
To conclude, increased availability of gridded digital elevation X
n

data, often referred to as DEMs, improved the estimation of the LS


C¼ ðSLRi $ EIi Þ = EIt (5)
i¼1
factor. The availability of high resolution DEM (Europe 25m but in
certain countries much higher (e.g., Switzerland DEM 2m)) has Although it is possible to compute C values for a wide set of
outpaced method development of flow algorithms (Seibert & tillage techniques and crop rotations within RUSLE, a large number
McGlynn, 2007), which need to define the contributing area. of input data are required (Gabriels, Ghekiere, Schiettecatte, &
Currently, four algorithms exist to calculate the upslope contrib- Rottiers, 2003). With the advent of GIS-based modelling and the
uting area: 1) single-direction flow algorithm, D8 (e.g., O'Callaghan ambitions to assess the potential impact on soil erosion over larger
and Mark (1984), p. 2) the multiple flow direction algorithm (MD8, areas, the C factor, as with most other factors within RUSLE, has
Quinn, Beven, Chevallier, and Planchon (1991)), 3) infinite possible experienced a process of computation transformation and simpli-
single-direction flow pathways (D∞), and 4) triangular multiple fication. At catchment- and regional-scale, most of the input pa-
flow direction algorithm (MD∞, Seibert and McGlynn (2007)). Ac- rameters for the RUSLE C sub-factors became hard, if not impossible
cording to our knowledge comparison of the different flow algo- to assess and quantify given that the model simulation is a specific
rithm to mapped erosion features is still missing. field plot. In recent years, the methods most commonly employed
Another main research gap related to the availability of high to compute the C factor and set large-scale RUSLE-based modelling
resolution DEM is the dependence of USLE-type estimates on grid are simple attributions of literature C-factor values to land use
size of the DEM. Soil erosion estimates were found to decrease with maps without further land sub-classification (Bakker et al. (2008);
decreasing resolution (Mondal et al., 2017). Consequently erosion Ma €rker et al. (2008), among others) or satellite remote sensing
estimates may be biased depending on the resolution of the DEM. approaches ((Borrelli, Panagos, Ma €rker, Modugno, & Schütt, 2017;
To investigate resolution dependence seems even more pending in De Jong, 1994; Scho € nbrodt, Saumer, Behrens, Seeber, & Scholten,
the light of drone derived DEM with high horizontal resolutions of 2010; Van der Knijff, Jones, & Montanarella, 1999), among others).
5 cm (Peralta et al., 2017). With increasing calls for more knowledge on soil erosion on macro-
regional scale, research on C factor based on remote sensing tech-
4.4. C factor niques has become a widely applied method across the globe
(Vrieling, 2006; Zhang, Zhang, Liu, Qiao, & Hu, 2011). While these
The cover-management factor (C) in USLE-type equations applications often inadequately represent the management aspect
measures the combined effect of all interrelated cover and of the C factor, the gained insights nevertheless provide a valuable
C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225 213

contribute to improve the understanding of the soil erosion po- certain slope. Furthermore, the impact of stone walls and grass
tential of a given region, especially when considering monthly or margins was modelled using the more than 226,000 observations
sub-monthly land use/land cover change. Most recent, regional- from the Land use/cover area frame statistical survey (LUCAS)
scale C factor computation approaches have moved towards carried out in 2012 in the European Union (Panagos et al., 2015c).
spatially and spatiotemporally descriptions using vegetation based
indices for land use classifications (Meusburger, Konz, Schaub, & 5. Model validation
Alewell, 2010b; Mo € ller, Gerstmann, & Gla €ßer, 2014, pp.
5072e5075; Panagos et al., 2014c; Schmidt, Alewell, & Meusburger, 5.1. General thoughts on model validation
2018; Scho €nbrodt et al., 2010; Teng et al., 2016) or weighted average
operations combining crop statistics with remote sensing and GIS ‘Validation’ refers to the testing of the model output to confirm
modelling techniques (Borrelli, Panagos, Langhammer, Apostol, & the results that should be produced to reality (Fishman & Kiviat,
Schütt, 2016b; Panagos et al., 2015b). A geo-statistical approach 1968). Ecosystem modelers suggested that a ‘valid’ model of a
based on weighted average operations was recently introduced to biological system would necessarily be an exact copy of the system
thoroughly incorporate the extent, types, spatial distribution of (Oreskes, Shraderfrechette, & Belitz, 1994). Alewell and
global croplands, and the effects of the different regional cropping Manderscheid (1998) argued that a direct validation of a model
systems into a global RUSLE-based soil erosion model (Diodato would prove that this model is an exact description of the modeled
et al., 2017). In these recent approaches, the C factor reflects the system, the model output and the system's output matching under
relative effectiveness of the soil and crop management systems in a any circumstance. As such, model validation would only be possible
given region in terms of their ability to reduce soil loss. for thermodynamically closed systems (Oreskes et al., 1994), where
all parameters and processes are described and measureable (e.g.
4.5. P factor within a mathematical proof). Following this line of thinking, a
model describing a thermodynamically open system cannot be
As in the case of the C factor, the support practice factor (P) in validated.
the RUSLE-type equations is based on the concept of deviation from A very different definition of validation derives from the
a standard. It expresses the ratio of soil loss in a field with specific meaning of the Latin word validus (¼ strong, healthy, powerful) and
support practice to the loss of soil under conditions of straight-row was used by Martin (1996). In that sense, a model is validated, if it
farming up and down the slope. The support practices affecting the has been proven itself as a strong political or economic tool (Martin,
erosion process by modifying the amount and dynamics of runoff 1996), e.g. putting a legislation into action or setting up a man-
are considered in the P factor. For cultivated land supporting agement plan. Foster, Yoder, Weesies, and Toy (2001), in presenting
practices generally considered as P factor are contour tillage/ the RUSLE2 concept, also connects model success to its usefulness,
planting, strip-cropping, terrace farming systems and stabilized when stating that a model which provides perfect erosion esti-
water ways because improved tillage practices such as sod-based mates but which somehow leads to a poor conservation planning
rotations, fertility treatments, and greater quantities of crop resi- decision has failed to meet its objectives. Indubitable, we have left
dues left on the field are already part of the C factor (Wischmeier & the realm of natural science with these lines of thinking. However,
Smith, 1965). The overall P factor is the product of the individual following the recent critical debate of soil erosion modelling, the
support practices. In the original paper of Wischmeier and Smith definition of Martin (1996) or the logic of Foster et al. (2001) has
(1965) empirical P values to consider contouring were provided been implicitly applied. E.g., one of the main concluding arguments
but the researchers recommended to consult further expert to dismiss the Pan-European modelling of Panagos et al. (2015d)
judgement to adapt the P value for specific field conditions espe- was, that “erosion rate figures can be used by politicians/policy
cially when small gullies or rills exist or one of the above mentioned makers to advocate measures that may be misguided” or that
measures was adapted. Within RUSLE, a list of P values obtained “modelled information can be used to suggest that policies to
from experimental data supplemented by analytical experiments is mitigate erosion are successfully tackling soil loss across Europe”
provided, offering a wide range of support practice conditions. (Evans & Boardman, 2016a; Fiener & Auerswald, 2016).
Nevertheless, according to (McCool et al., 1987) the values of the P If we accept models as tools in natural science, we can use them
factor are the least reliable of the USLE/RUSLE factors as the effec- to test hypothesis on process understanding, relative differences
tiveness observed in field studies conducted on given slopes between systems or potential development over time. Unlike a
showed wide ranges of reductions. mathematical proof, a scientific theory or hypothesis is empirical,
Typical P values range from about 0.2 for reverse-slope bench and is always open to falsification if new evidence is presented. A
terraces to 1.0 where there are no erosion control practices given model which describes a specific system significantly better
(Wischmeier & Smith, 1978, p. 58). The lower the P value, the more might be declared the ‘valid’ model compared to another model
successfully the support practice or the combination thereof pro- which might be rejected and the term ‘valid’ in this sense is then
motes the deposition of soil particles. From a conservation planning used that any model that could not be proven invalid would be a
perspective, the deposition of the soil that is close to the source is valid model for the system (Alewell & Manderscheid, 1998).
preferable but the effectiveness of the support practice depends on There is general agreement in modelling that the calibration
the characteristics of the agricultural area (e.g., the slope gradient). dataset must be independent from any dataset which is used later
With the transition from field-to GIS-based modelling, the lack of to validate the model, and that if the same dataset is used for both it
spatial information to compute the P factor implies that in the vast should be no surprise that the model is a perfect predictor (Nearing,
majority of the catchment-to regional-scale GIS-based models the 2004). Split sample approaches, in which the available data is
crucial effect of support practices on soil erosion has been omitted. separated into a calibration set and a separate validation set, are
However, recent large-scale studies proposed alternative ap- usually the solution to this problem.
proaches that are able to provide relative estimations of the P factor For the open systems described in ecosystem modelling it was
and assess the possible effects of conservation policy (Panagos suggested that a model can only be validated, if it was not cali-
et al., 2015c; Yue et al., 2016). Panagos et al. (2015c) considers the brated (Sverdrup, Warfvinge, Blake, & Goulding, 1995), implying
latest policy developments in the Common Agricultural Policy, and that measured input parameters implemented in the model will be
applies the rules set by Member States for contour farming over a directly compared to measured target values without any inverse
214 C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225

process step. Interestingly, soil erosion modelling using empirical multi-year erosion totals were used. Values of the coefficient of
models, as much as it has been criticized in the past, commonly variability ranged from nearly 150% for a soil loss of 1 Mg ha1yr1
follows this most rigorous of all approaches: e.g., USLE-type model to as low as 18% or less for soil-loss values greater than 100 Mg
applications derive the input parameters from measured data and, ha1yr1 (Nearing et al., 1999).
without the process of calibration, compare modelled soil erosion As such, trying to validate USLE-type modelling with measured
maps directly with measured data. data, needs to take the high variability of soil erosion measure-
ments into account. In addition, it has been noted all throughout
5.2. Validating USLE-type models with measured data the history of using the USLE that there is a striking lack of data to
rigorously determine the single factors for all needed regions, sit-
One major obstacle in validating the (R)USLE with measured uations and scenarios (Auerswald et al., 2003; Wischmeier and
data is that (R)USLE provides gross erosion rates while most Smith, 1965, 1978, p. 58). Many plot studies measuring soil
monitoring programs and measurements (except for experimental erosion rates investigated bare fallow plots as a baseline reference
approaches) provide net erosion rates (e.g. on-site measurements which hardly ever exist in reality. While the focus of measurements
as the budget of erosion and deposition of sediments or even off- is towards medium sloped (e.g. > 3% < 18%) arable soils, flat areas or
site sediment yields in waters). Thus, and as long as USLE is not steeper slopes >20% under grassland or forest used are hardly ever
connected to sediment transport models (which is mostly not investigated (Bonta & Sutton, 1983; Borst & McCall, 1945; Fan, 1987;
operational on large scales, see section 3), modelled and observed Kilinc & Richardson, 1973; Liu et al., 1994; Tresch, 2014, p. 132;
data represent different fluxes. Or, as (Trimble & Crosson, 2000) Zhang, Wang, & Yang, 2014). Furthermore, measurement plots
stated, USLE only presumes to predict the amount of soil moved on were either located in areas with relatively large rainfall erosivity or
a field (e.g., gross erosion), not necessarily the amount of soil experimental treatment with high irrigation rates and intensities
moved from a field (e.g., net erosion). As such, a RUSLE modelling was used. So, all in all, there is bias in the available measurement
was not successfully validated when compared to sediment yields data towards regions or situations with conditions prone to erosion
of 454 small catchments with various land covers and uses across (Auerswald et al., 2003; Wischmeier and Smith, 1965, 1978, p. 58).
the United States, Puerto Rico, U.S. Virign Island and Guam (Tan Nevertheless, ultimately we need a validation of soil erosion
et al., 2018). Instead, models that use the power function of modelling to be able to attribute the relative importance of sedi-
runoff, shear stress or stream power such as the Morgan, the tRIBS- ment source strength to specific kinds of land use and management
Erosion and the GUEST model were more successful to target as well as other triggering factors (e.g. climatic events, topography,
catchment sediment yields (Tan et al., 2018). and vegetation cover (Alewell, Birkholz, Meusburger, Wildhaber, &
Nearing (2004) state that observations will always be closer to Mabit, 2016; Meusburger, Banninger, & Alewell, 2010a;
the truth than modelling and must remain the most important Meusburger et al., 2010b; Panagos et al., 2015d; Scheurer, Alewell,
component of scientific investigation. However, Garcia-Ruiz et al. Banninger, & Burkhardt-Holm, 2009; Schindler Wildhaber,
(2015) observed a correlation of measured erosion rates with size Michel, Burkhardt-Holm, Baenninger, & Alewell, 2012). As such,
of the study area, measurement method and duration of experi- erosion models have been compared extensively to measured data
ment or observation period in a worldwide meta-analysis of 4000 sets.
sites. Furthermore, there are many regions of the world being un- Risse et al. (1993) applied the USLE to 1700 plot years of data
derrepresented and erosion measurements lack long term obser- from 208 natural runoff plots. Annual values of measured soil loss
vations (Garcia-Ruiz et al., 2015). Thus, it has to be taken into averaged 35.1 Mg ha1 with an average magnitude of prediction
account that observations, especially due to the highly heteroge- error of 21.3 Mg ha1, or approximately 60% of the mean (Risse
neous nature of soil erosion and the tendency of erosion mea- et al., 1993). Rapp (1994, pp. 1e95) followed the latter approach
surements being very prone to errors (Garcia-Ruiz et al., 2015), and determined the error associated with RUSLE from 21 US sites
might also be biased in regional distribution or time periods chosen representing 1704 years of measurements from 206 plots. The
by the researchers or due to high systematic errors or random average annual magnitude of error was 11.7 Mg ha1 and the model
uncertainty (Rüttimann, Schaub, Prasuhn, and Rüegg (1995), see efficiency coefficient was 0.73, while the prediction of the 1638
section 5.2). As such, quality and potential biases of modelled and individual annual erosion values had an average magnitude of error
measured data has to be equally scrutinized, when model valida- of 20.8 Mg ha1 and a model efficiency of 0.58 (Rapp, 1994, pp.
tion with measured data is aimed at. 1e95). Zhang, Nearing, Risse, and McGregor (1996b) applied the
One reason of high modelling uncertainty frequently addressed Water Erosion Prediction Project (WEPP) computer simulation
in the literature is certainly that erosion measurements itself are model to 290 annual values and obtained an average of 21.8 Mg
connected to considerable uncertainty. Rüttimann et al. (1995) re- ha1for the measured soil loss, with an average magnitude of
ported a statistical analysis of data from four sites, each with five to prediction error of 13.4 Mg ha1, or approximately 61% of the mean.
six reported treatments. Each treatment had three replications. In all cases the relative errors tended to be greater for the lower
Reported coefficients of variation of soil loss within treatments soil-loss values with an over prediction on plots with low erosion
ranged from 3.4% to 173.2%, with the ploughed sites with higher soil rates and under prediction on plots with high erosion rates. Even
losses having a tendency to lower coefficient of variation between though deviations from modelled and measured data might seem
3.4 and 71.4%. Work carried out by Wendt, Alberts, and Hjelmfelt substantial in absolute numbers, the deviation between modelled
(1986) using 40 erosion plots, all with the same treatment in and observed erosion rates is not larger than the variability within
Minnesota found that coefficients of variation for the 25 storms measured data.
ranged from 18% to 91%. The larger erosion events again showed Kinnell (2010) in evaluating data from Rosewell (1993) also
less variability. Very little of the variability could be attributed to concludes that even though the RUSLE works well for a number of
measured plot properties, and the plots did not perform in the agricultural conditions in New South Wales, Australia, on a wider
same manner relative to each other in subsequent events (Wendt geographic area and larger variety of agricultural systems, there is a
et al., 1986). Nearing, Govers, and Norton (1999) studied erosion tendency for the USLE and the RUSLE to over-predict low average
variability using data from replicated soil-loss plots from the Uni- annual soil losses and to under-predict high average annual soil
versal Soil Loss Equation (USLE) database. Data from replicated plot losses. Zhang, Nearing, Risse, and McGregor (1996c) also discussed
pairs for 2061 storms, 797 annual erosion measurements and 53 an over prediction of small soil losses as well an under-prediction of
C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225 215

large soil losses for USLE approaches as well as the WEPP model. ha1yr1 (Meusburger et al., 2013). Thus, even though the FRN
On-site erosion measurements are not available on large scale, method also determines net erosion rates while USLE-type
but Cerdan et al. (2010) compiled an extensive database from modelling results in gross erosion estimates, the promising re-
literature on short to medium-term erosion rates as measured on sults of these studies might corroborate the assumptions that
erosion plots in Europe under natural rainfall conditions. Panagos comparisons may be applicable over long time periods of several
et al. (2014c) compared the soil erosion estimates of this decades which are evaluated by the FRN methods. However,
compiled plot data base (Cerdan et al., 2010) with modelled values Haciyakupoglu, Hizal, Gokbulak, and Kiziltas (2010) compared 137Cs
from the Pan-European Soil Erosion Risk Assessment (PESERA) based erosion rates to USLE assessment in the Turkish Omerli
model (Kirkby et al., 2008) and the European Environment Infor- catchment and found substantial deviations between the two
mation and Observation Network for soil (EIONET-SOIL) database of methods (higher ULSE rates by a factor of 2-3). The latter over-
Europe (which are regional to national USLE-type model applica- estimation of soil losses by USLE compared to 137Cs based erosion
tions). The mean values of soil loss reported by the national in- rates was also noted by Warren et al. (2005), but was overcome
stitutes (EIONET-SOIL) are larger than the PESERA estimates, with when USPED was applied, which considers net erosion rates (see
the main differences being for sloping land (>2 ) and for the land section 3) and might thus be better comparable to 137Cs based
cover type forest and heterogeneous agricultural land cover erosion rates.
(Panagos et al., 2014c). A graphical illustration of this comparison Regardless of the latter promising FRN results the question may
results in a deviation of both modelling approaches from the plot be asked or, nevertheless, has to be asked: are the model pre-
based measurements at the high end of modelled soil losses (Fig. 5). dictions ‘good enough’? And what is actually ’good enough‘ or what
Cerdan et al. (2010) warned of extrapolations from local mea- are the reference values considering the high uncertainty and
surements to regions and in addition the comparison might illus- biases of measured data? Even though Bagarello et al. (2012) con-
trate the above discussed principally different nature of net erosion cludes that soil loss estimates are considered reasonably accurate
rates based on plot measurements versus larger scale modelled for most practical purposes, especially for wide spatial scale ap-
gross rates. Furthermore, there is no straightforward conclusion plications, when the errors of the predictions do not exceed a factor
from these comparisons of modelled and measured data: are the of two or three, erosion modelling remains assailable as long as
models invalidated by the measured data or is the measured data there is no general agreement on the ’good enough’.
biased and/or not reflecting large spatial scale and long temporal Nearing (2004) concluded that model validation is not just a
average situations? matter of comparing measured to modelled data, one must also ask
As the direct measurement of soil erosion on plots implies long the question: “How variable is nature?” We would like to add, that
term monitoring all year round, the indirect assessment using in bidding farewell to the idea of accurately predicting absolute
fallout radionuclides (FRN) as tracers of soil erosion has been used. values with models but rather concentrating on the prediction of
Meusburger et al. (2010b) with an adapted C factor implemented relative differences, trends over times and systems reactions to
from fractional vegetation cover map derived from QuickBird im- processes and management practices, we can use models as tools to
agery, validated USLE derived erosion rates for hot spots of erosion learn about the modelled systems and their reactions. In this con-
in Alpine grasslands against FRN derived erosion rates (20.1 Mg ceptual approach, modelling in general and large-scale modelling
ha1yr1 versus 16 Mg ha1yr1 for 137Cs based versus modelled specifically will per se not aim at an accurate prediction of point
rates, respectively). RUSLE was also successfully validated at sites of measurements, but will test hypotheses on process understanding,
South Korea, with the 137Cs method ranging from 0.9 Mg ha1yr1 relative spatial and temporal variations, scenario development and
to 7 Mg ha1yr1 and RUSLE (considering the steep slopes of up to controlling factors (Oreskes et al., 1994). In an agreement with the
40 and the erosive monsoon events (R factor of 6600 latter and the above discussion, many modelers accept that full
MJmmha1 h1 yr1)) yielding 0.02 Mg ha1yr1 to 5.1 Mg model validation is a logical impossibility and Morton and Su'arez
(2001), Refsgaard and Storm (1996) as well as Senarath, Ogden,
Downer, and Sharif (2000) suggest that in most practical contexts
the term ‘model’ should be thought of as being synonymous with
‘theory’ or ‘hypothesis’, with the added implication that they are
being confronted and evaluated with data. An interesting example
of such ’hypothesis testing‘ is a study in northern Bavaria where a
comparison of conventional with organic farming practices found
measured soil losses to be about an order of magnitude lower than
RUSLE based values due to best management practices adopted on
both farms (Auerswald et al., 2003). However, modelled and
measured values pointed to the same conclusion of organic agri-
culture resulting in 15% less erosion on arable land despite its
localization in areas with substantially higher erosion susceptibility
(Auerswald et al., 2003). Garcia-Ruiz et al. (2015) concluded from a
global meta-analysis that “Geomorphologists should be more
interested in processes, causes, internal relationships, and the role
of stream channels than in obtaining absolute erosion rates, the
interpretation of which can lead to numerous errors”. However, as
USLE-type modelling is often used by land scape managers and
stake holders, it is important to communicate that these modelling
endeavors should not be used to predict absolute erosion rates (e.g.
for comparison to legal limit values) but rather to optimize catch-
Fig. 5. Erosion estimates from plot based measurements (Cerdan et al., 2010) versus
ment management and land use in delineating best practice
modelled erosion rates from PESERA (Kirkby et al., 2008), USLE-type models (EIONET techniques.
Data Base; (Panagos et al., 2014c)) and RUSLE2015 (Panagos et al., 2015d). With the acceptance that at least at large scales we should not
216 C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225

be solely focused on an accurate prediction of absolute values but parameters (e.g. hydrological as well as vegetation parameters) did
rather see models as a tool to test hypotheses of relative differences not invoke a sensitive response from the model. A similar pattern
between systems, trends over time, systems reactions to driving can be observed for WEPP in the work of Brazier, Beven, Freer, and
factors and management practices, we should, however, not Rowan (2000). Quinton (2013), in discussing the uncertainty of
generally exclude the use of models to assess soil erosion magni- complex erosion models like EUROSEM and WEPP, suggests that
tudes at least at smaller scales with the aim to assess sustainability this uncertainty is reason enough to describe the erosion system
of management. An example would be a study by Alewell, Egli, and with simpler models than those that are currently in vogue. One
Meusburger (2015) where 137Cs as well as 239þ240Pu derived soil reason for the popularity of USLE-type modelling is certainly that it
erosion estimates were compared with USLE estimates from is a good compromise between applicability in terms of required
Meusburger et al. (2010b) and confronted with soil formation rates input data and relatively good reliability of obtainable soil loss es-
in Swiss Alpine grasslands at plot scale. The conclusion from the timates (Risse et al., 1993). As such, it has become the standard
latter was, that measured and modelled soil erosion rates compare technique of many soil conservation workers (Morgan, 2005, pp.
relatively well but that they are in any case considerably higher 365e376). However, with remote sensing techniques improving
than soil formation rates and that thus land management cannot the availability of high resolution data immensely in recent years
claim to be sustainable. (dynamic monitoring of vegetation development as well as esti-
mating dynamic of climate variables), the way is paved for process
6. Model comparisons based techniques in the future needing high resolution spatial
input data.
6.1. Model choice and suitability The choice of a soil erosion-prediction tool has to be dependent
on the spatial and temporal scale of the intended model applica-
A complete overview in a recent review on water erosion tion, as the question of scale is crucial in choosing the right
models identified 82 models (Karydas et al., 2014). As stated above modelling approach. Process-oriented models require an applica-
an Web of Science query (http://apps.isiknowledge.com/) for the tion of the used equations at a given spatial scale, ranging from plot
period 2007e2017 resulted in 1149 hits corresponding to the key- to basin, and at event temporal scale. Event and at-point or small
words “Universal Soil Loss Equation”, “USLE”, “Revised Universal scale (hillslope) process based models are not suitable to be applied
Soil Loss Equation” or “RUSLE”. In comparison, most popular ap- for simulating soil loss on a wide region not only because of the lack
proaches independent of the USLE technology, such as WEPP of required input data but also because of the required indepen-
(Laflen et al., 1997), LISEM (De Roo et al., 1996), EUROSEM (Morgan dency of spatially discrete units. At a large spatial scale, the area has
et al., 1998) and PESERA (Kirkby et al., 2008) totalized 254 hits. to be discretized using, for example, a square grid subdivision
Needless to say, that the fact that USLE-type models are the most (raster scheme), choosing a mesh size consistent with the scale of
commonly used erosion tools globally, does not rank them auto- the original model deduction. Using a raster scheme applied to the
matically as the best available technique and the suitability of USLE model corresponds to hypothesize that each cell is indepen-
model concept needs to be proven for each specific case. dent of the others with respect to soil loss. Even though this meets
In the choice of the soil erosion model to be applied, and the original requirements postulated by Wischmeier and Smith
congruent with the discussion of section 5, it has to be taken into (1965) that for larger scale applications the area needs to be
account that it is not always necessary to calculate the exact erosion “broken down into a series of tracts having relatively homogeneous
rate for a particular situation (Nearing, 2013, pp. 365e378), but land use and treatment.”, it requires independency of each of these
rather to compare among different situations or time periods. units. In other words, at large spatial scales such as a region or even
Choosing which model to apply and model development might a global perspective, a simple index-based model able to calculate
become a matter of what type of information is available as input an average soil loss at annual or mean annual scale, allows
data. As an example, Nearing (2013, pp. 365e378) noted that most computing the involved factors only using spatially distributed
applications of WEPP were developed in the United States because input values. However, the cells cannot be assumed to be inde-
of the availability of soil, climate and crop information for those pendent of each other when sediment delivery processes have to be
areas. Another important criterion for the choice of a soil erosion modeled (e.g., at basin scale). In this case, the USLE scheme has
model is the question of spatial and temporal scale of available been applied by coupling it with a mathematical operator
input parameters as well as envisaged model output. As most expressing the hillslope transport efficiency (Ferro, 1997; Ferro &
erosion parameters are scale-dependent, data collected on plot Porto, 2000).
scale are not always appropriate for calibrating or validating
models (see also section 5) and upscaling of models through data 6.2. Direct model comparison used implicitly as model validation
aggregation may cause biases (Karydas et al., 2014).
In today's modelling world, many of the current generation of Model comparison has been used implicitly as a way of model
erosion models are getting more complex, with ever more pro- validation. However, for a rigorous model comparison the question
cesses being incorporated into them (Quinton, 2013). However, of scale is decisive (see also section 7). In a European application of
even the most complex model of erosion and sedimentation pro- the models PESERA and SENSOR both models compared well,
cesses can still only be an approximate representation of reality and however when ‘validated’ to the national scale USLE derived
not reality itself. The problems with complex models are twofold: erosion map of Hungary, the European applications of SENSOR and
(i) unavailability of the complex set of input parameters especially PESERA underestimated (or USLE overestimated) erosion rates
at large temporal and spatial scales and (ii) over parameterization considerably (Fig. 6a; (Podmanicky et al., 2011)). The original
which results in a non-uniqueness of fit and/or insensitivity of thought of this comparison was, to validate the European applica-
model output to some of the input parameters. From a mathe- tions of SENSOR and PESERA with the higher resolution soil erosion
matical view point the calibration of ecosystem models in general rates from Hungary derived with USLE (Podmanicky et al., 2011).
are hardly ever unique and the non-uniqueness of fit increases with However, if model type and scale are changed simultaneously, it
increasing numbers of parameters (Alewell & Manderscheid, 1998). remains unsolved, whether the deviation originates from scaling
Quinton (2013) as well as Veihe and Quinton (2000) and Veihe, inconsistencies or from model deviations or even failures.
Quinton, and Poesen (2000) showed that many of EUROSEM's input A comparison of national scale USLE-type modelling (EIONET
C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225 217

Fig. 7. Erosion estimates at the Mediterranean climate of Chile from plot based mea-
surements versus modelled rates from USLE, RUSLE, EPIC and WEPP (all data from
Stolpe (2005) with two different parameterizations of the WEPP model).

to be not better than either the USLE or the RUSLE (Kinnell, 2010).
USLE-type application has been criticized in complex topo-
graphic terrain, because the LS factor does not consider terrain
shape of the upslope area (Desmet & Govers, 1996; Mitasova et al.,
1996). Mitasova et al. (1996) compared the USLE LS factor and the
topographic erosion/deposition index incorporated in the so called
unit stream power based approach USPED and concluded, that the
latter is more appropriate for landscape erosion modelling, espe-
cially when erosion risk and deposition potential needs to be
evaluated.
Using more than 1600 plot years of data from natural
Fig. 6. Model comparison of soil loss data (Mg ha1yr1) calculated with SENSOR and rainfallerunoff plots in the USA, Tiwari, Risse, and Nearing (2000)
PESERA calibrated to European scale with USLE estimates applied to the national data observed that for average annual soil losses, the model efficiency
of Hungary (data of Podmanicky et al. (2011), Fig. 6a) and of national USLE-type of WEPP determined using the (Nash & Sutcliffe, 1970) factor was
modelling (EIONET data base, Panagos et al. (2014c)) to PESERA (Kirkby et al., 2008)
0.71, whereas it was 0.80 and 0.72 for the USLE and the RUSLE,
and RUSLE2015 (Panagos et al., 2015d), Fig. 6b).
respectively. However, as the WEPP model derives its strength from
being a process based model and has the capability to predict
data base, Panagos et al. (2014c)) with the European scale appli- spatial and temporal distribution of soil loss and deposition, it has
cation of RUSLE2015 (Panagos et al., 2015d) or PESERA (Kirkby et al., to be considered in the latter study that the data set was biased
2008) resulted in a relatively good agreement for low values but towards the USLE (all uniform natural runoff plots), the parameters
high deviations of the models at the higher end of soil losses used in the USLE have undergone more refinement, and the WEPP
(Fig. 6b). model was not calibrated at all (Tiwari et al., 2000). WEPP, USLE and
Stolpe (2005) compared RUSLE, WEPP and EPIC at the Medi- RUSLE were observed to over-predict low and under-predict high
terranean climate of Chile for their consistency with measured annual soil losses by Kinnell (2010). However, the latter was not
erosion rates (2*20 m plots) and those calculated using the previ- confirmed for the above presented data from Europe (Fig. 7).
ously calibrated USLE. Their conclusion was that WEPP permitted Favis-Mortlock (1997) drew together the results from models
the most adequate parameterization for both climate and volcanic simulating annual soil loss for a field site in the USA. The models
soils and that WEPP erosion rates were in best agreement with ranged from simple models, such as CSEP (Kirkby and Cox, 1995)
measured rates as well as with calibrated USLE values (Fig. 7). and those based on the USLE (Wischmeier & Smith, 1978, p. 58)
Nearing (2004) in comparing USLE and WEPP in two case studies, such as EPIC (Williams et al., 1983) and GLEAMS (Leonard, Knisel, &
concluded that USLE can provide certain information which WEPP Still, 1987), to the more process based and complex WEPP model.
simply cannot provide because of the restrictions of model Favis-Mortlock's results show no significant difference between the
complexity, but that the WEPP model can be used in a way where performances of the models. Morgan and Nearing (2011) compared
only the complex model interactions will provide the information the performance of the USLE, RUSLE and WEPP using 1700 plot
needed regarding system response. years of data from 208 natural runoff plots. Once again it became
Kinnell (2010) criticizes that WEPP ignores any variation in the clear that the more complex model, in this case WEPP, performs no
kinetic energy per unit quantity of rain, even though it is dependent better than the empirical, five parameter USLE (Morgan & Nearing,
on rainfall intensity data. These intensities are generated by the 2011).
stochastic daily parameter weather generator CLIGEN (Nicks, Lane, Using the USLE basically comes down to testing the hypothesis,
& Gander, 1995), but the meteorological data required to parame- that only a small number of parameters control the response of an
terize CLIGEN are not available in many countries outside the USA. erosion model and the remainder become largely inconsequential
The ability of WEPP to predict soil losses in the USA has been found (Quinton, 2013). As such, model uncertainty does not seem to be
218 C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225

smaller in complex, process based models compared to simpler, as a graphical summary in maps. With a spatially discrete model-
empirical models. Quinton (2013) when discussing several erosion ling approach as the USLE-type algorithm uncertainty is per se in-
model studies increasing the complexity of models and comparing dependent from scale, because the uncertainty within each single
it to the original simpler version, concludes that from a scientific discrete pixel is independent from the number of these pixels. The
point of view, the inclusion of more processes and parameters uncertainty is merely coupled to the resolution of the modelling
make the models a better representation of reality, but it did not approach (e.g. size of pixels) and the connected heterogeneity
make the models better in predicting erosion rates. However, within these pixels. However, limitation of the modelling capacity
stating this, we have to keep in mind that as noted above, USLE- might be connected to scale, as demands on input data and
type modelling will always be less successful when models are computational capacity is of course increasing with increase in
compared to catchment sediment yields or off-site measurements resolution and scale.
in freshwaters (see discussion in section 5.2 and as good example Again, as discussed above, when aiming at large scales and
(Tan et al., 2018)). considering erosion risk of (large) catchments, nations, continents
To conclude, there is not one ”superior” model suitable for all or even global scales, often an misunderstanding evolves during
kind of situations, temporal and spatial scales as well as data res- this transfer from small to large scale and USLE-type modelling
olution. This basically leaves soil erosion modellers with two pos- output is implicitly compared to sediment yield and sediment
sibilities: (i) accept that there are no adequate tools to model soil transport capacities. As such, evaluating large-scale model appli-
erosion by water and give up modelling until modelling and the cation of USLE-type modelling needs the repeated reminder that
required input data sets are improved or (ii) recollect what was on-site erosion risk is simulated and not off-site sediment yield and
discussed in section 5: At least at large scales we should bid fare- transport.
well to the idea of accurately predicting absolute values with While modelling on catchment scale nowadays operates on a
models but should rather aim at the prediction of relative differ- resolution of 2 m (depending on DEM resolution, see section 4), the
ences, trends over times and systems reactions to driving factors latest large scale modelling assessments achieved a spatial reso-
and management practices and use models as tools to learn about lution of 100 m for European (Panagos et al., 2015d) and 250 m for
the modelled systems and their reactions. In this conceptual global scale (Borrelli et al., 2017; Diodato et al., 2017). The latter
approach, modelling in general and large-scale modelling specif- approaches use considerably coarser resolution than the original
ically will be understood as a tool to test hypotheses on process USLE-type plot of 22.1 m length. However, variability does not
understanding, relative spatial and temporal variations, scenario necessarily increase with increase in scale. Even though it might be
development and controlling factors (Oreskes et al., 1994). expected that measured soil loss variability should decrease with
increasing plot size, Rüttimann et al. (1995) could not confirm an
7. Upscaling: from plot to global and from event to long term effect of plot size on increase in variability but rather noticed a
averages striking absence of any effect of an increase in plot size on mea-
surement variability or uncertainty.
Application of environmental modelling always involves four So one question to be answered is, how homogenous the model
different scales: the geographic scale of a research area, the tem- units (‘tracts’ or better pixels) are and how valid the approach is in
poral scale related to the time period of research, the measurement terms of assuming homogenous R, K, LS, C and P factors within each
scale of parameters (input data resolution) and the model scale pixel. As spatial data availability is increasing continuously in res-
referring to both temporal and spatial scales when a model was olution and remote sensing tools shedding more and more light on
established (Zhang, Drage, & Wainwright, 2004b). small scale heterogeneity, the above question might eventually
The crucial importance of temporal scale becomes obvious dissipate in the future. In this, there is no principle difference be-
when considering that over and under-estimation of modelled tween a global scale modelling from a national scale modelling and
erosion rates might be connected merely to the length in obser- not even from large catchments: it always comes down to pixel size
vational data: when the observation period is too short to catch the and data quality versus the system's heterogeneity.
long-term average, observations either miss out on extreme events, Kinnell (2010) warns at applying the USLE/RUSLE model to
or might by chance catch an accumulation of extreme events. The extrapolate data collected at the plot scale to areas much larger area
USLE, designed to predict long-term average, will then of course than the field sized areas for which it was originally designed. He
overestimate or underestimate these short measurement periods, argues that at the hillslope scale, spatial variability in soil and
respectively. vegetation result in spatial variations in runoff, so that the
Regarding the spatial scale, the original USLE model and its modelling approach would need to consider those spatial varia-
factors were developed for plot scale of 22.1 m length and thus its tions in runoff, which models like WEPP and EUROSEM have an
validity for large scale application has been discussed in recent advantage over the USLE, RUSLE and RUSLE2 because they include
debates (Auerswald et al., 2015; Evans, 2013; Evans and Boardman, explicit consideration of runoff (Kinnell, 2010). Again, the implicit
2016a, 2016b; Fiener & Auerswald, 2016; Panagos et al., 2016a, transfer from on-site soil erosion risk to off-site sediment yields
2016b; Panagos et al., 2015e; Schwertmann et al., 1987, pp. 1e64). might have slipped into the discussion. However, the problem with
Schwertmann et al. (1987, pp. 1e64) when transferring the model increasing variability in vegetation cover at hill slope (or even
from the US to sites in Bavaria (Southern Germany) warned that the larger) scales might at least partially be overcome by advanced
model approach is limited to plot scale and other equations should approaches of considering high resolution spatial input data from
be used for catchments scale applications. However, Wischmeier remote sensing. One example would be a study of Meusburger et al.
and Smith (1965) already stated there is principally no logical (2010b) who implemented spatial patterns of fractional vegetation
reason why it should not be correct to “break down the drainage cover from Quickbird imagery into the C factor map and achieved
area into a series of tracts having relatively homogeneous land use (i) a good agreement of spatial patterns of soil erosion predicted
and treatment. The erosion equation is then used to approximate with USLE compared to observed remote sensing patterns and (ii)
the average annual rate of soil movement from each tract.” Large erosion rates which are at least in the same order of magnitude
scale modelling using GIS tools is basically doing nothing else, compared to 137Cs derived erosion rates (for hot spots of erosion
when approximating average annual rates of soil movement for 20.1 Mg ha1yr1 versus 16 Mg ha1yr1 for 137Cs based and USLE
single tracts (¼pixels) presenting the result of each of these pixels derived soil losses, respectively). The latter was neither possible
C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225 219

with an USLE application with uniform C factor rates not consid- climate and rain patterns, vegetation cover and management, as
ering the spatial heterogeneous distribution of vegetation cover well as the specific soil characteristics, US type systems and soils
with PESERA where the same spatially distributed vegetation cover are not a case sui generis and soil erosion from US soils is regulated
was considered but spatial soil loss patterns could not be captured by the same factors and processes as everywhere else in the world.
adequately and soil losses on hot spots were considerably under- Long term investigations of Schwertmann et al. (1987, pp. 1e64)
estimated (2.9 Mg ha1yr1 compared to the above mentioned already supported the latter assumption three decades ago.
137
Cs based estimates (Meusburger et al., 2010b)). Schmidt, Alewell, Recent studies confirmed that when appropriately parametrized,
and Meusburger (2019) followed the latter approach and combined uncertainty of USLE-type modelling all around the globe is not
the spatio-temporal heterogeneity from the C factor map with the bigger than within the US (Kinnell, 2010; Meusburger et al., 2010b,
spatio-temporal heterogeneity of the R factor to a soil erosion risk 2013; Stolpe, 2005; Yue et al., 2016).
map of Switzerland with monthly temporal and 100 m spatial Apart from the discussion whether or not modelled soil erosion
resolution. estimates are actually ‘good enough’ from a scientific perspective
and the question what actually ‘good enough’ exactly is, we would
8. Limitations of the USLE concept and the practical need to like to raise another issue: the practical need of soil erosion
model soil erosion modelling.
Without modelling there will be no large scale erosion assess-
Recent studies continue to question the use of models to predict ments. Without large scale erosion assessments, soil erosion as one
erosion potential, and argue that instead there is a need for more of the most serious threats to soils will not be taken into consid-
field-based assessment and monitoring of water erosion (Evans, eration by any major environmental and agricultural policy pro-
2013). Modelling cannot be an alternative to measurement and grams, as well as integrated land surface models (LSMs). For
monitoring but might be a powerful tool in understanding obser- example, the voluntary Guidelines for Sustainable Soil Manage-
vations and in developing and testing theories. Understanding ment of Food and Agriculture Organization (FAO, 2017) identify soil
processes, regulating parameters, trends and differences between erosion by water and wind as the most significant threat to global
spatially or temporally differing data is very often the ultimate goal soils and the ecosystem services they provide. At global scale, the
in science and models serve as tools towards that end. As such, Sustainable Development Goals (SDGs) include a sub-indicator on
models can be understood as a tool for understanding, as well as estimated soil erosion by water. Europe's Common Agricultural
tools for simulation and prediction, a virtual laboratory which Policy sets requirements to protect utilized agricultural areas
might as a positive aspect serve as an integrator within and be- against erosion, which established a framework of standards that
tween disciplines, as it brings together data, observations and aim, among others, to preventing soil erosion. The Good Agricul-
knowledge of different fields (Nearing, 2004). However, as models tural and Environmental conditions (GAEC) implemented in the
are also used as means of communicating science and the results of European Common Agricultural Policy reform (Borrelli et al., 2016c)
science, there is always the danger, that what they communicate is introduced the prevention of soil erosion and higher residues
weak or even wrong and these weaknesses might not be apparent restitution, increased soil cover and lowered tillage disturbance
to the model user, or the user of the model output (Nearing, 2004). which are among the most effective mitigation practices in Euro-
Trimble and Crosson (2000) criticized the lack of prediction of pean arable lands. All these policy programs can not consider soil
sediment delivery ratios by USLE and its failure to predict gully erosion if there are no tools for large scale modelling and mapping
erosion, which led them to conclude for the US that “the limitations available.
of the USLE (…) are such that we do not seem to have a truly Tolerable soil erosion rates all over the world are given between
informed idea of how much soil erosion is occurring in this country, 0.1 and 1 mm yr1 (for overview of all studies see Li, Du, Wu, and
let alone of the processes of sediment movement and deposition.” Liu (2009)). For sake of simplicity, these mm rates might be con-
As discussed above, USLE-type models were not designed to predict verted with a bulk density of 1 t m3 which is typical for the loose
gully erosion nor sediment delivery ratios and modelers, model soil structure of surface soils and would result in soil loss rates of
users and stakeholders should of course be aware of underlying 1e10 Mg ha1yr1. For conditions prevalent in Europe, the upper
model concepts and parameters, the above discussed limitations limit of soil formation has been estimated to be approximately
and the possible evaluation and interpretation of the model output. 1.4 Mg ha1 yr1 while the lower limit is given as 0.3 Mg ha1 yr1
Regarding the latter, the simplicity of the USLE-concept might be (Alewell et al., 2015; Verheijen, Jones, Rickson, & Smith, 2009). As
regarded as an advantage as each of the five USLE parameters might tolerable soil erosion rates should not exceed soil formation rates,
be evaluated separately even by non-expert stakeholders many countries in the world are actually performing non-
increasing thus transparency and objectiveness of evaluation. sustainable agriculture. Without large-scale soil erosion mapping
The lack of quantifying gully erosion or stream bank erosion has we do not have tools to address this problem and thus no possi-
been frequently addressed as a lack or even general failure of USLE- bilities to demand mitigation actions.
type modelling (Belyaev, Wallbrink, Golosov, Murray, & Sidorchuk,
2005; Evans and Boardman, 2016a, 2016b; Quinton, 2013; Trimble 9. Conclusions and outlook
& Crosson, 2000). Gully erosion and stream bank erosion involve
complex and highly heterogeneous hydrological and soil erosion Soil loss occurs not because of any lack of knowledge on how to
processes of channel formations difficult to model. We would like protect soils, but a lack in policy governance (Montanarella, 2015;
to state that the USLE was never meant to address gully, wind or Panagos et al., 2016b). The lack in policy governance is due to a lack
stream bank erosion, as is clearly stated in the very first publica- in research, knowledge and dissemination on spatial distribution of
tions of USLE (Wischmeier and Smith, 1965, 1978, p. 58). It would be soil erosion and sediment source attribution from water and wind
like discrediting a butterfly monitoring tool for missing out on erosion. As such, soil erosion modelling is crucially needed on both
important bird species. small and large spatial scales for planning, management and policy
USLE-type modelling is frequently discussed to be developed for measures. The USLE-type models have been extensively used and
US type soils and systems and thus questioned to be applicable for modified during the last decades reaching the ceiling of improve-
other regions (see above). However, as long as model parameters ment in recent years. Recent advancements in remote sensing,
are carefully considered and adapted to site or region specific release of climatic datasets, greater availability of earth
220 C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225

observations data, and increased processing of big datasets are the the exception of the pan-European application of the PESERA
promising factors for more dynamic and process-based modelling model (Kirkby et al., 2008).
approaches in the near future. Soil erosion modeling moves to- As measured soil loss data itself can be object to high un-
wards more dynamic approaches and can use the field parcels as certainties, future research should target reducing uncertainties in
objects due to availability of highly accurate spatial and climatic modelling and measurement. Comprehensive and well planned
data for monitoring intra-annual variability of climate, vegetation measurement campaigns at small as well as large scale are crucially
and management practices (Borrelli, Meusburger, Ballabio, needed for soil erosion modelling validation as well as deepening
Panagos, Alewell, 2018). our process understanding. As such, it is recommended to invest on
However, improving interpretation of the soil erosion processes monitoring campaigns and experimental designs towards more
at different spatial and temporal scales and testing or developing dynamic modelling. In addition, comparison of large scale model-
physically oriented modeling approaches is not only of socio- ling with local or regional studies might help to narrow possible
economic and political importance but has also an obvious scien- pitfalls, inconsistencies and uncertainty of modelling. A possible
tific aim of process understanding and general prediction of future model validation approach might be the comparison of
ecosystem dynamics (as soil erosion is an important process of continental scale modeling with local modelling data sets, which
ecosystem stability and dynamic, e.g., erosion rates in climate might give important insights into processes, trends and driving
change modelling of carbon dynamics, soil nutrient balance or factors of soil erosion. In the upcoming Land Use/Cover Area frame
pollution dissemination prediction). statistical Survey Soil (LUCAS Soil) data collection campaign, sur-
From our literature evaluation, we conclude that at today's state veyors will make a qualitative assessment of soil erosion in >26,000
of knowledge USLE-type models have no higher uncertainty than locations in Europe (Orgiazzi, Ballabio, Panagos, Jones, &
more complex based models provided modelling targets at on-site Fernandez-Ugalde, 2018). As LUCAS is repeated every 3 years, this
soil erosion risk (compared to off-site sediment transport and has the potential to evolve into a monitoring program, which can
yield). Simultaneously, they can be a choice with regard to data eventually be used to validate modeling on European scale.
availability concerns and/or ease of usage and interpretation. This Validation of the USLE-type applications are usually inexplicitly
has to be considered with care in case of increased data input carried out in the most rigorous of possible validation procedures as
availability in the near future which will allow modelling erosional the model algorithms are per se not calibrated, but the final out-
process in a more dynamic and potentially process based way. comes are compared to measured data.
Obtaining accurate and reliable soil loss estimates using Despite the large number of soil erosion models available for the
spatially distributed models on wide regions depends on both the prediction of both runoff and soil loss at a variety of scales, little
resolution (vertical and horizontal) of the input topographic in- quantification is made of uncertainty and error associated with
formation and the quality of the land-use input data. USLE-type model output. The latter seems crucially important to not only
models are a compromise for wide region application when both advance todays understanding but also the acceptance of soil
a mesh-size comparable with the original developed slope-length erosion modelling outputs on larger scales. It might be suggested
scale is applied and information on rainfall, soil and land-use sys- that future applications should use objective modelling criteria for
tems is available. evaluation (e.g. as has been suggested by Alewell and
The meta-analysis of published articles using USLE-type Manderscheid (1998) or Brazier et al. (2000)). However, the latter
modelling to estimate soil loss by water erosion from local to would not solve the paradigm of the principally different nature of
continental scale showed an increasing trend especially in the last modelled (gross) versus measured (net) erosion rates by USLE-type
20 years. The models are widely used on all continents and within models. As such, parallel to increasing research efforts in soil
109 countries. The long-tradition of articles published mainly in the erosion measurement and investing in rigorous application of
US in the early ‘80s and ‘90s, has been continued in Europe, evaluation criteria, the path where soil erosion is connected to
Australia and Southern America and more recently increasingly in topography, runoff and flow path dynamic as well as depositional
Asia and Africa with scientists addressing also emerging research regime should be further followed and solutions should be found
topics connected to soil erosion such as climate change, nutrients for large-scale applications.
balance, freshwater pollution, land use and management and pro-
tection measures.
As discussed above, soil erosion modelling is, like any other References
ecosystem model, only a representation of reality and not reality
Aiello, A., Adamo, M., & Canora, F. (2015). Remote sensing and GIS to assess soil
itself, and is thus prone to deviate from reality and predicted un- erosion with RUSLE3D and USPED at river basin scale in southern Italy. Catena,
certainty might be more or less substantial. As such, for each 131, 174e185.
modelling endeavor a deep understanding of underlying processes, Alewell, C., Birkholz, A., Meusburger, K., Wildhaber, Y. S., & Mabit, L. (2016).
Quantitative sediment source attribution with compound-specific isotope
assumptions and model structure is necessary to avoid misinter- analysis in a C3 plant-dominated catchment (central Switzerland). Bio-
pretation (e.g. in case of USLE-Type modelling the conceptual geosciences, 13(5), 1587e1596.
confusion between on-site gross erosion risk and off-site net Alewell, C., Egli, M., & Meusburger, K. (2015). An attempt to estimate tolerable soil
erosion rates by matching soil formation with denudation in Alpine grasslands.
sediment yields). However, the above discussed validation at- Journal of Soils and Sediments, 15(6), 1383e1399.
tempts showed that when appropriately parameterized soil loss Alewell, C., & Manderscheid, B. (1998). Use of objective criteria for the assessment
estimates with USLE-type models are within the order of magni- of biogeochemical ecosystem models. Ecological Modelling, 107(2), 213e224.
Amundson, R., Berhe, A. A., Hopmans, J. W., Olson, C., Sztein, A. E., & Sparks, D. L.
tude compared to measured soil loss rates and relative differences (2015). Soil and human security in the 21st century. Science, 348(6235), 6.
between areas, management scenarios and time periods seem to be Angulo-Martinez, M., Begueria, S., & Kysely, J. (2016). Use of disdrometer data to
captured correctly in many studies. Based on this, the modelling evaluate the relationship of rainfall kinetic energy and intensity (KE-I). Science
of the Total Environment, 568, 83e94.
approach can be recommended to be used to target risk areas,
Angulo-Martinez, M., Lopez-Vicente, M., Vicente-Serrano, S. M., & Begueria, S.
scenario calculation for conservation tools or other management (2009). Mapping rainfall erosivity at a regional scale: A comparison of inter-
plans, spatial comparison of regions, countries and continents and polation methods in the Ebro basin (NE Spain). Hydrology and Earth System
temporal comparisons of differing time periods. At the current state Sciences, 13(10), 1907e1920.
Arnhold, S., Lindner, S., Lee, B., Martin, E., Kettering, J., Nguyen, T. T., et al. (2014).
of knowledge, USLE-type models are the only model algorithms Conventional and organic farming: Soil erosion and conservation potential for
which have been used at large scales (e.g. continental, global), with row crop cultivation. Geoderma, 219, 89e105.
C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225 221

Arnold, J. G., Srinivasan, R., Muttiah, R. S., & Williams, J. R. (1998). Large area hy- Iowa. American Society of Agronomy Journal, 39, 65e73.
drologic modeling and assesment Part I: Model development. JAWRA Journal of Cao, L., Zhang, K., Dai, H., & Liang, Y. (2015). Modeling interrill erosion on unpaved
the American Water Resources Association, 34(1), 73e89. roads in the loess plateau of China. Land Degradation and Development, 26(8),
Arnoldus, H. M. J. (1980). An approximation of the rainfall factor in the universal soil 825e832.
loss equation. In M. De Boodts, & D. Gabriels (Eds.), Assessments of erosion (pp. Cerdan, O., Govers, G., Le Bissonnais, Y., Van Oost, K., Poesen, J., Saby, N., et al. (2010).
127e132). John Wiley and Sons Ltd. Rates and spatial variations of soil erosion in Europe: A study based on erosion
Aronica, G., & Ferro, V. (1997). Rainfall erosivity over the Calabrian region. Hydro- plot data. Studies on Large Volume Landslides, 122(1e2), 167e177.
logical Sciences Journal-Journal Des Sciences Hydrologiques, 42(1), 35e48. Cronshey, R. G., & Theurer, F. D. (1998). AnnAGNPS-non-point pollutant loading model
Auerswald, K., Fiener, P., Gomez, J. A., Govers, G., Quinton, J. N., & Strauss, P. (2015). (Las Vegas).
Comment on "Rainfall erosivity in Europe" by Panagos et al. (Sci. Total Environ., De Jong, S. M. (1994). Derivation of vegetative variables from a landsat tm image for
511, 801-814, 2015). Science of the Total Environment, 532, 849e852. modelling soil erosion. Earth Surface Processes and Landforms, 19, 165e178.
Auerswald, K., Fiener, P., Martin, W., & Elhaus, D. (2014). Use and misuse of the K De Roo, A. P. J., Wesseling, C. G., & Ritsema, C. J. (1996). Lisem: A single-event
factor equation in soil erosion modeling: An alternative equation for deter- physically based hydrological and soil erosion model for drainage basins. I:
mining USLE nomograph soil erodibility values. Catena, 118, 220e225. Theory, input and output. Hydrological Processes, 10(8), 1107e1117.
Auerswald, K., Kainz, M., & Fiener, P. (2003). Soil erosion potential of organic versus De Vente, J., & Poesen, J. (2005). Predicting soil erosion and sediment yield at the
conventional farming evaluated by USLE modelling of cropping statistics for basin scale: Scale issues and semi-quantitative models. Earth Science Reviews,
agricultural districts in Bavaria. Soil Use and Management, 19(4), 305e311. 71(1e2), 95.
Bagarello, V., Di Piazza, G. V., Ferro, V., & Giordano, G. (2008). Predicting unit plot Desmet, P., & Govers, G. (1996). A GIs procedure for automatically calculating the
soil loss in Sicily, south Italy. Hydrological Processes, 22(5), 586e595. USLE LS factor on topographically complex landscape units. Journal of Soil and
Bagarello, V., Di Stefano, V., Ferro, V., Giordano, G., Iovino, M., & Pampalone, V. Water Conservation, 51(5), 427e433.
(2012). Estimating the USLE soil erodibility factor in Sicily, South Italy. Applied Di Stefano, C., Ferro, V., & Pampalone, V. (2017). Testing the USLE-M family of
Engineering in Agriculture, 28(2), 199e206. models at the Sparacia experimental site in South Italy. Journal of Hydrologic
Bagarello, V., Di Stefano, C., Ferro, V., & Pampalone, V. (2017). Predicting maximum Engineering, 22(8), 05017012-05017012.
annual values of event soil loss by USLE-type models. Catena, 155, 10e19. Diodato, N., Borrelli, P., Fiener, P., Bellocchi, G., & Romano, N. (2017). Discovering
Bagarello, V., & Ferro, V. (2004). Plot-scale measurement of soil erosion at the historical rainfall erosivity with a parsimonious approach: A case study in
experimental area of Sparacia (southern Italy). Hydrological Processes, 18(1), Western Germany. Journal of Hydrology, 544.
141e157. Diodato, N., Knight, J., & Bellocchi, G. (2013). Reduced complexity model for
Bagarello, V., Ferro, V., & Pampalone, V. (2015). A new version of the USLE-MM for assessing patterns of rainfall erosivity in Africa. Global and Planetary Change,
predicting bare plot soil loss at the Sparacia (South Italy) experimental site. 100, 183e193.
Hydrological Processes, 29(19), 4210e4219. Dissmeyer, G. E., & Foster, G. R. (1980). A guide for predicting sheet and rill erosion on
Bakker, M. M., Govers, G., van Doorn, A., Quetier, F., Chouvardas, D., & Rounsevell, M. forest land.
(2008). The response of soil erosion and sediment export to land-use change in Doetterl, S., Van Oost, K., & Six, J. (2012). Towards constraining the magnitude of
four areas of Europe: The importance of landscape pattern. Geomorphology, 98, global agricultural sediment and soil organic carbon fluxes. Earth Surface Pro-
213e226. cesses and Landforms, 37(6), 642e655.
Barenblatt, G. I. (1979). Similarity, self-similarity and intermediate asymptotics. New Dokuchaev, V. V. (1892). Nashi stepi prezde i teper(Our steppes before and recently).
York: Consultants Bureau. Russia(in Russian). Sankt_Peterburg: Published House Evdokimov.
Barenblatt, G. I. (1987). Dimensional analysis (Amsterdam). Dunn, M., & Hickey, R. (1998). The effect of slope algorithms on slope estimates
Belyaev, V. R., Wallbrink, P. J., Golosov, V. N., Murray, A. S., & Sidorchuk, A. Y. (2005). within a GIS. Cartography, 27(1), 9e15.
A comparison of methods for evaluating soil redistribution in the severely Evans, R. (2013). Assessment and monitoring of accelerated water erosion of
eroded Stavropol region, southern European Russia. Geomorphology, 65(3e4), cultivated land - when will reality be acknowledged? Soil Use and Management,
173e193. 29(1), 105e118.
Ben-Dor, E., & Banin, A. (1995). Near-infrared analysis as a rapid method to Evans, R., & Boardman, J. (2016a). The new assessment of soil loss by water erosion
simultaneously evaluate several soil properties. Soil Science Society of America in Europe. Panagos P. et al., 2015 Environmental Science & Policy 54, 438-447-A
Journal, 59(2), 364e372. response. Environmental Science & Policy, 58, 11e15.
Bennett, H. H. (1939). A permanent loss to new england soil erosion resulting from Evans, R., & Boardman, J. (2016b). A reply to panagos et al., 2016 (Environmental
the hurricane. Geographical Review, 29(2), 196e204. science & policy 59 (2016) 53-57. Environmental Science & Policy, 60, 63e68.
Boardman, J. (2006). Soil erosion science: Reflections on the limitations of current Fan, J. C. (1987). Measurements of erosion on highway slopes and of the universal soil
approaches. Catena, 68(2e3), 73e86. loss erosion equation. West Lafayette: Purdue Univ. (Ind).
Bonilla, C. A., & Vidal, K. L. (2011). Rainfall erosivity in Central Chile. Journal of FAO. (2015). GLADIS - global land degradation information system.
Hydrology, 410(1e2), 126e133. FAO. (2017). Voluntary Guidelines for sustainable soil management food and agricul-
Bonta, J. V., & Sutton, P. (1983). Erosion and reclamation plots: Research on the ture organization of the united nations. http://www.fao.org/3/a-bl813e.pdfhttp://
hydroogy and water quality of watersheds subjected to surface mining. Bureau of www.fao.org/3/a-bl813e.pdf (Rome, Italy).
Mines, U.S: Dept. Interior. Favis-Mortlock, D. T. (1997). Validation of field scale soil erosion models using
Borrelli, P., Diodato, N., & Panagos, P. (2016a). Rainfall erosivity in Italy: A national common data sets. In J. Boardman, & D. T. F.-M (Eds.), Modeling soil erosion by
scale spatio-temporal assessment. International Journal of Digital Earth, 9(9). WaterNATO ASI series (pp. 89e127). Berlin: Springer-Verlag.
Borrelli, P., Meusburger, K., Ballabio, C., Panagos, P., & Alewell, C. (2018). Object- Ferro, V. (1997). Further remarks on a distributed approach to sediment delivery.
oriented soil erosion modelling: A possible paradigm shift from potential to Hydrological Sciences Journal, 42(5), 633e647.
actual risk assessments in agricultural environments. Land Degradation & Ferro, V. (2010). Deducing the USLE mathematical structure by dimensional analysis
Development, 29(4), 1270e1281. and self-similarity theory. Biosystems Engineering, 106(2), 216e220.
Borrelli, P., Panagos, P., Langhammer, J., Apostol, B., & Schütt, B. (2016b). Assessment Ferro, V., & Minacapilli, M. (1995). Sediment delivery processes at basin-scale.
of the cover changes and the soil loss potential in European forestland: First Hydrological Sciences Journal-Journal Des Sciences Hydrologiques, 40(6), 703e717.
approach to derive indicators to capture the ecological impacts on soil-related Ferro, V., & Porto, P. (1999). A comparative study of rainfall erosivity estimation for
forest ecosystems. Ecological Indicators, 60. southern Italy and southeastern Australia. Hydrological Sciences Journal-Journal
Borrelli, P., Panagos, P., Ma €rker, M., Modugno, S., & Schütt, B. (2017). Assessment of Des Sciences Hydrologiques, 44(1), 3e24.
the impacts of clear-cutting on soil loss by water erosion in Italian forests: First Ferro, V., & Porto, P. (2000). Sediment delivery distributed (SEDD) model. Journal of
comprehensive monitoring and modelling approach. Catena, 149, 770e781. Hydrologic Engineering, 5(4), 411e422.
Borrelli, P., Paustian, K., Panagos, P., Jones, A., Schütt, B., & Lugato, E. (2016c). Effect Fiener, P., & Auerswald, K. (2016). Comment on "The new assessment of soil loss by
of good agricultural and environmental conditions on erosion and soil organic water erosion in Europe" by Panagos et al. (Environmental Science & Policy 54
carbon balance: A national case study. Land Use Policy, 50, 408e421. (2015) 438-447). Environmental Science & Policy, 57, 140e142.
Borrelli, P., Robinson, D. A., Fleischer, L. R., Lugato, E., Ballabio, C., Alewell, C., Fishman, G. S., & Kiviat, P. J. (1968). The statistics of discrete event simulation.
Meusburger, K., Modugno, S., Schütt, B., Ferro, V., Bagarello, V., Oost, K. V., Simulation, 10, 185e191.
Montanarella, L., & Panagos, P. (2017b). An assessment of the global impact of Foley, J. (2017). Living by the lessons of the planet. Science, 356(6335), 251e252.
21st century land use change on soil erosion. Nature Communications, 8(1), Foster, G. R. (1982). Relation of Usle Factors to Erosion on Rangeland. Science and
2013. Education Administration Publications. ARM(NW-26).
Borst, H. L., & McCall, A. G. (1945). Investigations in erosion control and the recla- Foster, G. R., & Wischmeier, W. H. (1974). Evaluating irregular slopes for soil loss
mation of eroded land at the Northwest Appalachian conservation experiment prediction. Transactions of the Asae, 17(2), 305e309.
station. OH: Zanesville. Foster, G. R., Yoder, D. C., Weesies, G. A., McCool, D. K., McGregor, K. C., &
Brazier, R. E., Beven, K. J., Freer, J., & Rowan, J. S. (2000). Equifinality and uncertainty Bingner, R. L. (2008). Draft User's Guide, Revised Universal Soil Loss Equation
in physically based soil erosion models: Application of the GLUE methodology Version 2 (RUSLE-2). Washington, DC: USDA-Agricultural Research Service.
to WEPPethe water erosion prediction projectefor sites in the UK and USA. Foster, G. R., Yoder, D. C., Weesies, G. A., & Toy, T. J. (2001). The design philosophy
Earth Surface Processes and Landforms, 25(8), 825e845. behind RUSLE2: Evolution of an empirical model. In Soil erosion research for the
Brown, L. C., & Foster, G. R. (1987). Storm erosivity using idealized intensity dis- 21st century, proceedings (pp. 95e98).
tributions. Transactions of the Asae, 30(2), 379e386. Gabriels, D., Ghekiere, G., Schiettecatte, W., & Rottiers, I. (2003). Assessment of USLE
Browning, G. M., Parish, C. L., & Glass, J. A. (1947). A method for determining the use cover-management C-factors for 40 crop rotation systems on arable farms in
and limitation of rotation and conservation practices in control of soil erosion in the Kemmelbeek watershed, Belgium. Soil and Tillage Research, 74, 47e53.
222 C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225

Garcia-Ruiz, J. M., Begueria, S., Nadal-Romero, E., Gonzalez-Hidalgo, J. C., Lana- Larson, W. E., Lindstrom, M. J., & Schumacher, T. E. (1997). The role of severe storms
Renault, N., & Sanjuan, Y. (2015). A meta-analysis of soil erosion rates across the in soil erosion: A problem needing consideration. Journal of Soil and Water
world. Geomorphology, 239, 160e173. Conservation, 52, 90e95.
Gessesse, B., Bewket, W., & Bra €uning, A. (2015). Model based characterization and Lee, J. H., & Heo, J. H. (2011). Evaluation of estimation methods for rainfall erosivity
monitoring of runoff and soil erosion in response to land use/land cover based on annual precipitation in Korea. Journal of Hydrology, 409(1e2), 30e48.
changes in the. Land Degradation & Development, 26, 711e724. February 2014. Leitinger, G., Tasser, E., Newesely, C., Obojes, N., & Tappeiner, U. (2010). Seasonal
van de Giesen, N., Stomph, T. J., & de Ridder, N. (2005). Surface runoff scale effects in dynamics of surface runoff in mountain grassland ecosystems differing in land
west African watersheds: Modeling and management options. Agricultural use. Journal of Hydrology, 385(1e4), 95e104.
Water Management, 72(2), 109e130. Leonard, R. A., Knisel, W. G., & Still, D. A. (1987). GLEAMS: Groundwater loading
Griffin, M. L., Beasley, D. B., Fletcher, J. J., & Foster, G. R. (1988). Estimating soil loss effects of agricultural management systems. Transactions of the ASAE, 30(5),
on topographically non-uniform field and farm units. Journal of Soil and Water 1403e1418.
Conservation, 43(4), 326e331. Li, L., Du, S., Wu, L., & Liu, G. (2009). An overview of soil loss tolerance. Catena, 78(2),
Haciyakupoglu, S., Hizal, A., Gokbulak, F., & Kiziltas, S. (2010). A comparison of soil 93e99.
losses estimated using fallout radionuclides and the USLE for the Omerli Littleboy, M., Freebairn, D. M., Hammer, G. L., & Silburn, D. M. (1992a). Impact of
watershed, Turkey. In M. Zlatic (Ed.), Global change: Challenges for soil man- soil-erosion on production in cropping systems .2. simulation of production and
agement. Advances in geoecology (pp. 175e183). erosion risks for a wheat cropping system. Australian Journal of Soil Research,
Haileslassie, A., Priess, J., Veldkamp, E., Teketay, D., & Lesschen, J. P. (2005). 30(5), 775e788.
Assessment of soil nutrient depletion and its spatial variability on smallholders' Littleboy, M., Silburn, D. M., Freebairn, D. M., Woodruff, D. R., Hammer, G. L., &
mixed farming systems in Ethiopia using partial versus full nutrient balances. Leslie, J. K. (1992b). Impact of soil-erosion on production in cropping systems .1.
Agriculture Ecosystems & Environment, 108(1), 1e16. development and validation of a simulation-model. Australian Journal of Soil
Hart, G. E. (1984). Erosion from simulated rainfall on mountain rangeland in Utah. Research, 30(5), 757e774.
Journal of Soil and Water Conservation, 39, 330e334. Liu, B. Y., Nearing, M. A., & Risse, L. M. (1994). Slope gradient effects on soil loss for
Hengl, T., de Jesus, J. M., MacMillan, R. A., Batjes, N. H., Heuvelink, G. B. M., steep slopes. Transactions of the Asae, 37(6), 1835e1840.
Ribeiro, E., et al. (2014). SoilGrids1km-Global soil information based on auto- Liu, B. Y., Zhang, K. L., & Xie, Y. (2002). An empirical soil loss equation. In Proc. of 12th
mated mapping. PLoS One, 9(8), 17. ISCO. Beijing: Tsinghua Press.
Hickey, R. (2000). Slope angle and slope length solutions for GIS. Cartography, 29(1), Lloyd, C. H., & Eley, G. W. (1952). Graphical solution of probable soil loss formula for
1e8. northeastern region. Journal of Soil and Water Conservation, 7, 189e191.
Hussein, M. H., Kariem, T. H., & Othman, A. K. (2007). Predicting soil erodibility in Loureiro, N. D., & Coutinho, M. D. (2001). A new procedure to estimate the RUSLE
northern Iraq using natural runoff plot data. Soil & Tillage Research, 94(1), EI30 index, based on monthly rainfall data and applied to the Algarve region,
220e228. Portugal. Journal of Hydrology, 250(1e4), 12e18.
Ito, A. (2007). Simulated impacts of climate and land-cover change on soil erosion Lufafa, A., Tenywa, M. M., Isabirye, M., Majaliwa, M. J. G., & Woomer, P. L. (2003).
and implication for the carbon cycle, 1901 to 2100. Geophysical Research Letters, Prediction of soil erosion in a Lake Victoria basin catchment using a GIS-based
34(9). Universal Soil Loss model. Agricultural Systems, 76(3), 883e894.
Jetten, V., Govers, G., & Hessel, R. (2003). Erosion models: Quality of spatial pre- Luleva, M. I., van der Werff, H., Jetten, V., & van der Meer, F. (2011). Can infrared
dictions. Hydrological Processes, 17, 887e900. spectroscopy Be used to measure change in potassium nitrate concentration as
Joel, A., Messing, I., Seguel, O., & Casanova, M. (2002). Measurement of surface water a proxy for soil particle movement? Sensors (Basel, Switzerland), 11(4),
runoff from plots of two different sizes. Hydrological Processes, 16(7), 4188e4206.
1467e1478. Ma €rker, M., Angeli, L., Bottai, L., Costantini, R., Ferrari, R., Innocenti, L., et al. (2008).
Johnson, C. W., Savabi, M. R., & Loomis, S. A. (1984). Rangeland erosion measure- Assessment of land degradation susceptibility by scenario analysis: A case study
ments for the USLE. Transactions of the ASAE, 27, 1313e1320. in southern Tuscany, Italy. Geomorphology, 93(1e2), 120e129.
Kara, O., Şensoy, H., & Bolat, I. (2010). Slope length effects on microbial biomass and Martin, P. H. (1996). Physics of stamp-collecting? Thoughts on ecosystem model
activity of eroded sediments. Journal of Soils and Sediments, 10(3), 434e439. design. Science of the Total Environment, 183(1e2), 7e15.
Karydas, C. G., & Panagos, P. (2016). Modelling monthly soil losses and sediment McCool, D. K., Brown, L. C., Foster, G. R., Mutchler, C. K., & Meyer, L. D. (1987).
yields in Cyprus. International Journal of Digital Earth, 9(8), 766e787. Revised slope steepness factor for the universal soil loss equation. Transactions
Karydas, C. G., Panagos, P., & Gitas, I. Z. (2014). A classification of water erosion of the Asae, 30(5), 1387e1396.
models according to their geospatial characteristics. International Journal of McCool, D., Foster, G., & Weesies, G. (1997). Slope length and steepness factors (LS).
Digital Earth, 7(3), 229e250. In K. Renard, G. Foster, G. Weesies, D. McCool, & D. Yoder (Eds.), Predicting soil
Keesstra, S. D., Bouma, J., Wallinga, J., Tittonell, P., Smith, P., Cerda , A., et al. (2016). erosion by water: A guide to conservation planning with the revised universal soil
The significance of soils and soil science towards realization of the United loss equation (RUSLE) (pp. 101e141). USA: Department of Agriculture, Agricul-
Nations Sustainable Development Goals. SOIL, 2(2), 111e128. tural Research Service, Washington, District of Columbia.
Kilinc, M., & Richardson, E. V. (1973). Mechanics of soil erosion from overland flow Merritt, W. S., Letcher, R. A., & Jakeman, A. J. (2003). A review of erosion and
generated by simulated rainfall. Ft. Collins: Colorado State Univ. sediment transport models. Environmental Modelling & Software, 18(8e9),
King, C., & Delpont, G. (1993). Spatial assessment of erosion: Contribution of remote 761e799.
sensing, a review. Remote Sensing Reviews, 7(3e4), 223e232. Meusburger, K., Banninger, D., & Alewell, C. (2010a). Estimating vegetation
Kinnell, P. I. A. (1981). Rainfall intensity kinetic-energy relationships for soil loss parameter for soil erosion assessment in an alpine catchment by means of
prediction. Soil Science Society of America Journal, 45(1), 153e155. QuickBird imagery. International Journal of Applied Earth Observation and Geo-
Kinnell, P. I. A. (2008). Sediment delivery from hillslopes and the universal soil loss information, 12(3), 201e207.
equation: Some perceptions and misconceptions. Hydrological Processes, 22(16), Meusburger, K., Konz, N., Schaub, M., & Alewell, C. (2010b). Soil erosion modelled
3168e3175. with USLE and PESERA using QuickBird derived vegetation parameters in an
Kinnell, P. I. A. (2010). Event soil loss, runoff and the universal soil loss equation alpine catchment. International Journal of Applied Earth Observation and Geo-
family of models: A review. Journal of Hydrology, 385(1e4), 384e397. information, 12(3), 208e215.
Kinnell, P. I. A. (2014). Applying the QREI30 index within the USLE modelling Meusburger, K., Mabit, L., Park, J. H., Sandor, T., & Alewell, C. (2013). Combined use of
environment. Hydrological Processes, 28(3), 591e598. stable isotopes and fallout radionuclides as soil erosion indicators in a forested
Kinnell, P. I. A., & Risse, L. M. (1998). USLE-M: Empirical modeling rainfall erosion mountain site, South Korea. Biogeosciences, 10(8), 5627e5638.
through runoff and sediment concentration. Soil Science Society of America Meusburger, K., Steel, A., Panagos, P., Montanarella, L., & Alewell, C. (2012). Spatial
Journal, 62(6), 1667e1672. and temporal variability of rainfall erosivity factor for Switzerland. Hydrology
Kirkby, M. J., & Cox, N. J. (1995). A climatic index for soil-erosion potential (csep) and Earth System Sciences, 16.
including seasonal and vegetation factors. Catena, 25(1-4), 333e352. Mitasova, H., Hofierka, J., Zlocha, M., & Iverson, L. R. (1996). Modelling topographic
Kirkby, M. J., Irvine, B. J., Jones, R. J. A., Govers, G., Boer, M., Cerdan, O., et al. (2008). potential for erosion and deposition using GIS. International Journal of
The PESERA coarse scale erosion model for Europe. I. - model rationale and Geographical Information Systems, 10(5), 629e641.
implementation. European Journal of Soil Science, 59(6), 1293e1306. Mitasova, H., Hofierka, J., Zlocha, M., & Iverson, L. (1997). Reply to comment by
Knisel, W. G. (1980). Creams: A field scale model for chemicals, runoff and erosion from Desmet and Govers. International Journal of Geographical Information Science,
agricultural management systems; conservation research report No. 26 (pp. 11(6), 611e618.
1e640). US Departement of Agriculture. Mitasova, H., & Mitas, L. (1999). Modeling soil detachmentwith RUSLE 3D using GIS.
Laflen, J. M., Elliot, W. J., Flanagan, D. C., Meyer, C. R., & Nearing, M.a. (1997). WEPP- University of Illinois at Urbana-Champaign. date accessed February 2018 http://
predicting water erosion using a process-based model. Journal of Soil and Water skagit.meas.ncsu.edu/~helena/gmslab/erosion/usle.html.
Conservation, 52, 96. Mo €ller, M., Gerstmann, H., & Gla €ßer, C. (2014). Dyna C : A framework for the dynamic
Laflen, J. M., & Flanagan, D. C. (2013). The development of U. S. soil erosion pre- derivation of the usle C -factor using high temporal multi-spectral satellite imagery.
diction and modeling. International Soil and Water Conservation Research, 1(2), Mondal, A., Khare, D., Kundu, S., Mukherjee, S., Mukhopadhyay, A., & Mondal, S.
1e11. (2017). Uncertainty of soil erosion modelling using open source high resolution
Laflen, J. M., Foster, G. R., & Onstad, C. A. (1985). Simulation of individual- storm soil and aggregated DEMs. Geoscience Frontiers, 8(3), 425e436.
loss for modeling the impact of soil erosion on crop productivity. Soil Erosion Montanarella, L. (2015). Govern our soils. Nature, 528(7580), 32e33.
and Conservation, 285e295. Moore, I. D., & Burch, G. J. (1986). Physical basis of the length-slope factor in the
Laflen, J. M., & Moldenhauer, W. C. (2003). Pioneering soil erosion prediction e the universal soil loss equation. Soil Science Society of America Journal, 50(5),
USLE story. Bejing, China: World Association of Soil and Water Conservation. 1294e1298.
C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225 223

Morgan, R. P. C. (2005). Soil loss conservation: Problems and prospects. Chichester: Total Environment, 479e480(1), 189e200.
John Wiley & Sons. Panagos, P., Meusburger, K., Ballabio, C., Borrelli, P., Begueria, S., Klik, A., et al.
Morgan, R. P. C., & Nearing, M. A. (2011). Handbook of Erosion Modelling (pp. 1e401). (2015e). Reply to the comment on "Rainfall erosivity in Europe" by Auerswald
John Wiley and Sons. et al. Science of the Total Environment, 532, 853e857.
Morgan, R. P. C., Quinton, J. N., Smith, R. E., Govers, G., Poesen, J. W. A., Auerswald, K., Panagos, P., Meusburger, K., Van Liedekerke, M., Alewell, C., Hiederer, R., &
et al. (1998). The European soil erosion model (EUROSEM): A process-based Montanarella, L. (2014c). Assessing soil erosion in Europe based on data
approach for predicting soil loss from fields and small catchments. , 527-544. collected through a European network. Soil Science and Plant Nutrition, 60(1),
Earth Surface Processes and Landforms, 23(6), 527e544. 15e29.
Morton, A., & Súarez, M. (2001). Kinds of models. In M. G. Anderson, & P. D. Bates Peralta, C., Beriain, E., Iglesias, C., Marangunic, C., Marangunic, A., & Casassa, G.
(Eds.), Model validation: Perspectives in hydrological science (pp. 11e21). Chi- (2017). Ground topography validation of a high-resolution DEM and ortho-
chester: John Wiley & Sons. photos obtained with a commercial UAS. In 2017 first IEEE international sym-
Musgrave, G. W. (1947). The quantitative evaluation of factors in water erosion, a posium of geoscience and remote sensing (GRSS-Chile) (Vol. 4, p. 4).
first approximation. Journal of Soil and Water Conservation, 2, 133e138. Pineiro, G., Paruelo, J. M., Oesterheld, M., & Jobbagy, E. G. (2010). Pathways of
Nash, J. E., & Sutcliffe, J. E. (1970). River flow forecasting through conceptual models. grazing effects on soil organic carbon and nitrogen. Rangeland Ecology &
Part 1 - a discussion of principles. Journal of Hydrology, 10, 282e290. Management, 63(1), 109e119.
Nearing, M. A. (1997). A single, continuous function for slope steepness influence on Podmanicky, L., Balazs, K., Belenyesi, M., Centeri, C., Kristof, D., & Kohlheb, N. (2011).
soil loss. Soil Science Society of America Journal, 61(3), 917e919. Modelling soil quality changes in Europe. An impact assessment of land use
Nearing, M. (2004). Soil erosion and conservation. In J. Wainwright, & M. Mulligan change on soil quality in Europe. Ecological Indicators, 11(1), 4e15.
(Eds.), Environmental modelling: Finding simplicity in complexity (pp. 277e290). Poesen, J. W., Torri, D., & Bunte, K. (1994). Effects of rock fragments on soil-erosion
John Wiley & Sons, Ltd. by water at different spatial scales - a review. Catena, 23(1e2), 141e166.
Nearing, M. A. (2013). Soil erosion and conservation. Porto, P. (2016). Exploring the effect of different time resolutions to calculate the
Nearing, M. A., Govers, G., & Norton, L. D. (1999). Variability in soil erosion data from rainfall erosivity factor R in Calabria, southern Italy. Hydrological Processes,
replicated plots. Soil Science Society of America Journal, 63(6), 1829. 30(10), 1551e1562.
Nearing, M. A., Yin, S. Q., Borrelli, P., & Polyakov, V. O. (2017). Rainfall erosivity: An Quinn, P., Beven, K., Chevallier, P., & Planchon, O. (1991). The prediction of hillslope
historical review. Catena, 157. flow paths for distributed hydrological modelling using digital terrain models.
Nicks, A. D., Lane, L. J., & Gander, G. A. (1995). Weather generator (chapter 2). In Hydrological Processes, 5(1), 59e79.
D. C. Flanagan, & M. A. Nearing (Eds.), USDA water erosion prediction project: Quinton, J. (2013). Erosion and sediment transport. In J. Wainwright, & M. Mulligan
Hillslope profile and watershed model documentation- NSERL report. USDA-ARS (Eds.), Environmental modelling: Finding simplicity in complexity (pp. 187e196).
National Soil Erosion Research Laboratory. John Wiley & Sons, Ltd.
Nyssen, J., Vandenreyken, H., Poesen, J., Moeyersons, J., Deckers, J., Haile, M., et al. Rapp, J. F. (1994). Error assessment of the revised universal soil loss equation using
(2005). Rainfall erosivity and variability in the northern Ethiopian highlands. natural runoff plot data, Master Thesis. Arizona: The University of Arizona.
Journal of Hydrology, 311(1e4), 172e187. Refsgaard, J. C., & Storm, B. (1996). Construction, calibration and validation of hy-
O'Callaghan, J. F., & Mark, D. M. (1984). The extraction of drainage networks from drological models. In M. B. Abbott, & J. C. Refsgaard (Eds.), Distributed hydro-
digital elevation data. Computer Vision, Graphics, and Image Processing, 28(3), logical modelling (pp. 41e45). Dordrecht: Reidel.
323e344. Rejman, J., & Brodowski, R. (2005). Rill characteristics and sediment transport as a
Oldeman, L. R. (1992). Global extent of soil degradation. ISRIC Bi-annual report 1991- function of slope length during a storm event on loess soil. Earth Surface Pro-
1992. cesses and Landforms, 30(2), 231e239.
Oliveira, P. T. S., Wendland, E., & Nearing, M. A. (2013). Rainfall erosivity in Brazil: A Renard, K. G., & Ferreira, V. A. (1993). Rusle model description and database
review. Catena, 100, 139e147. sensitivity. Journal of Environmental Quality, 22(3), 458e466.
Oreskes, N., Shraderfrechette, K., & Belitz, K. (1994). Verification, validation, and Renard, K. G., & Foster, G. R. (1985). Managing rangeland soil resources: The uni-
confirmation of numerical-models in the earth-sciences. Science, 263(5147), versal soil loss equation. Rangelands, 7(3), 118e122.
641e646. Renard, K., Foster, G., Weesies, G., McCool, D., & Yoder, D. (1997). Predicting soil
ndez-Ugalde, O. (2018). LUCAS
Orgiazzi, A., Ballabio, C., Panagos, P., Jones, A., & Ferna erosion by water: A guide to conservation planning with the revised universal soil
soil, the largest expandable soil dataset for Europe: A review. European Journal loss equation (RUSLE). Agriculture handbook 703. Washington, District of
of Soil Science, 69(1), 140e153. Columbia, USA: Department of Agriculture, Agricultural Research Service.
Ozsoy, G., Aksoy, E., Dirim, M. S., & Tumsavas, Z. (2012). Determination of soil Renard, K. G., Foster, G. R., Weesies, G. A., & Porter, J. P. (1991). Rusle - revised
erosion risk in the mustafakemalpasa river basin, Turkey, using the revised universal soil loss equation. Journal of Soil and Water Conservation, 46(1), 30e33.
universal soil loss equation, geographic information system, and remote Risal, A., Bhattarai, R., Kum, D., Park, Y. S., Yang, J. E., & Lim, K. J. (2016). Application
sensing. Environmental Management, 50(4), 679e694. of Web ERosivity module (WERM) for estimation of annual and monthly R
Panagos, P., Ballabio, C., Borrelli, P., Meusburger, K., Klik, A., Rousseva, S., et al. factor in Korea. Catena, 147, 225e237.
(2015a). Rainfall erosivity in Europe. Science of the Total Environment, 511, Risse, L. M., Nearing, M. A., Laflen, J. M., & Nicks, A. D. (1993). Error assessment in
801e814. the universal soil loss equation. Soil Science Society of America Journal, 57(3),
Panagos, P., Borrelli, P., Meusburger, K., Alewell, C., Lugato, E., & Montanarella, L. 825.
(2015b). Estimating the soil erosion cover-management factor at the European Rodriguez, J. L. G., & Suarez, M. C. G. (2012). Methodology for estimating the
scale. Land Use Policy, 48, 38e50. topographic factor LS of RUSLE3D and USPED using GIS. Geomorphology, 175,
Panagos, P., Borrelli, P., Meusburger, K., van der Zanden, E. H., Poesen, J., & Alewell, C. 98e106.
(2015c). Modelling the effect of support practices (P-factor) on the reduction of € mkens, M., Poesen, J., & Wang, J. (1988). Relationship between the USLE soil
Ro
soil erosion by water at European scale. Environmental Science and Policy, 51. erodibility factor and soil properties. Proceedings, 371e385.
Panagos, P., Borrelli, P., Meusburger, K., Yu, B., Klik, A., Jae Lim, K., et al. (2017a). Rosewell, C. J. (1993). SOILOSS: A program to assist in the selection of management
Global rainfall erosivity assessment based on high-temporal resolution rainfall practices to reduce erosion. Technical handbook (Vol. 11). Soil Conservation Ser-
records. Scientific Reports, 7(1), 4175. vice of New South Wales.
Panagos, P., Borrelli, P., Meusburger, K., Yu, B., Klik, A., Lim, k.j., et al. (2017b). Global Rossel, R. A. V., & Behrens, T. (2010). Using data mining to model and interpret soil
rainfall erosivity assessment based on high-temporal resolution rainfall records. diffuse reflectance spectra. Geoderma, 158(1), 46e54.
Scientific Reports. In review(February): 1-12. Rüttimann, M., Schaub, D., Prasuhn, V., & Rüegg, W. (1995). Measurement of runoff
Panagos, P., Borrelli, P., Poesen, J., Ballabio, C., Lugato, E., Meusburger, K., et al. and soil erosion on regularly cultivated fields in Switzerland d some critical
(2015d). The new assessment of soil loss by water erosion in Europe. Environ- considerations. Catena, 25(1), 127e139.
mental Science & Policy, 54, 438e447. Sadeghi, S. H., Zabihi, M., Vafakhah, M., & Hazbavi, Z. (2017). Spatiotemporal
Panagos, P., Borrelli, P., Poesen, J., Meusburger, K., Ballabio, C., Lugato, E., et al. mapping of rainfall erosivity index for different return periods in Iran. Natural
(2016b). Reply to the comment on "The new assessment of soil loss by water Hazards, 87(1), 35e56.
erosion in Europe" by Fiener & Auerswald. Environmental Science & Policy, 57, Scheurer, K., Alewell, C., Banninger, D., & Burkhardt-Holm, P. (2009). Climate and
143e150. land-use changes affecting river sediment and brown trout in alpine countries-
Panagos, P., Borrelli, P., Poesen, J., Meusburger, K., Ballabio, C., Lugato, E., a review. Environmental Science and Pollution Research, 16(2), 232e242.
Montanarella, L., & Alewell, C. (2016a). Reply to "The new assessment of soil loss Schindler Wildhaber, Y., Michel, C., Burkhardt-Holm, P., Baenninger, D., & Alewell, C.
by water erosion in Europe. In P. Panagos, et al. (Eds.), 2015 Environ. Sci. Policy 54 (2012). Measurement of spatial and temporal fine sediment dynamics in a small
(Vol. 59, pp. 438e447). Environmental Science & Policy. A response" by Evans river. Hydrology and Earth System Sciences, 16(5), 1501e1515.
and Boardman Environ. Sci. Policy 58, 11-15. 53-57. Schmidt, S., Alewell, C., & Meusburger, K. (2018). Mapping spatio-temporal dy-
Panagos, P., Christos, K., Cristiano, B., & Ioannis, G. (2014a). Seasonal monitoring of namics of the cover and management factor (C-factor) for grasslands in
soil erosion at regional scale: An application of the G2 model in Crete focusing Switzerland. Remote Sensing of Environment, 211, 89e104.
on agricultural land uses. International Journal of Applied Earth Observation and Schmidt, S., Alewell, C., & Meusburger, K. (2019). Monthly RUSLE soil erosion risk of
Geoinformation, 27, 147e155. Swiss grasslands. Journal of Maps, 1e10.
Panagos, P., Imeson, A., Meusburger, K., Borrelli, P., Poesen, J., & Alewell, C. (2016c). Schmidt, S., Alewell, C., Panagos, P., & Meusburger, K. (2016). Regionalization of
Soil conservation in Europe: Wish or reality? Land Degradation & Development, monthly rainfall erosivity patterns in Switzerland. Hydrology and Earth System
27(6), 1547e1551. Sciences, 20(10), 4359e4373.
Panagos, P., Meusburger, K., Ballabio, C., Borrelli, P., & Alewell, C. (2014b). Soil Scho€nbrodt, S., Saumer, P., Behrens, T., Seeber, C., & Scholten, T. (2010). Assessing the
erodibility in Europe: A high-resolution dataset based on LUCAS. Science of the USLE crop and management factor C for soil erosion modeling in a large
224 C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225

mountainous watershed in Central China. Journal of Earth Science, 21, 835e845. processing of digital elevation data using a Cþþ executable. Computers &
Schwertmann, U., Vogl, W., & Kainz, M. (1987). Bodenerosion durch Wasser - Geosciences, 30(9e10), 1043e1053.
vorhersage des Abtrags und Bewertung von Gegenmaßnahmen. Stuttgart: Verlag Van Rompaey, A. J. J., Verstraeten, G., Van Oost, K., Govers, G., & Poesen, J. (2001).
Eugen Ulmer. Modelling mean annual sediment yield using a distributed approach. Earth
Seibert, J., & McGlynn, B. L. (2007). A new triangular multiple flow direction algo- Surface Processes and Landforms, 26(11), 1221e1236.
rithm for computing upslope areas from gridded digital elevation models. Van der Knijff, J., Jones, R. R. J., & Montanarella, L. (1999). Soil erosion risk assessment
Water Resources Research, 43(4). in Italy. Luxemburg: Off. Publ. Eur. Communities EUR 19022.
Senarath, S. U. S., Ogden, F., Downer, C. W., & Sharif, H. O. (2000). On the calibration Vannoppen, W., Vanmaercke, M., De Baets, S., & Poesen, J. (2015). A review of the
and verification of two-dimensional, distributed, Hortonian, continuous mechanical effects of plant roots on concentrated flow erosion rates. Earth-
watershed models. Water Resources Research, 36, 1495e1510. Science Reviews, 150, 666e678.
Shepherd, K. D., & Walsh, M. G. (2002). Development of reflectance spectral libraries Veihe, A., & Quinton, J. (2000). Sensitivity analysis of EUROSEM using Monte Carlo
for characterization of soil properties. Soil Science Society of America Journal, simulation I: Hydrological, soil and vegetation parameters. Hydrological Pro-
66(3), 988e998. cesses, 14(5), 915e926.
Silburn, D. M., & Freebairn, D. M. (1992). Evaluations of the creams model .3. Veihe, A., Quinton, J., & Poesen, J. (2000). Sensitivity analysis of EUROSEM using
simulation of the hydrology of vertisols. Australian Journal of Soil Research, Monte Carlo simulation II: The effect of rills and rock fragments. Hydrological
30(5), 547e564. Processes, 14(5), 927e939.
Silburn, D. M., & Loch, R. J. (1989). Evaluation of the creams model .1. sensitivity Verheijen, F. G. A., Jones, R. J. A., Rickson, R. J., & Smith, C. J. (2009). Tolerable versus
analysis of the soil-erosion sedimentation component for aggregated clay soils. actual soil erosion rates in Europe. Earth-Science Reviews, 94(1e4), 23e38.
Australian Journal of Soil Research, 27(3), 545e561. Vrieling, A. (2006). Satellite remote sensing for water erosion assessment: A review.
Simanton, J., Roger, H. B., Herbert, B., & Renard, K. G. (1980). Application of the USLE Catena, 65(1), 2e18.
to southwestern rangelands. Hydrology and water resources in Arizona and the Wall, G. J., Coote, D. R., Pringle, E. A., & Shelton, I. J. (2002). RUSLEFAC d revised
Southwest. universal soil loss equation for application in Canada: A handbook for estimating
Sivertun, A., & Prange, L. (2003). Non-point source critical area analysis in the soil loss from water erosion in Canada. ECORC contribution number 02-92, AAFC
Gisselo watershed using GIS. Environmental Modelling & Software, 18(10), contribution number AAFC/AAC2244E. Ottawa, Ontario: Research Branch Agri-
887e898. culture and Agri-Food Canada.
Smith, D. D. (1941). Interpretation of soil conservation data for field use. Agricultural Wall, G. J., Dickinson, W. T., Rudra, R. P., & Coote, D. R. (1988). Seasonal soil erod-
Engineering, 22, 173e175. ibility variation in southwestern ontario. Canadian Journal of Soil Science, 68(2),
Smith, D. D., & Whitt, D. M. (1948). Evaluating soil losses from field areas. Agri- 417e424.
cultural Engineering, 29, 394e396. Wang, G. Q., Fang, Q. Q., Teng, Y. G., & Yu, J. S. (2016). Determination of the factors
Smith, S. J., Williams, J. R., Menzel, R. G., & Coleman, G. A. (1984). Prediction of governing soil erodibility using hyperspectral visible and near-infrared reflec-
sediment yield from southern plains grasslands with the modified universal soil tance spectroscopy. International Journal of Applied Earth Observation and Geo-
loss equation. Journal of Range Management, 37(4), 295e297. information, 53, 48e63.
Smith, D. D., & Wischmeier, W. H. (1957). Factors affecting sheet and rill erosion. Wang, B., Zheng, F., & Wang, Y. (2012). Adaptability analysis on soil erodibility
American Geophysical Union Transaction, 38, 889e896. models in typical thin layer black soil area of Northeast China. Nongye Gong-
Spaeth, K. E., Pierson, F. B., Weltz, M.a., & Blackburn, W. H. (2003). Evaluation of cheng Xuebao/Transactions of the Chinese Society of Agricultural Engineering,
USLE and RUSLE estimated soil loss on rangeland. Journal of Range Managment, 28(6), 126e131.
56, 234e246. Warren, S. D., Mitasova, H., Hohmann, M. G., Landsberger, S., Iskander, F. Y.,
Stenberg, B., & Rossel, R. A. V. (2010). Diffuse reflectance spectroscopy for high- Ruzycki, T. S., et al. (2005). Validation of a 3-D enhancement of the Universal
resolution soil sensing. In R. A. Viscarra Rossel, A. B. McBratney, & B. Minasny Soil Loss Equation for prediction of soil erosion and sediment deposition. Ca-
(Eds.), Proximal soil sensing (pp. 29e47). Dordrecht: Springer Netherlands. tena, 64(2e3), 281e296.
Stolpe, N. B. (2005). A comparison of the RUSLE, EPIC and WEPP erosion models as Wawer, R., & Nowocien, E. (2006). Cyfrowa mapa wystepowania erozji wodnej
calibrated to climate and soil of south-central Chile. Acta Agriculturae Scandi- powierzchniowej obszaru Polski z wykorzystaniem CORINE Land Cover 2000.
navica Section B Soil and Plant Science, 55(1), 2e8. Rozdzielczosc przestrzenna 500 m. Roczniki Akademii Rolniczej w Poznaniu.
Sverdrup, H., Warfvinge, P., Blake, L., & Goulding, K. (1995). Modeling recent and Rolnictwo, 65, 207e213.
historic soil data from the rothamsted-experimental-station, UK using safe. Weltz, M. A., Kidwell, M., & Fox, H. D. (1998). Influence of abiotic and biotic factors
Agriculture Ecosystems & Environment, 53(2), 161e177. in measuring and modeling soil erosion on rangelands: State of knowledge.
Tan, Z., Leung, L. R., Li, H.-Y., & Tesfa, T. (2018). Modeling sediment yield in land Journal of Range Management, 51, 482e495.
surface and earth system models: Model comparison, development, and eval- Weltz, M. A., Renard, K. G., & Simanton, J. R. (1987). Revised universal soil loss
uation. Journal of Advances in Modeling Earth Systems, 0(0). equation for western rangelands, General technical report RM-Rocky Mountain
Teng, H. F., Rossel, R. A. V., Shi, Z., Behrens, T., Chappell, A., & Bui, E. (2016). Forest and Range Experiment Station. USA): US Department of Agriculture, Forest
Assimilating satellite imagery and visible-near infrared spectroscopy to model Service.
and map soil loss by water erosion in Australia. Environmental Modelling & Wendt, R. C., Alberts, E. E., & Hjelmfelt, A. T. (1986). Variability of runoff and soil loss
Software, 77, 156e167. from fallow experimental plots. Soil Science Society of America Journal, 50(3),
Tiwari, A. K., Risse, L. M., & Nearing, M. A. (2000). Evaluation of WEPP and its 730e736.
comparison with USLE and RUSLE. Transactions of the Asae, 43(5), 1129e1135. Williams, J. R. (1975). Sediment-yield prediction with universal equation using runoff
Torri, D., Poesen, J., & Borselli, L. (1997). Predictability and uncertainty of the soil energy factor, present and prospective technology for predicting sediment yield and
erodibility factor using a global dataset. Catena, 31(1), 1e22. sources. US Department of Agriculture.
Toy, T. J., F, G. R., & R, K. G. (2002). Soil erosion: Processes, prediction, measurement, Williams, J. R., & Berndt, H. D. (1977). Sediment yield prediction based on watershed
and control. New York: John Wiley & Sons, Inc. hydrology. Transactions of the Asae, 20(6), 1100e1104.
Tresch, S. (2014). Determination of the S-factor in steep slopes. Master Thesis,. Basel: Williams, J. R., Renard, K. G., & Dyke, P. T. (1983). Epic: A new method for assessing
University Basel. erosion's effect on soil productivity. Journal of Soil and Water Conservation,
Trimble, S. W., & Crosson, P. (2000). U.S. Soil erosion rates–myth and reality. Science, 38(5), 381e383.
289(5477), 248e250. Winchell, M., Jackson, S., Wadley, A., & Srinivasan, R. (2008). Extension and vali-
USDA. (1961). US agricultural research service. a universal equation for predicting dation of a geographic information system-based method for calculating the
rainfall-erosion losses. U.S. Dept. Agr. ARS, 22e66, 1e11. Revised Universal Soil Loss Equation length-slope factor for erosion risk as-
USDA. (2008). Revised universal soil loss equation version 2 (RUSLE2). Science docu- sessments in large watersheds. Journal of Soil and Water Conservation, 63(3),
mentation. Washington, D.C: USDA-Agricultural Research Service. 105e111.
USDA. (2018). Revised universal soil loss equation 2 - download RUSLE2. https://www. Wischmeier, W. H. (1959). A rainfall erosion index for a universal SOIL-loss equa-
ars.usda.gov/southeast-area/oxford-ms/national-sedimentation-laboratory/ tion. Soil Science Society of America Proceedings, 23, 246e249.
watershed-physical-processes-research/research/rusle2/revised-universal-soil- Wischmeier, W. H. (1960). Cropping-management factor evaluations for a universal
loss-equation-2-download-rusle2/. soil-loss equation. Soil Science Society of America Proceedings, 24, 322e326.
Van Oost, K., Govers, G., & Desmet, P. (2000a). Evaluating the effects of changes in Wischmeier, W. H. (1966). Relation of field-plot runoff to management and physical
landscape structure on soil erosion by water and tillage. Landscape Ecology, Factors1. Soil Science Society of America Journal, 30(2), 272e277.
15(6), 577e589. Wischmeier, W. H. (1975). Estimating the soil loss equations cover and management
Van Oost, K., Govers, G., Van Muysen, W., & Quine, T. A. (2000b). Modeling trans- factor for undisturbed areas, Present and prospective technology for predicting
location and dispersion of soil constituents by tillage on sloping land. Soil Sci- sediment yields and sources. Oxford.
ence Society of America Journal, 64(5), 1733e1739. Wischmeier, W. H. (1976). Use and misuse of the universal soil loss equation. Journal
Van Oost, K., Quine, T. A., Govers, G., De Gryze, S., Six, J., Harden, J. W., et al. (2007). of Soil and Water Conservation, 3, 5e9.
The impact of agricultural soil erosion on the global carbon cycle. Science, 318, Wischmeier, W. H., Johnson, C. B., & Cross, B. V. (1971). A soil erodibility nomograph
626e629. for farmland and construction sites. Journal of Soil and Water Conservation, 26,
Van Remortel, R. D., Hamilton, M. E., & Hickey, R. J. (2001). Estimating the LS factor 189e193.
for RUSLE through iterative slope length processing of digital elevation data Wischmeier, W. H., & Smith, D. D. (1958). Evaluation of factors in the soil-loss
within arclnfo grid. Cartography, 30(1), 27e35. equation. Agricultural Engineering, 39(8), 458e462.
Van Remortel, R. D., Maichle, R. W., & Hickey, R. J. (2004). Computing the LS factor Wischmeier, W. H., & Smith, D. D. (1965). Predicting rainfall-erosion losses from
for the Revised Universal Soil Loss Equation through array-based slope cropland east of the Rocky Mountains. Agriculture Handbook (Vol 282).
C. Alewell et al. / International Soil and Water Conservation Research 7 (2019) 203e225 225

Washington DC: US Department of Agriculture. risk. Land Degradation & Development, 20(1), 84e91.
Wischmeier, W. H., & Smith, D. D. (1978). Predicting rainfall erosion losses - a guide to Zhang, X., Drage, N. A., & Wainwright, J. (2004b). Scaling issues in environmental
conservation planning, agriculture handbook No. 537. USDA/Science and Educa- modelling. In J. Wainwright, & M. Mulligan (Eds.), Environmental modelling:
tion Administration, US. Govt. Printing Office, Washington D.C. Finding simplicity in complexity (pp. 319e334). John Wiley & Sons, Ltd.
Xu, X. L., Liu, W., Kong, Y. P., Zhang, K. L., Yu, B., & Chen, J. D. (2009). Runoff and Zhang, K., Li, S., Peng, W., & Yu, B. (2004a). Erodibility of agricultural soils on the
water erosion on road side-slopes: Effects of rainfall characteristics and slope loess plateau of China. Soil and Tillage Research, 76(2), 157e165.
length. Transportation Research Part D: Transport and Environment, 14(7), Zhang, X. C., Nearing, M. A., Risse, L. M., & McGregor, K. C. (1996b). Evaluation of
497e501. runoff and soil loss predictions using natural runoff plot data. Transactions of the
Xu, X., Miao, C., & Liao, Y. (2008). Quantitative remote sensing study on regional soil American Society of Agricultural Engineers, 39(3), 855e863.
erosion. Wuhan University Journal of Natural Sciences, 13(3), 377e384. Zhang, X. C., Nearing, M. A., Risse, L. M., & McGregor, K. C. (1996c). Evaluation of
Yair, A., & Raz-Yassif, N. (2004). Hydrological processes in a small arid catchment: WEPP runoff and soil loss predictions using natural runoff plot data. Trans-
Scale effects of rainfall and slope length. Geomorphology, 61(1e2), 155e169. actions of the Asae, 39(3), 855e863.
Yang, D., Kanae, S., Oki, T., Koike, T., & Musiake, K. (2003a). Global potential soil Zhang, L., O'Neill, A. L., & Lacey, S. (1996a). Modelling approaches to the prediction
erosion with reference to land use and climate changes. Hydrological Processes, of soil erosion in catchments. Environmental Software, 11(1), 123e133.
17(14), 2913e2928. Zhang, K. L., Shu, A. P., Xu, X. L., Yang, Q. K., & Yu, B. (2008). Soil erodibility and its
Yang, D. W., Kanae, S., Oki, T., Koike, T., & Musiake, K. (2003b). Global potential soil estimation for agricultural soils in China. Journal of Arid Environments, 72(6),
erosion with reference to land use and climate changes. Hydrological Processes, 1002e1011.
17(14), 2913e2928. Zhang, F. B., Wang, Z. L., & Yang, M. Y. (2014). Validating and improving interrill
Yin, S., Xie, Y., Liu, B., & Nearing, M. A. (2015). Rainfall erosivity estimation based on erosion equations. PLoS One, 9(2), 11.
rainfall data collected over a range of temporal resolutions. Hydrology and Earth Zhang, H., Yang, Q., Li, R., Liu, Q., Moore, D., He, P., et al. (2013). Extension of a GIS
System Sciences, 19(10), 4113e4126. procedure for calculating the RUSLE equation LS factor. Computers & Geo-
Young, R. A., Onstad, C. A., Bosch, D. D., & Anderson, W. P. (1989). AGNPS - A sciences, 52, 177e188.
nonpoint-source pollution model for evaluating agricultural watersheds. Jour- Zhang, W., Zhang, Z., Liu, F., Qiao, Z., & Hu, S. (2011). Estimation of the USLE cover
nal of Soil and Water Conservation, 44(2), 168e173. and management factor C using satellite remote sensing: A review. In Pro-
Yue, Y., Ni, J. R., Ciais, P., Piao, S. L., Wang, T., Huang, M. T., et al. (2016). Lateral ceedings - 2011 19th international conference on geoinformatics, geoinformatics
transport of soil carbon and land-atmosphere CO2 flux induced by water 2011.
erosion in China. Proceedings of the National Academy of Sciences of the United Zingg, A. W. (1940a). Degree and length of land slope as it affects soil loss in runoff.
States of America, 113(24), 6617e6622. Agricultural Engineering, 21(2), 59e64.
Zhang, Y., Degroote, J., Wolter, C., & Sugumaran, R. (2009). Integration of modified Zingg, A. W. (1940b). Degree and length of land slope as it affects soil loss in runoff.
universal soil loss equation (musle) into a gis framework to assess soil erosion Agricultural Engineering, 21, 59e64.

You might also like