You are on page 1of 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/51185057

The assessment of spatial distribution of soil salinity risk using neural


network

Article  in  Environmental Monitoring and Assessment · June 2011


DOI: 10.1007/s10661-011-2132-5 · Source: PubMed

CITATIONS READS

41 182

2 authors:

Akmal Akramkhanov Paul L. G. Vlek


International Center for Agricultural Research in the Dry Areas (ICARDA) University of Bonn
39 PUBLICATIONS   615 CITATIONS    370 PUBLICATIONS   12,094 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

GLOWA Volta View project

Soil quality degradation assessment View project

All content following this page was uploaded by Akmal Akramkhanov on 22 April 2015.

The user has requested enhancement of the downloaded file.


Environ Monit Assess
DOI 10.1007/s10661-011-2132-5

The assessment of spatial distribution of soil salinity risk


using neural network
Akmal Akramkhanov · Paul L. G. Vlek

Received: 14 September 2010 / Accepted: 5 May 2011


© Springer Science+Business Media B.V. 2011

Abstract Soil salinity in the Aral Sea Basin is one vs. 1.04 dS m−1 ). Visual comparison of the maps
of the major limiting factors of sustainable crop suggests that the estimated map had soil salinity
production. Leaching of the salts before planting that was uniform in distribution. The upscaling
season is usually a prerequisite for crop estab- proved to be satisfactory; depending on critical
lishment and predetermined water amounts are salinity threshold values, around 70–90% of loca-
applied uniformly to fields often without discern- tions were correctly estimated.
ing salinity levels. The use of predetermined wa-
ter amounts for leaching perhaps partly emanate Keywords Irrigated agriculture · Upscaling ·
from the inability of conventional soil salinity Validation · Spatial variation ·
surveys (based on collection of soil samples, lab- Environmental correlation
oratory analyses) to generate timely and high-
resolution salinity maps. This paper has an ob-
jective to estimate the spatial distribution of soil Introduction
salinity based on readily or cheaply obtainable en-
vironmental parameters (terrain indices, remote Soil salinity in the Aral Sea Basin is one of the ma-
sensing data, distance to drains, and long-term jor limiting factors of sustainable crop production.
groundwater observation data) using a neural net- Leaching of the salts before planting season is
work model. The farm-scale (∼15 km2 ) results usually a prerequisite for crop establishment and
were used to upscale soil salinity to a district area predetermined water amounts are applied uni-
(∼300 km2 ). The use of environmental attributes formly to fields often without discerning salinity
and soil salinity relationships to upscale the spatial levels. The use of predetermined water amounts
distribution of soil salinity from farm to district for leaching perhaps partly emanate from the in-
scale resulted in the estimation of essentially sim- ability of conventional soil salinity surveys (based
ilar average soil salinity values (estimated 0.94 on collection of soil samples, laboratory analy-
ses) to generate timely and high-resolution salinity
maps. The spatial distribution of soil salinity over
the landscape depends on many environmental
A. Akramkhanov (B) · P. L. G. Vlek factors (Evans and Caccetta 2000; McKenzie and
Center for Development Research (ZEF),
Ryan 1999). While intensive soil sampling can be
University of Bonn, Walter-Flex-Str. 3,
Bonn, Germany 53113 eliminated or significantly reduced by an elec-
e-mail: akmal@zef.uzpak.uz, api001@yahoo.com tromagnetic (EM) induction tool that measures
Environ Monit Assess

electrical conductivity of the soil (Lesch et al. sify soil salinity risk potential. Within the soil-
1995; Rhoades et al. 1999), there are few studies landscape dimension, Zhu (2000) showed that the
on relationship between environmental variables neural network approach reveals much greater
and EM values to estimate soil salinity in irrigated spatial detail and has higher quality than that
areas. derived from the conventional soil map. Similarly,
Similar to many other successful attempts McBratney et al. (2000) gave an example of clay
(Lavado et al. 2006; Leij et al. 2004; Moore et al. content estimation where neural networks per-
1993) to develop analogues of conventional sur- formed among the best. In all these studies, as well
veys by terrain attributes, this study builds on as in the majority of other studies, feed-forward
readily available or easily extractable data to ac- networks have been used. Feed-forward neural
count for the spatial distribution of soil salinity. networks are composed of layers of neurons, in
The focus here was whether it would be possible which the input layer of neurons is connected to
to develop models using the available existing the output layer of neurons. The training process
data sets regardless their intrinsic shortcomings. If is undertaken by changing the weights such that a
so, it would help decision making under the con- desired input–output relationship is realized.
ditions encountered by farmers or water resource Multilayer perceptrons (MLPs) are layered
managers. feed-forward networks typically trained with sta-
The salinity distribution was modeled using the tic backpropagation (Principe et al. 2000). With
identified factors at the farm scale (Akramkhanov backpropagation, the input data is repeatedly pre-
2005) and this model was used to upscale the salin- sented to the neural network. With each presenta-
ity distribution to the district level. Soil salinity is tion, the output of the neural network is compared
a dynamic soil property particularly in irrigated to the desired output and an error is computed.
areas of Khorezm region of Uzbekistan. Emphasis This error then is fed back (backpropagated) to
was made to test the feasibility of the model to the neural network and used to adjust the weights
estimate soil salinity during summer period, when such that the error decreases with each iteration
most irrigation takes place and soil salinity in the and the neural model gets closer and closer to pro-
area is subject to variation. The model developed ducing the desired output (this process is known
in such variable environment will be robust and as “training”). Principe et al. (2000) emphasize
applicable for other time periods. that their main advantage is in being easy to use,
The complex nature of the environmental vari- and the key disadvantages are that they train
ables and assumptions in classic statistical analysis slowly and require a large amount of training
often make the use of statistical theories difficult. data (typically three times more training samples
An alternative is the use of artificial neural than network weights). However, MLPs with a
network (ANN), the computerized information sufficient number of hidden units can approximate
processing model, the development of which was any continuous function to a pre-specified accu-
motivated by the way the brain works. It is able racy; in other words, MLP networks are universal
to modify its structure by learning from the data approximators (Cherkassky and Mulier 1998).
rather than being programmed, and is thus able to In this study, ANN is used as an alternative
solve complicated relations (Principe et al. 2000). to regression techniques for prediction of the soil
In salinity studies, ANN have been used for a salinity distribution. More than 400 cases of field-
variety of purposes, including the prediction of a collected data are used for model development.
salt build-up in the crop root zone (Patel et al. The overall research objective of the study was
2002), dielectric constant–soil water relationships to develop a model to estimate soil salinity distri-
(Persson et al. 2002), and river water salinity fore- bution at the farm and district level with an em-
casting (Bowden et al. 2002; Maier and Dandy phasis on using readily available and cheaply ob-
1999). tainable measurements, so that real-time decision
In studies with a spatial domain, Eklund et al. making can be supported both on-farm (by farm-
(1998) applied a neural network algorithm for ers) and off-farm (by policy makers) in Khorezm,
inductive inference from GIS layers to clas- Uzbekistan.
Environ Monit Assess

Materials and methods district scale, with the difference that rice is grown
in many areas of the district, whereas on the
Area description study farm rice was not grown during the sampling
year.
The data for training of the model at the farm The climate in Khorezm is arid with an annual
scale was collected in 2002 on the research farm precipitation of about 100 mm ranging from 35 to
(41◦ 36 N, 60◦ 31 E) of the Urgench State Uni- 170 mm during dry and wet years, respectively.
versity and the surrounding farm southwest of About 70% of it falls in the winter and spring.
the city of Khiva, in Khorezm, Uzbekistan. The The topography of the district is similar to that of
sampling area (ca. 15 km2 ) consists of sands, sandy the farm, i.e., flat, with elevation points normally
loams, and silt loams at equal proportions. Most distributed ranging from 85 to 110 m above sea
of the area is cultivated land. However, bare or level (mean 93 m and standard deviation 2.9). The
abandoned land was also included in the study landscape is dissected by an extensive network
to possibly increase the variability in soil salinity. of drains and collectors (Fig. 1). The soils in the
More details of the farm-scale description can be district are classified by the local classification sys-
found in Akramkhanov (2005) and Akramkhanov tem as desert zone meadow-oasis soils and by the
et al. (2008). Resulting trained neural network World Reference Base as Cambisols (Abdullaev
model from farm scale in 2002 is applied to up- 2003). At finer resolution some soils match
scaling area in 2003. Solonchak description but depending on location,
The upscaling area is a district. Khiva district is once leached, fall into a complex of arenosols–
one of 11 districts of the Khorezm region and is fluvisols–anthrosols descriptions. These soils have
situated in the southwest of the region. The total moderately high humus content (0.8–1.4%) com-
area of upscaling in this study is approximately pared to the rest of the soils in this desert zone.
300 km2 (Fig. 1), but data is gathered from sur- Since the land in the area has been constantly
rounding areas as well to avoid edge effect dur- under development, many studies distinguish soils
ing interpolation analysis. A range of attributes by irrigation duration, basically referring to them
present at farm scale are also represented at the as soils of ancient or recent development.

Fig. 1 Location of study


area for model
development and district
area for upscaling
Environ Monit Assess

Soil salinity and environmental variables Data on groundwater table depth and its salin-
ity were obtained from over 500 observation wells
Soil salinity (CMv) was estimated by electromag- installed in the study area by the Hydrological
netic induction device (CM-138) which measures Melioration Expedition of the Khorezm Depart-
apparent electric conductivity (ECa ) of a soil ment of Land and Water Resources. The average
profile down to 1.5 m depth in vertical mode. value of groundwater table depth and salinity in
The following variables from selected environ- groundwater observation wells collected in July
mental factors were derived from varying sources for the period from 1990 to 2002 were averaged
for the farm and district scale inputs. The soil and used for interpolation and feeding into the
texture for the farm scale input was analyzed on model. It was assumed that long-term average
samples collected during a survey from top 30 cm. data represent the spatial distribution of ground-
We assume that topsoil is representative of the water table depth and salinity better than single-
1 m layer. The soil texture for the district level year observations.
was estimated using the dataset from 1996 soil sur-
vey provided by Soil Science Institute where the ANN simulation model characteristics
average depth of the top horizons of 51 sampling
sites was 31 cm. For the ease of transferability of The PC-based software package NeuroSolutions
the results in the local context, clay was assumed 4.31 (NeuroDimension, Inc.) was used in this
to consist of soil separates with a size of less study for the artificial neural network operation.
than 0.001 mm and silt to be between 0.001 and The package is integrated into a spreadsheet. The
0.05 mm (Russian classification scheme). steps outlined in Maier and Dandy (1999) and
Elevation data from a 1:10,000-topographic Bowden et al. (2002) were used as a guide in this
map were obtained to create a digital elevation work. The default values suggested in the software
model (DEM). Terrain indices were calculated were used unless stated otherwise.
from a 30 × 30-m raster-based DEM. The fol- The dataset of 441 cases was split into three
lowing terrain indices were calculated using the subsets, namely training, cross-validation, and
software DiGeM 2.0 (Olaf Conrad, Gottingen, testing. The training subset is used to train the
Germany): aspect (as), slope (sl), profile curva- network, the cross-validation subset is used to
ture (profc), plan curvature (planc), curvature avoid neural network overtraining (training is ter-
(curv) (Zevenbergen and Thorne 1987); diver- minated when mean squared error (MSE) of the
gence/convergence (dc) indices, flow accumula- cross-validation subset increases), and the testing
tion (upslope contributing area, ua), wetness (wt), subset is used to test the built model on unseen
and erosivity (based on universal soil loss equa- data. The number of epochs (iterations) was set
tion, ls) (Moore et al. 1993). Additionally, curva- to 10,000 and the network was trained using cross-
ture (curv7) and the terrain characterization in- validation ten times. Bowden et al. (2002) showed
dices (tci) were calculated according to Park et al. that data division into subsets is important for
(2001). Elevation values in meters (alt) above mean the model performance, and three subsets should
sea level and solar radiation (solar) were also used. represent the same population. For example, in-
Agricultural maps (scale 1:10,000 and 1:25,000) dependent subsets were split provided that each
were used to obtain the drain water collector net- included the lowest and the highest values for each
work. The layers were digitized and the nearest dis- variable. Out of 441 cases, the training subset con-
tance from drain collector (dcoll) to sampling point sisted of 265 cases, cross-validation 66 cases, and
was obtained using ArcView 3.2 (ESRI Inc., USA). the testing subset 110 cases, the detailed statistics
Remote sensing parameters were obtained are given in Akramkhanov (2005). Some variables
from a Landsat 7 satellite image acquired on the could not be divided into equal subsets, which can
July 12, 2002. Three variables were selected, the be attributed to singular or rare events present in
ratio vegetation index which is known to delineate the dataset. However, in general, the statistics of
reduced reflectance due to salinity (Wang et al. the subset are in good agreement and appear to
2002) and the raw bands 3 and 5. come from the same population.
Environ Monit Assess

Fig. 2 Relationship
between ECa measured
with electromagnetic
induction device CM-138
in vertical mode and ECe
of the soil for various soil
textures

A multi-layer perceptron with one hidden layer is generated that summarizes the variation of each
with 12 processing elements was built with the output with respect to the variation in each input.
help of a wizard provided in the software. Non-
linearity was introduced by the hyperbolic tan-
Validation survey of the district scale map
gent function (TanhAxon), and the momentum
learning rule was used for updating the weights. The study area was represented by 1,755 points
Predictions of the neural network were assessed randomly generated by the Arcview sample gen-
with the correlation coefficient R between mea- erator extension (DNR 2004). This number of
sured and estimated values and mean squared points was arbitrarily chosen to cover the whole
error. The best weights during these runs were area of the district, and it was expected to provide
applied for the testing subset. The best weights a sufficient number of points. The trained model
were then applied to the production (dataset to be was applied to these 1,755 points with the same
predicted) subset. attributes used in training but derived for the
In addition, sensitivity analysis was performed district area. The resulting map from predicted
to identify those input variables that had more of values of soil salinity was interpolated using the
an effect on estimating soil salinity. It provides inverse distance weighting in ArcMap.
a measure of the relative importance among the The validation of the generated salinity map
inputs of the neural model and illustrates how the was done by ground-truthing at 315 randomly
model output varies in response to input variation. selected locations. Sampling sites were randomly
The first input is varied between its mean ±1 selected but with a view to logistics and accessi-
standard deviations while all other inputs are fixed bility and were required to cover the whole area
at their respective means. The network output is of the delimited district area. Their coordinates
computed for 50 steps above and below the mean. were uploaded into a handheld GPS in advance
This process is repeated for each input. A report and navigation by GPS was used to locate the

Table 1 Neural network results of training; estimates between measured and estimated soil salinity values (dS m−1 )
Subset Measured Estimated Correlation MSE
CMv mean CMv mean coefficient, R
Training 0.51 0.51 0.82 0.04
Cross-validation 0.46 0.51 0.85 0.03
Testing 0.51 0.53 0.83 0.03
Environ Monit Assess

1.8 1.4 1.4


Training subset Cross-validation subset Testing subset
1.6 1.2 1.2
predicted CMv, dS m-1

predicted CMv, dS m-1

predicted CMv, dS m-1


1.4
1 1
1.2
1 0.8 0.8
0.8 0.6 0.6
0.6
0.4 0.4
0.4
0.2 0.2 0.2
0 0 0
0 0.5 1 1.5 2 2.5 0 0.2 0.4 0.6 0.8 1 1.2 1.4 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
measured CMv, dS m-1 measured CMv, dS m-1 measured CMv, dS m-1

Fig. 3 Scatter plots between measured and estimated CMv for different subsets in the neural network

sampling points on the ground. The timeframe and low mean square error between measured
for the validation survey was scheduled to be in and estimated soil salinity values. MSE values are
summer, running for about 4–5 weeks and start- comparable between subsets, which indicate that
ing in July. Vertical electromagnetic conductivity the division into subsets yielded relatively uniform
measurement (CMv, dS m−1 ) was used in the samples in these subsets.
analysis. Sensitivity analysis showed that the response
To compare estimated and measured CMv, we of the model was highly dependent on the fol-
used relationship presented in Fig. 2. For the lowing terrain variables, listed in decreasing order
study area the conversion for loamy texture of of importance: curvature (curv), plan curvature
Rhoades et al. (1999) and Bennett et al. (1995) (planc), profile curvature (profc), and solar radi-
is appropriate, where CMv of 0.6, 1.0, 1.5, and ation (solar). It appears that soil salinity is largely
2.3 dS m−1 roughly correspond to ECe of 4, 6, 8, influenced by the micro-topography, convexity, or
and 12 dS m−1 , respectively. Soils with ECe above concavity, which tends to influence surface water
4 dS m−1 are considered saline. Since soils in the retention.
study area are a mix of light and medium texture,
transition range of critical salinity CMv values
above which soils are considered saline could be Validation results
within 0.6–0.9 dS m−1 . Therefore, we classify CMv
values into saline and not saline levels in esti- The statistics of the estimated and measured
mated and measured datasets using three criti- soil salinity datasets are summarized in Table 2.
cal salinity threshold values such as 0.7, 0.8, and The mean values for both, estimated and mea-
0.9 dS m−1 . For each critical salinity threshold we sured sets, are close to each other (0.94 and
compare estimated with measured CMv classified 1.04 dS m−1 ); MSE and RMSE are 0.05 and 0.25,
as saline (i.e., >0.7 dS m−1 ) and not saline (i.e., respectively. Despite the fact that the average
<0.7 dS m−1 ) classes. We calculate precision of
correctly classified saline values at each selected
critical salinity value.
Table 2 Statistics of trained, measured, and estimated val-
ues of soil salinity (dS m−1 )

Results Trained Estimated Measured


Mean 0.51 0.94 1.04
Neural network model training performance Median 0.51 0.93 1.02
Mode 0.20 1.86 0.96
SD 0.33 0.11 0.19
Statistics of the performance of the neural net- Skewness 0.79 2.21 0.67
work model are summarized in Table 1. The scat- Minimum 0.03 0.41 0.59
ter plots of the measured and estimated salinity Maximum 1.92 1.86 1.78
values for each subset are shown in Fig. 3. The Percentiles 25 0.20 0.88 0.91
results indicate that the model performed well 50 0.51 0.93 1.02
taking into account high correlation coefficients 75 0.73 0.99 1.14
Table 3 Comparison table of estimated and measured salinity maps using three critical threshold values 0.7, 0.8, and 0.9 dS m−1
Threshold Measured Predicted
Environ Monit Assess

0.7 dS m−1 White – not saline Dark – saline

0.8 dS m−1

0.9 dS m−1
Environ Monit Assess

one feature: deviation in the lower and higher


range between measured and estimated values.
The neural network seems to overestimate values
in the low salinity range and to underestimate
those in the higher range. Underestimations of soil
salinity, especially in the vicinity of peak values,
could be caused by the lack of samples with high
values in the training set for the model to learn.
The modeling procedure has to deal with several
issues, namely change of scales and the complex
relationship among the variables involved.
Data fed into the model come from a vari-
Fig. 4 Precision of classifying saline soils of estimated ety of sources and often represent the proxies
maps based on critical salinity threshold values. Threshold
values of 0.7, 0.8, and 0.9 discussed in text are marked of the true covariates obtained by interpolation.
Interpolation has been extensively studied, and
errors resulting from interpolation itself can be
quite substantial. For example, the RMSE for
value of the training subset (0.51 dS m−1 ) was different kriging methods of soil salinity examined
considerably lower than that of the estimated set by Triantafilis et al. (2001) varied from 1 to 1.3
(0.94 dS m−1 ), the model performed well, i.e., it in conditions very similar to this study area. Also,
predicted the average salinity for the district close interpolation is known to smooth the surface and
to the average value of the measured set. the resulting dataset represents less the low and
According to Rhoades et al. (1999) and Bennett high values. Perhaps the lack of such low and high
et al. (1995) average CMv of estimated and mea- values in the number of environmental covari-
sured datasets indicate saline soils in the study ates included in the model training contributed to
area. Spatial distribution of saline and not saline over- or under-estimation of soil salinity.
soils in estimated and measured maps is presented The strong influence of the terrain attributes
in Table 3 using three critical salinity threshold indicated by the sensitivity analysis is encour-
values. Light color indicates not saline soils, dark aging, because it supports the concept of the
color—saline soils. Change in precision using crit- soil–landscape relationship in this study. It ap-
ical salinity threshold values is shown in Fig. 4. pears that soil salinity variability is mainly due to
If we assume that critical salinity threshold the spatial variability of curvatures (curv, planc,
above which soils are considered saline to be equal profc), which tend to control water movement
to 0.7 dS m−1 then the precision of estimating and solar radiation. The neural network was more
saline soils is close to 90%. High precision can also sensitive to these micro-scale controlling factors
be seen from two maps in Fig. 4, where both es- than to those that are obvious at a macro-scale.
timated and measured maps are uniformly saline Kachanoski et al. (1985) observed that A-horizon
throughout. variability was due to spatial variability of surface
With the increase in the critical salinity thresh- curvature, but the relationship was complicated
old to 0.8 and 0.9 dS m−1 precision is reduced due to the presence of positive and negative cor-
to around 80% and 70%, respectively. Estimated relations over different frequency ranges.
maps have more areas classified as saline com- The explanation of spatial distribution of soil
pared to measured dataset. salinity by terrain information, i.e., shape, is at-
tractive for empirical model building, because it
does not require detailed soil surveys over the
Discussion landscape. The major drawback of such models,
however, is that they are scale specific or range
The scatter plots (Fig. 3) of all three subsets (of values it had been trained on) specific, which
used in training of the model appear to have is also noticeable in the scatter plots (Fig. 3).
Environ Monit Assess

The difference between average salinity val- terrain and texture, and on additional parameters
ues of training and measured datasets stands out. from remote sensing and drainage networks. The
Accurate prediction of the average soil salinity emphasis was on terrain attributes because, as
perhaps was due to the presence of the lowest and Moore et al. (1993) and McKenzie et al. (2000)
highest salinity ranges in the training subset dur- proved, these provide a much better basis for
ing the model training. This ability of the neural visualizing other data layers and allow generating
network to model complicated functions demon- sets of data that have value in their own right (e.g.,
strates that they are excellent approximators. slope and solar radiation), even when the exact
The use of environmental attributes and the form of the relationship is unknown.
soil salinity relationship to upscale spatial distri- This study demonstrates that the ANN ap-
bution of soil salinity from farm to district scale proach is suitable for upscaling purposes, and re-
essentially estimated similar mean soil salinity val- sults are valid within a given range of soil salinity.
ues (0.94 vs. 1.04 dS m−1 ). The neural network Timing of measurement is important, and as the
model used to spatially estimate soil salinity val- study was conducted during the summer period,
ues at selected points was able to approximate soil further application of this model to other dis-
salinity despite the possible errors in intermediate tricts of the Khorezm region and validation should
processing of data. This shows that the model involve the summer period only. Application of
based on the included variables can effectively the results to other seasons must be carefully as-
estimate soil salinity. The major advantage of this sessed before making any generalizations about
approach is that available data is used. soil salinity distribution using the model presented
Since soil texture change is transitional rather here. For this, conducting field survey and retrain-
than abrupt the application of one critical salinity ing model with dataset obtained in, for example,
threshold value might not be appropriate. There- autumn.
fore, the precision of the estimated map in identi- The influence of soil and irrigation manage-
fying saline soils is around 70–90% depending on ment on soil salinity is a dynamic feature and
the critical salinity threshold value used. From the should be included in models of soil salinity dis-
experience in the area and other surveys with EM tribution. Capturing and parameterizing this dy-
devices threshold values of 0.8 and 0.9 dS m−1 will namic feature into a model could potentially im-
certainly indicate saline soils, although with lower prove model performance. Therefore, further re-
precision. search is needed to quantify these parameters,
As technology and models are developing which can then be included in the further devel-
rapidly, many methodologies emphasize the opment of such models.
flexibility and adeptness to modifications. The
model can be fine-tuned by incorporation of other
related factors, for instance, geodesic data such as Conclusions
the extent of impermeable layers. By incorporat-
ing new correlating factors, further complication The environmental correlation model built for the
of the model should be avoided. By fine-tuning a farm-scale estimated soil salinity using a neural
“window of opportunity” (Bouma and Hoosbeek network with a correlation coefficient between es-
1996) is rather created for improvement of the timated and measured soil salinity of 0.83. Explicit
model by replacing with either a better explaining prediction of salinity was satisfactory taking into
factor or a more easily obtainable factor. account contrasting scales of soil salinity varia-
Hoosbeek and Bouma (1998) emphasize that tion and environmental data derived from varying
to be able to define suitable indicators for certain scales and unknown accuracy.
soil and land qualities, it is important to have The use of environmental attributes and soil
insight into the characteristics of an indicator. As salinity relationship to upscale spatial distribution
previously attempted by many others, the Jenny of soil salinity from farm to district scale essen-
model was simplified by the assumption that other tially estimated identical mean soil salinity values
parameters are constant and was based largely on (0.94 vs. 1.04 dS m−1 ). Precision of estimation
Environ Monit Assess

of spatial distribution of saline soils was in the Evans, F. H., & Caccetta, P. A. (2000). Broad-scale spatial
range of 70–90% depending on the critical salinity prediction of areas at risk from dryland salinity. Car-
tography, 29, 33–40.
threshold applied.
Hoosbeek, M. R., & Bouma, J. (1998). Obtaining soil and
In general, the results indicate that significant land quality indicators using research chains and geo-
correlations between quantifiable terrain at- statistical methods. Nutrient Cycling in Agroecosys-
tributes and soil salinity exist. This study shows tems, 50, 35–50.
that such a relationship was successfully used to Kachanoski, R. G., Rolston, D. E., & de Jong, E. (1985).
Spatial and spectral relationships of soil properties
estimate soil salinity at farm scale solely based and microtopography: I. Density and thickness of A
on quantified environmental variables. Therefore, horizon. Soil Science Society of America Journal, 49,
the environmental variables must be available for 804–812.
the study area in high resolution or easily measur- Lavado, F., Maneta, M., & Schnabel, S. (2006). Prediction
of near-surface soil moisture at large scale by digital
able. terrain modelling and neural networks. Environmental
Monitoring and Assessment, 121(1–3), 211–230.
Leij, F. J., Romano, N., Palladino, M., Schaap, M. G., &
Acknowledgement This study was funded by the German
Coppola, A. (2004). Topographic attributes to pre-
Ministry for Education and Research (BMBF; project
dict soil hydraulic properties along a hillslope transect.
number 0339970A).
Water Resources Research, 40, W02407.
Lesch, S. M., Strauss, D. J., & Rhoades, J. D. (1995).
Spatial prediction of soil salinity using electromag-
netic induction techniques. 2. An efficient spatial sam-
References pling algorithm suitable for multiple linear regression
model identification and estimation. Water Resources
Abdullaev, U. (2003). Republic of Uzbekistan. Land degra- Research, 31, 387–398.
dation assessment in drylands (LADA). Research and Maier, H. R., & Dandy, G. C. (1999). Empirical com-
Design Institute Uzgipromeliovodkhoz, pp. 43. parison of various methods for training feed-forward
Akramkhanov, A. (2005). The spatial distribution of soil neural networks for salinity forecasting. Water Re-
salinity: Detection and prediction. ZEF Series in Ecol- sources Research, 35, 2591–2596.
ogy and Development No. 32, Bonn, Germany, Uni- McBratney, A. B., Odeh, I. O. A., Bishop, T. F. A.,
versity of Bonn. Dunbar, M. S., & Shatar, T. M. (2000). An overview
Akramkhanov, A., Sommer, R., Martius, C., Hendrickx, of pedometric techniques for use in soil survey. Geo-
J. M. H., & Vlek, P. L. G. (2008). Comparison and sen- derma, 97, 293–327.
sitivity of measurement techniques for spatial distrib- McKenzie, N. J., & Ryan, P. J. (1999). Spatial prediction of
ution of soil salinity. Irrigation and Drainage Systems, soil properties using environmental correlation. Geo-
22, 115–126. derma, 89, 67–94.
Bennett, D. L., George, R. J., & Ryder, A. (1995). Soil McKenzie, N. J., Gessler, P. E., Ryan, P. J., & O’Connell,
salinity assessment using the EM38: Field operating D. A. (2000). The role of terrain analysis in soil map-
instructions and data interpretation. Miscellaneous ping. In J. P. Wilson, & J. C. Gallant (Eds.), Terrain
Publication 4/95, Western Australia, Department of analysis: Principles and applications (p. 245–265). New
Agriculture. York: Wiley.
Bouma, J., & Hoosbeek, M. R. (1996). The contribution Moore, I. D., Gessler, P. E., Nielsen, G. A., & Peterson,
and importance of soil scientists in interdisciplinary G. A. (1993). Soil attribute prediction using terrain
studies dealing with land. In R. J. Wagenet, & J. analysis. Soil Science Society of America Journal, 57,
Bouma (Eds.), The role of soil science in interdscipli- 443–452.
nary research (pp. 1–15). SSSA Special Publication 45. Park, S. J., McSweeney, K., & Lowery, B. (2001). Iden-
Bowden, G. J., Maier, H. R., & Dandy, G. C. (2002). Opti- tification of the spatial distribution of soils using
mal division of data for neural network models in wa- a process-based terrain characterization. Geoderma,
ter resources applications. Water Resources Research, 103, 249–272.
38, 1–11. Patel, R. M., Prasher, S. O., Goel, P. K., & Bassi, R. (2002).
Cherkassky, V., & Mulier, F. (1998). Learning from data: Soil salinity prediction using artificial neural networks.
Concepts, theory, and methods. U.S.A.: Wiley. Journal of the American Water Resources Association,
Department for Natural Resources (2004). DNR random 38, 91–100.
sample generator. Release 2.2. Minnesota: Department Persson, M., Sivakumar, B., Berndtsson, R., Jacobsen,
for Natural Resources. O. H., & Schjonning, P. (2002). Predicting the dielec-
Eklund, P. W., Kirkby, S. D., & Salim, A. (1998). Data tric constant–water content relationship using artificial
mining and soil salinity analysis. International Journal neural networks. Soil Science Society of America Jour-
of Geographical Information Science, 12, 247–268. nal, 66, 1424–1429.
Environ Monit Assess

Principe, J. C., Euliano, N. R., & Lefebvre, C. W. (2000). Wang, D., Wilson, C., & Shannon, M. C. (2002). Inter-
Neural and adaptive systems: Fundamentals through pretation of salinity and irrigation effects on soybean
simulations. New York: Wiley. canopy reflectance in visible and near-infrared spec-
Rhoades, J. D., Chanduvi, F., & Lesch, S. (1999). Soil salin- trum domain. International Journal of Remote Sensing,
ity assessment: Methods and interpretation of electri- 23, 811–824.
cal conductivity measurements. FAO Irrigation and Zevenbergen, L. W., & Thorne, C. R. (1987). Quantita-
Drainage Paper 57. tive analysis of land surface topography. Earth Surface
Triantafilis, J., Odeh, I. O. A., & McBratney, A. B. (2001). Processes and Landforms, 12, 47–56.
Five geostatistical models to predict soil salinity from Zhu, A. X. (2000). Mapping soil landscape as spatial con-
electromagnetic induction data across irrigated cotton. tinua: The neural network approach. Water Resources
Soil Science Society of America Journal, 65, 869–878. Research, 36, 663–677.

View publication stats

You might also like