ArticlePDF Available

Abstract and Figures

Flood irrigation after crop harvest, e.g. autumn irrigation (AI), is a common irrigation practice in arid and semi-arid regions like Hetao Irrigation District (HID) in Northwest China to increase soil moisture and leach soil salt. Detailed information about the extent, timing, and amount of AI is imperative for modeling agro-hydrological processes and irrigation management. However, little attention is given to the identification of the above AI factors. There are basically three major difficulties in estimating the annual changes in AI, including a suitable index to identify AI, temporal instability of thresholds, and an effective validation method for irrigation timing. Therefore, this study proposes a simple and effective threshold-based method to extract the extent and timing of AI in the HID using MODIS water indices at a daily timescale. The Multi-Band Water Index (MBWI) time series is first reconstructed using an adaptive weighted Savitzky-Golay filter and then used to identify the AI extent and time. The proposed model has a stronger generalization capability both in time and space due to robust thresholds selected from the Z-score normalized feature variable. The model is validated both at pixels generated by the segmentation of Sentinel-derived MBWI using a threshold-based model and at sampling points from the field survey. Results show that the model performed well with an overall accuracy of more than 90.0% for the irrigation area. The overall accuracies of irrigation timing are 76.4% and 91.7% based on the middle-to-late and whole irrigation periods, respectively. We found a decreasing trend in the AI area and a gradual delay in the starting time of AI in the HID, mainly due to changes in cropping patterns, climate, and irrigation fees. Overall, the model is promising in identifying flood irrigation extent and timing in large irrigation districts and is helpful for irrigation scheduling.
This content is subject to copyright. Terms and conditions apply.
RESEARCH ARTICLE
Multi-year mapping of ood autumn irrigation extent and timing in harvested
croplands of arid irrigation district
Ximin Qian
a
, Hongwei Qi
a
, Songhao Shang
a
, Heyang Wan
a
and Ruiping Wang
b
a
State Key Laboratory of Hydroscience and Engineering, Department of Hydraulic Engineering, Tsinghua University, Beijing, China;
b
Department of Scientific Research, Bayannur Institute of Water Conservancy Research, Bayannur, China
ABSTRACT
Flood irrigation after crop harvest, e.g. autumn irrigation (AI), is a common irrigation practice in arid
and semi-arid regions like Hetao Irrigation District (HID) in Northwest China to increase soil
moisture and leach soil salt. Detailed information about the extent, timing, and amount of AI is
imperative for modeling agro-hydrological processes and irrigation management. However, little
attention is given to the identication of the above AI factors. There are basically three major
diculties in estimating the annual changes in AI, including a suitable index to identify AI,
temporal instability of thresholds, and an eective validation method for irrigation timing.
Therefore, this study proposes a simple and eective threshold-based method to extract the
extent and timing of AI in the HID using MODIS water indices at a daily timescale. The Multi-
Band Water Index (MBWI) time series is rst reconstructed using an adaptive weighted Savitzky-
Golay lter and then used to identify the AI extent and time. The proposed model has a stronger
generalization capability both in time and space due to robust thresholds selected from the
Z-score normalized feature variable. The model is validated both at pixels generated by the
segmentation of Sentinel-derived MBWI using a threshold-based model and at sampling points
from the eld survey. Results show that the model performed well with an overall accuracy of more
than 90.0% for the irrigation area. The overall accuracies of irrigation timing are 76.4% and 91.7%
based on the middle-to-late and whole irrigation periods, respectively. We found a decreasing
trend in the AI area and a gradual delay in the starting time of AI in the HID, mainly due to changes
in cropping patterns, climate, and irrigation fees. Overall, the model is promising in identifying
ood irrigation extent and timing in large irrigation districts and is helpful for irrigation scheduling.
ARTICLE HISTORY
Received 06 May 2022
Accepted 14 September 2022
KEYWORDS
Flood irrigation; transient
water bodies; irrigation
timing; MODIS; water index;
arid agricultural region
1. Introduction
Flood irrigation in harvested croplands is benecial in
increasing soil water storage, regulating soil tempera-
ture, reducing pests in cropland, and/or leaching salt
(Xiong et al. 2021). Flood irrigation has been adopted
in many regions of the world (Jing et al. 2020; Negri
et al. 2020), especially in the irrigation districts of arid
and semi-arid regions in Northwest China. It is also
known as autumn or winter irrigation (AI) because it is
generally applied in late autumn to early winter after
crop harvest. AI is quite dierent from irrigation dur-
ing the crop growing season. The surface water
inundated by AI often reaches several decimeters
and takes several days to fully inltrate. AI, therefore,
consumes a large amount of water, e.g. AI accounts
for about 1/3 of the annual irrigation water diversion
to the Hetao Irrigation District (HID) from Mid-
September to early November (Lu et al. 2019).
Irrigation during the crop growing season
requires rational management, and so does AI. On
the one hand, improper time and amount of AI may
lead to secondary soil salinization issues and harm
crop production (Mao et al. 2017). On the other
CONTACT Songhao Shang shangsh@tsinghua.edu.cn
GISCIENCE & REMOTE SENSING
2022, VOL. 59, NO. 1, 1598–1623
https://doi.org/10.1080/15481603.2022.2126342
© 2022 The Author(s). Published by Informa UK Limited, trading as Taylor & Francis Group.
This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use,
distribution, and reproduction in any medium, provided the original work is properly cited.
hand, large irrigation districts in the arid regions of
China are subjected to water scarcity problems due
to the limited available water for irrigation.
Therefore, agricultural water conservation and sali-
nity control have become major issues for the sus-
tainable development of irrigation districts in arid
regions (Xu et al. 2010). Under the current circum-
stances of limited water availability, the ecient use
of available water to address the soil salinity and its
leaching eect is a major concern in these regions
(Wen et al. 2020). In addition, the decrease in water
supply for irrigation has led to changes in the spa-
tiotemporal allocation of irrigation water and crop-
ping patterns, which then have signicant impacts
on the hydrological processes, agricultural produc-
tion, and the environment at a regional scale (Shi
et al. 2020). Detailed irrigation information is an
indispensable input to quantifying these changes in
the regional agro-hydrological simulation models
(Wen et al. 2022). Therefore, timely and accurate AI
information is necessary for irrigation and agriculture
administrations as well as for researchers.
Recently, the applications of remote sensing tech-
niques (satellite data) to investigate irrigation during
the crop growing season are signicantly increased
(Karthikeyan, Chawla, and Mishra 2020; Ozdogan et al.
2010). The investigation is performed at dierent spa-
tial scales by adopting optical data (Deines et al. 2019;
Salmon et al. 2015), microwave data (Dari et al. 2021;
Gao et al. 2018), or both (Bazzi et al. 2019; Elwan et al.
2022). Moreover, the investigation of irrigation during
the growing season is mainly focused on the dier-
ences in the greenness index (Xie and Lark 2021) and/
or soil moisture between rain-fed and irrigated crops.
However, very few studies have investigated AI to
date because the above-mentioned methods may
not be able to detect AI. Therefore, there is
a signicant research gap in identifying AI on a large
scale. AI identication is almost similar to the identi-
cation of water bodies because water accumulates
to the depth of several decimeters after the AI appli-
cation. Water index (WI) can be used for the identi-
cation of AI. However, the identication of AI is more
dicult than that of water bodies because the surface
water ponding as a result of AI lasts only for several
days. Therefore, the identication of AI requires high
temporal resolution data, where Moderate Resolution
Imaging Spectroradiometer (MODIS) with the daily
temporal resolution is the most appropriate remote
sensing data for studying short-term inundation (Dao,
Mong, and Chan 2019).
Meanwhile, dierentiation between water, ice, and
snow is also an important aspect in the identication
of AI because the inundated irrigation water after AI
can freeze in some meteorological conditions, where
ice and snow may have WI values close to or higher
than water (Klein and Barnett 2003). Although freez-
ing of water manifests the condition of being irri-
gated, the increase in WI due to freezing can lead to
signicant inconsistencies in feature values for lands
irrigated at dierent times, which hinders the precise
identication of AI extent and time. Therefore, snow
and ice should be handled carefully while dealing
with AI. Although many studies have compared the
performances of dierent WIs for water body detec-
tion (Ji, Zhang, and Wylie 2009), little is known about
their sensitivity to surface water change and freezing.
Therefore, a robust index is required for AI identica-
tion which should be sensitive to changes in surface
water but not to freezing.
Threshold-based models are one of the main
approaches for identifying irrigated land or water
bodies based on the selected index (Pervez and
Brown 2010). The threshold-based model requires the
selection of appropriate feature variables, which can be
divided into two categories. The rst category is based
on the feature value in a specied day or the maxi-
mum/minimum values in the time series (L. Wang et al.
2021). The second category is based on the dynamic
time series, where the combined index of dierent
periods in a specic time prole is selected, or the
time variation of the index is considered (Yeom et al.
2021). It has been proved that the time prole is more
suitable to identify irrigation than the static spectral
signatures at a one-time point (Y. Chen et al. 2018), but
it emphasizes the processing of feature time series.
Another drawback of the threshold-based method is
that the thresholds cannot be compared temporally
(Xie and Lark 2021). Therefore, pre-classication (e.g.
wet and dry years classication (Pervez, Budde, and
Rowland 2014) or data normalization techniques (Xie
and Lark 2021)) were used to obtain thresholds avail-
able for multiple years to track annual changes in
irrigation extent and timing.
The Z-score normalization method is one of the
most promising techniques for normalizing tem-
poral trajectories (Santana et al. 2018). It provides
a statistical measure of deviations from the mean
GISCIENCE & REMOTE SENSING 1599
after taking into account the natural variability
(Kreyszig 2010), and facilitates the user to dene
thresholds without considering the variations in
detected objects due to disturbing factors such as
satellite view angle (DeVries et al. 2020). Z-score has
been successfully applied so far for re detection (De
Carvalho Júnior et al. 2015), tree planting detection
(Chen, Jin, and Brown 2019), ood detection
(DeVries et al. 2020), etc. Therefore, the Z-score-
based threshold model was considered in this
study to identify AI. Similar to the identication of
irrigated paddy rice elds and irrigation timing
(Jeong et al. 2012), the AI threshold can be deter-
mined by statistical data for a particular study area,
whereas AI time refers to the rst time at which the
threshold is reached.
In addition, a sucient number of widely distributed
and representative reference data is critical to identify
and assess the accuracy of remote sensing-based AI
extent and time. Currently, irrigation products are only
available during the growing season and therefore
cannot be used to validate AI identication. Visual
interpretation of the high spatial resolution image or
Google Earth is another common method of obtaining
validation samples for the detection of irrigation areas
(Jin et al. 2016). However, it is dicult to provide
validation for irrigation timing due to the requirements
of short satellite revisit periods and suciently accurate
auxiliary information. Therefore, a simple and fast
method that can extract irrigation distribution and
irrigation time from high-resolution images on a large
scale could be useful for the validation of AI mapping.
To summarize, there are still several key challenges in
detecting the multi-year variations in AI area and timing,
such as appropriate feature extraction, temporal incom-
parability of thresholds, and limited data available for
accuracy assessment for the irrigation time. In order to
address these problems, the main objective of this study
is to develop a threshold-based model capable of
extracting multi-year AI area and timing using remote
sensing-based daily WI time series data, where
a hierarchical decision tree classier is used to determine
the irrigation time for all irrigated pixels. The model is
applied to the HID and validated from multiple perspec-
tives using agricultural statistics, eld survey data, and
information extracted from high-resolution Sentinel-2
images on Google Earth Engine (GEE).
2. Study area
Hetao Irrigation District (HID), located in Inner
Mongolia, is one of the three largest irrigation dis-
tricts in China and the largest in the arid regions of
Northwest China. It is a representative irrigation dis-
trict that relies on AI for storing water and leaching
salt. It consists of ve sub-irrigation districts, includ-
ing Wulanbuhe (WLBH), Jiefangzha (JFZ), Yongji (YJ),
Yichang (YC), and Wulate (WLT) (Figure 1). The
whole HID covers an area of about 1.189 million
ha, where 0.733 million ha (62% of the total) is
irrigated cropland that is mainly sown with corn,
sunower, and wheat. Natural land is another
major land-use type and covers approximately 19%
of the total area of HID, which consists of sparse
grassland with scattered shrubs and abandoned
cropland due to high soil salinity. The groundwater
table depth in HID is generally shallow (Wen et al.
2020), with large annual cyclic variations from 0.5
to 3.0 m.
Agricultural production in HID is highly dependent
on irrigation due to the arid and semi-arid climate
conditions (low precipitation and strong evaporation
potential). As reported in the Bayannur water resources
bulletin, irrigation water is primarily diverted from the
Yellow River, accounting for about 1/7 of the total
water diversion in the Yellow River basin. The total
amount of water diverted from the Yellow River is
approximately 4.0–4.8 billion m
3
per year from 2010
to 2020. Water available for irrigation has decreased in
recent years as a result of decreased runo and inte-
grated water resources management in the Yellow
River basin. A proper irrigation management strategy
is therefore essential for high-ecient use of limited
irrigation water, particularly for AI, which accounts for
more than one-third of the total diversion of irrigation
water each year (Lu et al. 2019).
3. Materials and methods
The owchart (Figure 2) shows that the methodology
for this study can be divided into four major steps: (a)
data collection and pre-processing, (b) feature extrac-
tion, (c) development of the algorithm for retrieving
irrigation information (extent and time), and (d) accu-
racy assessment.
1600 X. QIAN ET AL.
3.1. Data sources
3.1.1 Remote sensing data and preprocessing
Remote sensing data used in this study include the
daily land surface reectance of the MODIS
(MOD09GA) from 2010 to 2020 and Sentinel-2 from
2016 to 2020.
The MOD09GA has 500 m spatial and daily tem-
poral resolutions, which can be downloaded from
https://ladsweb.modaps.eosdis.nasa.gov/. Sentinel-2
has a spatial resolution of 10–60 m and a temporal
resolution of 2–5 days. We used Sentinel-2ʹs Top of
Atmosphere (TOA) reectance product processed
Figure 1. Sketch of the Hetao Irrigation District and field sampling points. WLBH, JFZ, YJ, YC, and WLT refer to Wulanbuhe, Jiefangzha,
Yongji, Yichang, and Wulate sub-irrigation districts, respectively; DK, HH, LH, WY, and WLTQ refer to Dengkou, Hangjinhouqi, Linhe,
Wuyuan, and Wulateqianqi counties, respectively; AI and NAI refer to Autumn irrigation and Non-Autumn irrigation, respectively.
Figure 2. Flowchart of the proposed methodology. MODIS, MBWI, and MMAVE represent Moderate Resolution Imaging
Spectroradiometer, Multi-Band Water Index, and maximum value of the three-day moving average, respectively.
GISCIENCE & REMOTE SENSING 1601
using GEE because the surface reectance data is
available for a very limited time on GEE. Several stu-
dies have reported a promising accuracy for TOA data
in detecting surface water inundation (Seaton, Dube,
and Mazvimavi 2020). The remote sensing data pre-
processing included re-projection, clipping to the
study area, and cloud and snow removal. For
MOD09GA, the daily State Quality Assurance (QA)
ags were used to mask pixels aected by cloud and
snow. On the other hand, the QA60 bitmask band was
used for Sentinel-2 to remove the cloud.
We also evaluated the potential of another MODIS
product, MCD43A4, as a data source for AI identica-
tion. The MCD43A4 product corrects the problem of
MOD09GA being aected by sensor viewing angle
and solar zenith angle (Thompson and Paull 2017).
However, it has a higher percentage of missing data in
areas with high precipitation, aerosol concentration,
or snow cover (Colditz et al. 2018).
The study period of September 15 to November 30
of each year was selected based on the historical
statistics of starting and ending dates of AI. The com-
parison of the proportions of valid observations of the
above-mentioned three products in the study period
(given in Appendix A1, Figure 13) shows that
MOD09GA is more suitable for AI identication,
while Sentinel-2 can be used as a supplement to the
validation dataset.
3.1.2 Field sampling data
For validation purposes, sampling points were col-
lected in two eld surveys from 19th to
24 October 2019 and 21st to 25 October 2020. The
surveyed elds were divided into irrigation elds and
non-irrigation elds. The information of each sam-
pling eld including the longitude and latitude of
points along the eld boundary and azimuth is
recorded in detail. The longitude and latitude are
recorded with a Juno SA Handheld Computer with
GNSS (Global Navigation Satellite System) (Trimble,
USA). The eld-scale sampling data were then aggre-
gated within a spatial window of 500 m to determine
whether the corresponding MODIS pixel has been
irrigated. If more than 50% of the pixel area is irri-
gated, it is taken as an AI sample. If the irrigated area
of the pixel is less than 50% for the whole irrigation
period, then it is considered an NAI sample (Hao et al.
2022; Tsela et al. 2014). The numbers of irrigated and
non-irrigated pixels extracted from the eld survey in
2019 and 2020 is given in Table 1.
3.1.3 Other auxiliary data
Land-use maps of the study area (with 30 m spatial
resolution) for the years 2010 (Figure 3a) and 2020
(Figure 3b) are obtained from GlobeLand30 data sets
(http://globallandcover.com/). Globeland30 is
a global land cover product developed by the
Ministry of Natural Resources of the People’s
Republic of China. The overall accuracies (Kappa coef-
cients) of the GlobeLand30 datasets in 2010 and
2020 are 83.50% (0.78) and 85.72% (0.82), respectively
(http://globallandcover.com/). In addition, the classi-
cation accuracy of cropland was 92% based on the
data collected from an agricultural sample survey in
HID in 2020, where the remaining 8% of the cropland
is mainly classied as grassland.
The soil texture data (Figure 3d) obtained from the
Harmonized World Soil Database within the territory
of China in 1:1,000,000, was provided by the Cold and
Arid Regions Sciences Data Center at Lanzhou, China
(http://westdc.westgis.ac.cn/).
Statistical data of the AI area of each sub-irrigation
district from 2010 to 2020 (where the data for WLBH
and WLT are temporarily unavailable from 2019 to
2020) were provided by the Bayannur Institute of
Water Conservancy Research. Furthermore, the statis-
tical data of the AI area in the WLT after 2014 was not
used for model calibration and validation, as the data
for this period included only the Yellow River irriga-
tion area and was incongruent with other areas.
3.2. Masking of potentially irrigated area
In order to be consistent with MODIS data, the land-
use maps (Figures 3(a and 3b) were aggregated to
500-m pixel size using the nearest-neighbor
approach. All the croplands and grasslands north of
Table 1. The number of validation pixels extracted from field
sampling and Sentinel-2 images from 2016 to 2020.
Source Class
Year
2016 2017 2018 2019 2020 total
Field sampling AIL 56 117 173
NAIL 16 49 65
Sentinel-2 AIL 342 490 883 497 1577 3789
NAIL 464 484 510 725 511 2694
Note: AIL and NAIL represent the Autumn irrigation land and Non-Autumn
irrigation land, respectively.
1602 X. QIAN ET AL.
the ood bank in 2010 and 2020 land-use maps were
used as a mask of potentially irrigated areas for
further irrigation events detection. This is because
croplands and grasslands are often confused in land
cover classication (Shazadeh-Moghadam et al.
2021) and they are often converted to each other in
HID. Furthermore, croplands between the Yellow
River and the ood bank (Figure 1) generally do not
irrigate in autumn. However, these croplands exhibit
similar spectral characteristics to irrigated croplands
as a result of shallow groundwater levels and high soil
moisture due to the Yellow River and General Main
Canal close to the ood bank. Therefore, these crop-
lands were not considered to avoid misjudgment of
the AI location.
3.3. Method to extract irrigation information
3.3.1. Reconstruction of WI time series and WI
selection
A daily continuous time series of each index was rst
generated by linear interpolation based on cloudless
observation data and then ltered using an adaptive
weighted Savitzky-Golay (S-G) lter, a noise reduction
process proposed by Hu et al. (2020). The lter origin-
ally proposed for the MODIS-based Normalized
Dierence Vegetation Index (NDVI) is applied to the
MOD09GA. This lter eliminates the eect of negative
bias and improves the quality of the index time series.
In the ltering process, weights of indices on dierent
days were determined from the quality control para-
meters of the MOD09GA products. The four-bit range
of each reectance band provided by the QC band
has ve situations in the study area, i.e. 0000, 1000,
1011, 1101, and 1110. According to MODIS Surface
Reectance User’s Guide Collection 6 (Vermote,
Roger, and Ray 2015), they represent dierent quality
control aspects. In short, 0000 and 1000 are of high
quality; 1101 are of average quality; 1011 and 1110
are of low quality. Combined with the cloud and snow
conditions provided by the State band, the weight
(ω
i
) corresponding to the i-th point, is specied in
Table 2.
Initially, the lter window size is set to 7. In order to
reduce the computation complexity and reserve the
original high-quality data, the index values for points
with weight ω
i
= 1 remain unchanged in the tting
process. The lter window will expand adaptively
depending on the number of valid values (values
Table 2. Weights for the adaptive weighted Savitzky-Golay filter
determined from quality flags of the MOD09GA product.
Quality
band
name Flag Data quality description Weights
QC 500 m Bands 17
data
quality
Quality of all used bands is high 1.0
Quality of all used bands is above
average but not all high
0.8
Quality of all used bands is poor 0.2
Value obtained by linear
interpolation
0.1
The others 0.4
Figure 3. The Hetao Irrigation District maps: (a) Land-use map in 2010; (b) Land-use map in 2020; (c) Land-use changes from 2010 to
2020; (d) soil texture map.
GISCIENCE & REMOTE SENSING 1603
with weights greater than 0.1). The maximum lter
window size should not exceed 15, because data
beyond seven days before and after the current date
has little inuence on the central value.
We selected seven indices that could potentially be
used for AI identication, including Normalize
Dierence Water Index (NDWI) using bands 4 and 6
(NDWI46) and bands 1 and 7 (NDWI17) (Boschetti
et al. 2014), Multi-Band Water Index (MBWI) (X.
Wang et al. 2018), the 2015 water index (WI2015)
(Fisher, Flood, and Danaher 2016), Automated Water
Extraction Index for images with shadows (AWEIsh)
and without shadows (AWEInsh) (Feyisa et al. 2014),
and Soil Moisture Monitoring of Remote Sensing
index (SMMRS) (Zhan et al. 2006). We compared the
sensitivity of these indices to changes in water surface
and ice by analyzing all the eld samples from 2019
and 2020. Overall, MBWI is the most suitable index for
AI identication (see Appendix A2 for more details).
3.3.2. Selection of classification features
The feature variable derived from the MBWI curve is
selected based on the following two hypotheses: 1)
the peak MBWI value of irrigated land is higher than
that of non-irrigated land in the same sub-irrigation
district; 2) the uctuation of MBWI caused by rainfall/
snowfall, ice, and other factors during the study per-
iod is unlikely to cause a larger MBWI increase than AI
under normal circumstances.
However, we did not directly nd the maximum
MBWI value in the ltered MBWI time series in some
cases, which is mainly due to two reasons. First, some
of the peaks are not eective even after ltering the
MBWI because some uctuations may be caused by
external disturbances, e.g. subpixel clouds and incor-
rect ags in MODIS quality information (Y. Chen et al.
2018). Second, since the weighted S-G lter was
mainly employed to improve the quality of the
index time series without considering the smoothing
of curves, the ltered series may still have signicant
uctuations, making it dicult to extract the temporal
information of AI. The water surface in HID after AI can
be generally kept for 3–14 days depending mainly on
the applied irrigation depth and soil texture.
Therefore, a pixel having surface water for more
than three consecutive days should be identied as
an AI pixel. We further calculated the three-day mov-
ing average of the ltered MBWI time series, which
can make the time series smoother with a more
obvious changing trend, and thus help to determine
the AI time more precisely. Finally, we selected the
maximum value of the three-day moving average of
the ltered MBWI time series (MMAVE-MBWI) as the
feature variable.
3.3.3. Algorithm for identifying autumn irrigation
land
The key point of the threshold-based method is to
determine a threshold that can be used for multiple
years. The variation of reectance values of MOD09GA
products with sensor view-angle and solar zenith angle
and the application of S-G lter on MBWI time series
may lead to inconsistent inter-class separability across
years (Shao et al. 2016). Therefore, processing the data
is necessary to stabilize the thresholds. Several studies
have shown that the Z-score normalization method
can be used to nd outliers within the year, where
the resulting maps developed from Z-score normaliza-
tion are better than those derived from simple empiri-
cal threshold methods (B. Chen, Jin, and Brown 2019;
DeVries et al. 2020). Therefore, we used the Z-score
normalization method to eliminate the abnormal inter-
annual uctuations in each sub-irrigation district. The
Z-score can be calculated using the following equation:
Z¼aμ
δ(1)
where a and Z are the initial and normalized values of
the selected feature; μ and δ are the average and
standard derivation of the feature variable for
a reference land-use type in the study region. As an
objective baseline, the reference land-use type should
be able to reect inter-annual variations in feature
values due to factors unrelated to the study objectives,
and it should be available for each sub-irrigation dis-
trict. We compared three possible reference land-use
types, i.e. water bodies, all the croplands and grass-
lands, and part of croplands and grasslands with higher
MMAVE-MBWI, to nd an appropriate reference land-
use type (see Appendix A3 for more details). After
Z-score normalization, the threshold value of the fea-
ture variable (MBWI
Te
) was calibrated using the statis-
tical data for AI areas of each sub-irrigation district.
3.3.4. Algorithm for determining autumn irrigation
time
As a result of fragmented croplands owned by dier-
ent farmers, a cropland pixel of 500 m by 500 m would
1604 X. QIAN ET AL.
usually contain several pieces of cropland that may or
may not be irrigated at the same time. More and more
cropland in one pixel gets inundated with the appli-
cation of irrigation, which allows the coarse-
resolution sensors to start detecting cropland with
water cover. The change of MBWI in an AI pixel is
comprised of the following stages (Figure 4).
First, before the application of AI, a cropland pixel
changes from crop cover to bare soil immediately after
crop harvest. MBWI curve shows a valley point shortly
before AI (Point A in Figure 4) because the MBWI values
for cropland are generally higher than those of bare soil
(X. Wang et al. 2018). The MBWI at point A increases
when the cropland in the pixel is irrigated. After reach-
ing a specied threshold (Point B in Figure 4), a certain
percentage of the cropland in the pixel is irrigated and
thus the entire pixel can be considered to reach the
level of AI. Therefore, the irrigation time is determined
as the date of point B. The specied threshold (MBWI
T
)
corresponding to B is obtained by inverse normaliza-
tion using the calibrated MBWI
Te
, which is
MBWIT¼MBWITeδþμ(2)
MBWI rises continuously with AI until the irrigation
stops. At this specic point, the MBWI reaches its peak
(Point C in Figure 4) and the water depth of the entire
pixel reaches its maximum level. After that, MBWI
begins to decrease and reaches the threshold for the
last time (Point D in Figure 4). At this specic thresh-
old, it can be considered that the whole pixel is no
longer inundated, i.e. no longer in the AI state. MBWI
continues to decline and reaches the rst valley point
(Point E in Figure 4).
According to the characteristic points A, B, C, D,
and E of the MBWI time series, characteristic periods
can be dened, including peak range and initial, mid-
dle, and late irrigation stages. Peak range (points A to
E in Figure 4) is dened as the time range to the peak
between two valleys. If the ending valley is greater
than the threshold, the time range should extend to
the next valley below the threshold. Initial irrigation
stage (points A to B in Figure 4) illustrates that the
cropland in a pixel starts to be irrigated until the pixel
is in the state of AI with a certain proportion of
irrigated cropland. Middle irrigation stage (points B
to C in Figure 4) is the period when the water content/
depth of the entire pixel keeps increasing with con-
tinued irrigation until it reaches the peak and the
irrigation stops. Late irrigation stage (points C to
D in Figure 4) is the period when the whole pixel is
still inundated although irrigation has stopped, i.e.
the whole pixel is still in the AI state. Overall, the
whole irrigation period (points A to D in Figure 4)
combines the initial, middle, and late irrigation stages.
In general, the MBWI time series of a pixel may
have one or more peaks exceeding the threshold.
The following steps (Figure 5) are used to determine
the date of AI for dierent cases.
(1) For pixels with only one peak exceeding the
threshold (Figure 4a), the rst day when the
MBWI reaches the threshold is taken as the AI
date.
(2) For pixels having multiple peaks that exceed
the threshold and the highest peak appears
rst (Figure 4b), the rst day when the MBWI
Figure 4. Illustration of MBWI curves for three representative AI pixels and irrigation stages. Point A corresponds to The
Valley of the smoothed MBWI curve before autumn irrigation, Points B and D correspond to the threshold value in the
increasing and decreasing segments of the MBWI curve after autumn irrigation, Point C corresponds to the peak value of
smoothed MBWI curve, and Point E corresponds to the first valley of the smoothed MBWI curve after autumn irrigation. DOY
represents Day of the Year.
GISCIENCE & REMOTE SENSING 1605
reaches the threshold is taken as the AI date
because the subsequent peak(s) may be caused
by icing or other factors.
(3) For pixels having multiple peaks that exceed
the threshold where the highest peak did not
appear rst (Figure 4c), we developed
a “proximity algorithm” based on the assump-
tion that the AI dates of adjacent pixels should
be close or roughly the same. The proximity
algorithm is based on a 3 by 3 window centered
on this pixel with the following steps:
(a) If more than half of the AI dates of other pixels
in a 3 by 3 window are within a certain peak time
range of the central pixel (points A to E in Figure 5),
the AI dates in this 3 by 3 window are considered to
be consistent. In this case, the AI date of the central
pixel is the date when the MBWI rst reaches the
threshold within this consistent peak range.
(b) If the conditions in step (a) are not met, the AI
date is taken as the date when MBWI rst reaches the
threshold within the peak time range where the max-
imum rising segment of the MBWI curve intersects.
(c) Step (a) is performed iteratively until the irriga-
tion dates cannot be determined for more pixels.
(d) The irrigation date of the nal remaining points
is determined as the date when MBWI rst reaches the
threshold within the peak range of the maximum
peak.
3.3.5. Sensitivity analysis of the threshold
Sensitivity analysis of the threshold is performed
by taking the calibrated threshold in each sub-
irrigation district as the baseline (T
b
) and the
range of variation is set from −0.3 to 0.3. Then
the AI areas in the sub-irrigation districts are esti-
mated with dierent thresholds, and the sensitiv-
ity of the threshold (S
T
) can be calculated from
Wen et al. (2022)
ST¼ΔA
Ab
=ΔT
Tb
(3)
where S
T
represents the ratio of the relative change in
model output to the relative change in the threshold,
A
b
is the estimated AI area corresponding to T
b
, ΔA is
the variation of the simulated AI area, and
Figure 5. The hierarchical structure of the decision tree classifier for irrigation timing. AI and FDRT represent autumn irrigation and the
first day when the MBWI reaches the threshold.
1606 X. QIAN ET AL.
ΔT represents the change of the threshold. Generally,
a greater |S
T
| indicate the threshold is more sensitive.
3.4. Generating validation points from Sentinel-2
images
AI distribution maps are also veried using validation
points extracted from Sentinel-2 . The Sentinel-2
MBWI distribution maps from 2016 to 2020 were
generated using the GEE platform to produce refer-
ence AI points. MBWI was originally proposed based
on the Landsat 8 surface reectance product, where
the zero value represents the default threshold for
separating water from non-water bodies (X. Wang
et al. 2018). Since the water depth and surface area
of AI cropland are shallower and smaller than those of
water bodies, it can be assumed that the MBWI
threshold of water bodies is greater than or equal to
that of AI cropland. Therefore, the threshold of water
bodies is selected as the threshold of AI cropland to
ensure the accuracy of the validation sample for AI
lands. Since the TOA reectance product of Sentinel-2
is used, the default threshold from Landsat 8 surface
reectance product has to be redened by tting the
linear relationship between Landsat-based MBWI
(LMBWI) and Sentinel-2-based MBWI (SMBWI) with
images on the day when data is available from both
datasets. AI date of the corresponding pixel is the
date when the MBWI rst exceeds the threshold to
become a water-like body. Further, to comply with
the MODIS-derived results, the percentage of irri-
gated Sentinel-2 pixels to the total number of
Sentinel-2 pixels within each MODIS pixel (500 m)
was recorded.
A pixel with a percentage of AI over 50% is
extracted as an irrigated pixel (Hao et al. 2022) and
its irrigation date is recorded as the time when more
than 50% of the pixel is irrigated. However, the
threshold method is no longer applicable for the
selection of non-irrigated sample points. The multiple
snapshots of landscape from Google Earth, Sentinel-2,
and Landsat8 imageries are visually interpreted to
supplement the eld sampling data.
3.5. Calibration and validation of the model
Available studies have shown that the choice of
calibration and validation period has a signicant
impact on the evaluation of model performance
and further simulations (e.g. Myers et al. 2021).
Instead of arbitrarily choosing calibration and vali-
dation periods, we divide the dataset into training
and testing sets. The training dataset includes infor-
mation about the AI area of ve sub-irrigation dis-
tricts from 2010 to 2017. The testing dataset
includes recorded AI area data from 2018 to 2020,
eld sampling points from 2019 to 2020, and vali-
dation points extracted from Sentinel-2 from 2016
to 2020. A k-fold cross-validation method is used to
determine the appropriate threshold and evaluate
the performance of the model by iteratively with-
holding samples from the training dataset (Y. Chen
et al. 2018). Moreover, the testing dataset is used to
evaluate the accuracy of the model in greater detail.
3.5.1. Cross-validation in each sub-irrigation district
The thresholds in each sub-irrigation district are cali-
brated by minimizing the mean absolute relative error
(MARE) between simulated and observed annual
autumn irrigated areas, which is
MARE ¼1
nX
n
i¼1
^
yiyi
yi
(4)
where ^
yi and yi are the simulated and observed areas
at the year i, and n is the total number of years.
In order to select the threshold, we used k-fold
cross-validation, which involves training the model
using k−1 folds of training data and validating the
trained model using the remaining data. The data of
8 years acquired from 2010 to 2017 were divided into
four folds (k = 4), where data of 6 years is used to train
the model, and the remaining 2 years’ records are
used for model validation. However, the data from
2010 to 2014 was used for calibration and validation
in WLT, where data for three years was selected ran-
domly for calibration and the other two years for
validation. Consequently, a total of 28 models are
obtained for WLBH, JFZ, YJ, and YC, while 10 models
are for WLT after the cross-validation. The model with
the lowest MARE for each sub-irrigation district was
selected to further evaluate the accuracy.
3.5.2. Validation with testing dataset in the whole
HID
In the testing dataset, the irrigation extent is vali-
dated using statistical AI area data. Further, both
the irrigation extent and time are validated using
GISCIENCE & REMOTE SENSING 1607
eld sampling points and points extracted from
Sentinel-2.
Relative error (RE) and overall accuracy (OA) are
used to assess the accuracy of AI area and specic
locations, respectively, which are expressed as
REi¼^
yiyi
yi
(5)
OA ¼ ðNm=NÞ 100%(6)
where N
m
is the number of simulation objects match-
ing the observation object, and N is the number of the
observation objects.
OA is also used to assess the accuracy of the
determined AI date, where two dierent OAs are
used based on dierent irrigation periods. For the
middle to late period of irrigation (MTLP-based
OA), N
m
in the OA is the number of samples
whose observed irrigation dates are within the
date intervals of the Mid-to-late irrigation period
(points B to D in Figure 4). However, N
m
in OA
based on the whole period of irrigation (WP-based
OA) is the number of samples whose observed
dates are within the date intervals of the whole
irrigation period (points A to D in Figure 4).
WP-based OA is used in the validation of eld
sampling points because the recorded date may be
in any stage of AI for a pixel. For validation points
extracted from Sentinel-2, the recorded date is the
date on which most of the areas (50% or more) in
the pixel have been irrigated. Although the thresh-
old is high, it may also be lower than the value of
MBWI when the cropland has just begun to be
irrigated. Due to the revisit cycle and cloud cover,
Sentinel-2 may not be able to capture dates with
maximum MBWI value for many pixels. Therefore,
MTLP-based OA is more reasonable for points
extracted from Sentinel-2.
4. Results
4.1. Cross-validation results based on statistical
autumn irrigation areas
Based on the results of the inter-annual uctuations of
average feature variable (MMAVE-MBWI) for three dif-
ferent reference land-use types, i.e. water bodies, all
the croplands and grasslands, and part of croplands
and grasslands with higher MMAVE-MBWI, there is
a good agreement among the three land-use types
from 2010 to 2020 (Figure 15, Appendix A3). Since
most of the study area is covered by cropland and
grassland , we selected the average MMAVE-MBWI of
all the cropland and grassland pixels which is accurate
enough to reect the inter-annual changes of
MMAVE-MBWI for the study area.
After MBWI normalization with reference to the
average MMAVE-MBWI of all the cropland-grassland
pixels, 4-fold cross-validation was used to calibrate
and validate the model. It is observed the model per-
forms well when data from dierent periods was used
for calibration. The average MAREs during the calibra-
tion period of all models developed in the cross-
validation processes in each sub-irrigation district
are smaller and range from 2.7% to 7.6% (Figure 6,
left panel). The comparison shows that dierent cali-
bration periods results in slight dierences in the
MAREs . The dierence between maximum and mini-
mum MAREs in JFZ is only 2%, and the dierences in
other sub-irrigation districts are about 5%.
However, the ranges of MAREs in the validation
period are signicantly greater than those in the cali-
bration period (Figure 6, middle panel). These dier-
ences in MAREs might be associated with fewer years
(only 2 years) for validation. When the calibration and
validation periods were considered as a whole, the
variations in MAREs in each sub-irrigation district
becomes smaller and almost concentrated near the
mean or median (Figure 6, right panel). The MARE of
JFZ is the smallest with a mean value of 2.8%, while
WLBH, YJ, YC, and WLT have mean MAREs close to
each other ranging from 6% to 8%. These results
suggest that the threshold-dependent classier was
quite robust and produced reasonably good results.
Among the models developed for each sub-
irrigation district, we selected the model with mini-
mum MARE during the whole calibration and valida-
tion period as the nal model for further validation of
AI location and calculation of AI time. The nal MBWIT
and MBWITe of ve sub-irrigation districts are shown
in Table 3. The minimum and maximum MBWI
T
are
observed in WLBH and WLT, respectively. This is
mainly because WLBH has more sand while WLT has
more clay (Figure 3(d), resulting in dierent surface
water retention capacities.
The results of sensitivity analysis illustrated that the
values of S
T
are −0.22, 0.04, 0.21, −0.13, and −0.37 for
WLBH, JFZ, YJ, YC, and WLT, respectively. Since the
1608 X. QIAN ET AL.
autumn irrigated area decreases with the threshold,
the signs of S
T
values are opposite to those of MBWI
Te
(Table 3). In other words, a 100% increase in the
absolute values of the threshold results in the
decrease of identied AI area by 22%, 4%, 21%, 13%,
and 37%. In general, the models for dierent sub-
irrigation districts have a low to moderate sensitivity
to the threshold. Among the irrigation districts, JFZ
has the lowest sensitivity, probably due to its higher
percentages of AI in all years.
4.2. Model accuracy assessment based on testing
datasets
4.2.1. Validation points generated from Sentinel-2
images
The tting results between SMBWI and LMBWI are
SMBWI = 0.88*LMBWI−0.015, R
2
= 0.83, p < 0.05.
Therefore, we set the threshold value of SMBWI as
−0.015. If the cropland pixel’s MBWI value is greater
than −0.015 on a given day, the pixel is then identied
as an AI pixel. A total of 6483 testing samples (3789
irrigated and 2694 non-irrigated) from 2016 to 2020
were selected from Sentinel-2 images to further verify
the accuracy of the model (Table 1). The distribution
of validation points is shown in Appendix A4
(Figure 16).
4.2.2. Accuracy assessment of autumn irrigation area
and its distribution
The model performed well in the testing years, as
illustrated by smaller relative errors between
observed and simulated irrigation areas at WLBH in
2018 and JFZ, YJ, and YC from 2018 to 2020 (Table 4).
The relative errors are within the acceptable range,
among which 52% are in the range of −5.0% to 5.0%
and 81% from −10.0% to 10.0%. The MAREs for the
testing year (2018–2020) are 2.2%, 7.3%, 6.0%, and
4.5% for WLBH, JFZ, YJ, and YC, respectively. The
MAREs of sub-irrigation districts and the whole HID
from 2010 to 2020 are between 4% and 7%.
Therefore, the threshold-based model developed
using data from 2010 to 2017 can be applied to
identify AI in other years with satisfactory precision.
Moreover, the overall accuracy for MMAVE-MBWI-
based classication model was 90.0% and 97.0%
according to the eld sampling data and the refer-
ence data extracted from Sentinel-2, respectively
(Table 5). In summary, the good performance of
Figure 6. Box plot for MAREs of each sub-irrigation district for calibration period (left), validation period (middle), and the whole
calibration and validation periods (right).
Table 3. Threshold of each sub-irrigation district from 2010 to
2020.
Sub-irrigation district Range of MBWI
T
MBWI
Te
WLBH −0.53 to −0.46 0.26
JFZ −0.41 to −0.33 −0.08
YJ −0.51 to −0.40 −0.44
YC −0.40 to −0.30 0.24
WLT −0.37 to −0.28 0.51
Note: MBWI
T
and MBWI
Te
represent the threshold and the normalized
threshold, respectively.
GISCIENCE & REMOTE SENSING 1609
model validations with independent datasets sup-
ports the robustness and potential of the method
for extracting AI attributes.
4.2.3. Accuracy assessment of autumn irrigation
timing
The accuracy of estimating irrigation time is given in
Table 6. For eld sampling points, WP-based OA is
higher than 90.0% in both 2019 and 2020, with an
average of 91.7%., However, the MTLP-based OA for
points extracted from Sentinel-2 varies from 66.5% to
82.9% with an average of 76.4% when 50% of irri-
gated area is taken as the lower limit of the irrigated
fraction to dene an irrigated MODIS pixel. WP-based
OA has higher values than those from MTLP-based OA
because of the longer time interval.
Furthermore, the MTLP-based OA from 2017 to 2020
can be signicantly higher than in 2016. This may be
related not only to the identication results but also to
the reference data used for assessment. The pixels in
study area have average Sentinel-2 images of 7.4, 12.2,
13.6, 13.5, and 12.7 days during the study period from
2016 to 2020. A signicantly smaller number of images
in 2016 may cause a loss of eective information, thus
failing to provide a more realistic assessment of the
model quality. Therefore, we attempted to increase the
eective information by increasing the lower limit
(80%) for the irrigation fraction of the validated pixels
to account for the dierence in MTLP-based OA in
dierent years. It can be found that the average MTLP-
based OA will rise to 88.9%, with a signicant increase
in 2016 in particular (Table 6).
Overall, the observed AI time matches well with the
irrigation dates simulated by our model, which indicates
the method used to calculate irrigation time is reliable.
4.3. Spatial-temporal changes in the distribution of
irrigated croplands across HID
Figure 7 depicts the spatial distributionof AI areas in
HID from 2010 to 2020, where the irrigated area
depends mainly on the availability of runo from
Table 4. Relative error (%) of each sub-irrigation district and the whole HID from 2010 to 2020.
Year WLBH JFZ YJ YC WLT HID
2010 −4.1 −6.2 −3.8 −17.0 −8.7 −9.4
2011 0.2 −2.0 −3.7 −7.8 −14.1 −5.5
2012 −6.6 0.1 −0.5 −2.3 0.4 −1.5
2013 0.2 2.7 0.2 8.5 −0.5 3.1
2014 13.2 6.7 16.4 13.6 6.5 12.2
2015 16.1 −2.7 7.0 −8.2
2016 −12.0 1.2 −3.5 0.2
2017 9.9 0.2 9.3 0.3
2018 −2.2 16.2 13.3 6.3
2019 5.5 2.7 −2.6
2020 −0.1 −1.9 −4.7
MARE (%) 7.2 4.0 5.7 6.5 6.0 6.3
Note: MARE refers to mean absolute relative error. Underlined numbers are for the testing dataset.
Table 5. The overall accuracy of irrigated land mapping using reference data.
Data source Identified class
Actual class
AIL NAIL Total Correct ratio
Field sampling data AIL 154 5 159 96.9%
NAIL 19 60 79 75.9%
Total 173 65 238 OA = 90.0%
Sentinel-2 reference data AIL 3625 30 3655 99.2%
NAIL 164 2664 2828 94.2%
Total 3789 2694 6483 OA = 97.0%
Note: AIL and NAIL represent Autumn irrigation land and Non-Autumn irrigation land, respectively.
Table 6. Overall accuracy (OA) of irrigation timing from 2016 to 2020.
Source OA
Year
2016 2017 2018 2019 2020 Average
Field sampling WP-based 92.1% 91.3% 91.7%
Sentinel-2(50%) MTLP-based 66.5% 79.1% 82.9% 81.1% 72.4% 76.4%
Sentinel-2(80%) MTLP-based 100.0% 75.8% 96.7% 100.0% 86.5% 88.9%
Note: 50% and 80% are lower limits of the irrigated fractions to define an irrigated MODIS pixel.
1610 X. QIAN ET AL.
the Yellow River that varied by as much as 20%
among dierent years. Over the 11 hydrological
years, the irrigated area varied between a maximum
of 4758 km
2
in 2011 and a minimum of 4121.5 km
2
in
2019. The mean autumn irrigated area from 2010 to
2020 was 4444.3 km
2
with a standard deviation of
215.25 km
2
, providing insights into the inter-annual
variability in the extent of the autumn irrigated area.
The proportion of the total AI area in each sub-
irrigation district has remained stable for these years,
with average values of 10.54% for WLBH, 26.70% for
JFZ, 21.99% for YJ, 29.27% for YC, and 11.48% for WLT.
Although there have been slight uctuations in some
years, the irrigated area has shown a downward trend
during the past 11 years (Figure 8), with an average
decreasing rate of 53.76 km
2
/year in the whole HID.
Besides, YC and YJ also have higher decreasing rates
of 20.50 and 17.44 km
2
/year, respectively. Based on
the statistical data of planting areas acquired from
Water Conservancy Development Center of HID and
the estimated autumn irrigated areas in Figure 8, JFZ
has the highest proportion of AI within the total
planting area (83.5%), while WLBH has the lowest
proportion (58.1%). Despite the largest cropland
area in YC, its proportion of AI (70.8%) is considerably
lower than that of JFZ. Overall, the proportion of AI
has reached 72.0% for the whole HID.
Superimposing multi-year images of AI area can
eectively detect the maximum extent of irrigation
area and nd the autumn irrigated times for each
pixel from 2010 to 2020 (Figure 9 a). Some lands
have never been irrigated in autumn, mainly grass-
land, wasteland, and sunowers in WLBH and WLT.
Highly concentrated AI lands are mainly distributed in
the JFZ and YJ, the southern WLBH, eastern YC, and
northern WLT. The distribution of AI in the WLBH is
mainly concentrated near the main canal.
By aggregating the 500 m irrigation map into
a uniform 9 km
2
pixel and performing linear regression
on the irrigation area of each aggregated pixel, we
further studied the spatial trends in the variations of
irrigated area over time on a ne scale (Figure 9(b).
From Figures 7 and 9b, we can nd that the autumn
irrigated areas in the southern YC and the northern YJ
have signicantly decreased since 2014. The western
part of the WLT has hardly been irrigated in autumn
since 2016. The northeast corner of the JFZ has also
seen a reduction in AI area since 2015.
4.4. Spatial-temporal changes in autumn irrigation
timing
Figure 10 depicts the spatial distribution of AI dates
across HID from 2010 to 2020. The gure shows
Figure 7. Spatial distributions of autumn irrigated areas in the Hetao Irrigation District from 2010 (a) to 2020 (k).
GISCIENCE & REMOTE SENSING 1611
obvious dierences in the spatial distribution of AI
time among dierent years. In general, earlier irriga-
tion was found around the boundary between YC and
WLT and some areas in the north of HID. Compared
with highly variable irrigation time in areas with small
and scattered cropland, more synchronized irrigation
was found in the southern part of HID (mainly in JFZ
and YJ). The identied irrigation time of WLBH varies
greatly from year to year.
As an example, Figure 11 shows evolution process
of the irrigation area in 2019. A small-scale AI area rst
appeared in YJ and the east of YC. Before
October 15th, fewer areas in the HID have been irri-
gated, while a wide range of AI has appeared in JFZ,
YJ, and YC since October 15th. By November 15th, the
whole HID has almost completed AI. The results
(cumulative area diagram shown in Appendix A5,
Figure 17) show that irrigation was earlier in 2010,
2012, and 2018, while delayed in 2015 and 2020.
Figure 12 shows that the start time of AI is delayed
gradually from 2010 to 2020. Meanwhile, the propor-
tion of cropland irrigated from mid-September to
mid-October showed a downward trend and reached
a new low of 6% in 2020. Since 2012, more than 70%
of the cropland is irrigated from mid-October to mid-
November, and this proportion is still increasing.
5. Discussion
Our study found that AI in HID showed dierent
spatial distribution patterns each year, with
a downward trend in the area, and a delayed starting
time in the study period. This is consistent with our
Figure 9. Spatial distribution of autumn irrigation times (a) and variation trends in the irrigated area for aggregated pixels of 9 km
2
(b)
in the 11 years from 2010 to 2020.
Figure 8. Inter-annual changes of the autumn irrigation areas (S) in each sub-irrigation district and the whole HID from 2010 to 2020.
t is the year.
1612 X. QIAN ET AL.
Figure 10. Spatial distributions of autumn irrigation time in the Hetao Irrigation District for the period of 2010 (a) to 2020 (k).
Figure 11. The time transition diagram of the spatial distribution of autumn irrigation in 2019, divided into five periods with half-
month intervals: September 15th to September 30th (a), September 15th to October 15th (b), September 15th to October 31st (c),
September 15th to November 15th (d), and September 15th to November 30th (e).
GISCIENCE & REMOTE SENSING 1613
eld investigation and previous literature results (L.
Wang et al. 2021; Yang, Liu, and Liu 2017).
The distribution pattern of AI is inuenced by crop-
ping patterns and soil texture. The cropping patterns
dier in dierent sub-irrigation districts, especially the
percentage of sunower planting. The sunower is
more salt-tolerant and planted later than other
major crops, which can be irrigated in the spring
before sowing to leach salt and increase soil moisture.
Therefore, sunower has less demand for AI. As
a result, JFZ with a lower percentage of sunower
(Wen, Shang, and Rahman 2019) has the highest pro-
portion of AI within the total planting area. On con-
trary, YC and WLT with higher percentages of
sunower have lower proportions of AI.
Furthermore, WLBH has the lowest proportion of AI
mainly due to sandy soil texture (Figure 3(d). The
sandy soil with poor moisture retention capacity
results in poor eect in increasing the root zone
moisture, which reduces the farmers’ enthusiasm for
AI and leads to great variation in the AI time
(Figure 10) dependent on the farmers’ decisions.
The decreasing trend in the AI area (Figure 8) is
mainly related to the change in cropping patterns and
irrigation water fees. Increased area and extent of
sunower planting in recent years (Yu and Shang
2017) is the main reason for the decreasing trend of
AI area. The decreasing rate was greater in YC and YJ
and smaller in WLBH, JFZ, and WLT. This is mainly
because the southern part of YC and WLT has always
been the growing area of sunowers, but recently
sunowers are expanding toward the middle of HID,
i.e. the rest of YJ and YC. Besides, the water fee of AI is
higher than that of spring irrigation in HID, which
results in the decreased enthusiasm of the farmers
for AI and also leads to a decreasing trend in the AI
area.
The postponement of AI time is related to not only
cropping patterns but also some other factors such as
climate change. In recent years, the proportion of late
autumn crops tends to increase, such as maize har-
vested at the end of September, which leads to
delayed AI time in these croplands. Meanwhile, cli-
mate change characterized by temperature rising and
shortened freezing periods leads to the poorer eect
of early irrigation (Yang, Liu, and Liu 2017).
The changes in AI area and time signicantly
impact the hydrological cycle, agriculture, ecosystem,
and environment in HID. On the one hand, a decrease
in the AI area leads to a drop in groundwater levels.
Consequently, road slurry and land collapse caused
by frost heave and thaw collapse in mid-spring will
become less frequent (Yang, Liu, and Liu 2017). In
addition, although the area irrigated by water diver-
sion from the Yellow River is decreasing year by year
in WLT, the total AI area is still stable (Figure 8). It
indicates that the proportion of AI using other irriga-
tion methods is increased annually in WLT, which has
a good eect on saving irrigation water and control-
ling soil salinity (Mao et al. 2017). On the other hand,
the changes in irrigation extent and timing will lead to
corresponding changes in salinity and moisture
Figure 12. Proportions of irrigated areas from September 15th to October 15th and from October 15th to November 15th from 2010 to
2020 and their corresponding trend lines.
1614 X. QIAN ET AL.
distribution (Li et al. 2012). The multi-year water-salt
balance in the HID will be disrupted, requiring arti-
cial adjustment of irrigation strategies.
Our results can be used as a reference for irrigation
and agricultural management in the HID. First, it can
provide AI information in near-real-time (by the end
of AI season) to guide spring irrigation and AI in the
coming year. Second, the results provide information
on croplands that are not irrigated in autumn, which
are useful for determining the main planting areas for
sunowers and guiding the rational adjustment of the
cropping patterns to prevent the single planted vari-
ety from aecting the sustainable development of
local agricultural production (J. Liu et al. 2015).
The present model has been applied successfully in
the HID. However, because it is an initial exploration
in identifying the extent and timing of AI, the model
still has some limitations and needs to be further
improved.
To better detect the AI time, MODIS data with the
daily temporal resolution is used. However, the mod-
erate spatial resolution of MODIS data may inuence
the identication precision because of the inevitable
problem of mixed pixels (Wardlow and Callahan
2014). The proportion of AI in HID is high, exceeding
70% of the crop planting area. Furthermore, the irri-
gated land is basically in large blocks, which provides
a better opportunity to use MODIS data to identify the
irrigated area. Although the model performs well in
the entire HID and four of ve sub-irrigation districts,
its performance in WLBH is not as good as in other
districts. This is mainly because the distribution of
cropland in WLBH is more fragmented and MODIS
cannot capture it well. The selection of remote sen-
sing data with appropriate temporal and spatial reso-
lutions depends mainly on the distribution of
cropland in the irrigation district and research pur-
poses (Velpuri et al. 2009). Generally, products whose
spatial resolution matches the cropland scale for irri-
gation application should be used (Dari et al. 2022;
Massari et al. 2021). However, we often need data
with high temporal resolution (e.g. daily scale) that
may not have the high spatial resolution to match the
irrigation scale (Durgun et al. 2020). In such cases, the
percentage of cropland in each pixel can be consid-
ered in the classication in future studies or an appro-
priate method for irrigation land fraction estimation
in pixel scale can be further developed (W. Liu et al.
2018; Zhang, Chen, and Lu 2015).
Another source of uncertainty may be associated
with the threshold, which is also an unavoidable pro-
blem of the threshold-based method. The key is how
to choose the threshold value. In some previous stu-
dies using MODIS data for identifying irrigation, the
thresholds came from typical sample plots (L. Wang
et al. 2021) that require the matching between sam-
ple purity and actual purity of the classication target;
otherwise, it may lead to misclassication (Y. Chen
et al. 2016). Therefore, we chose the conservative
calibration method with statistical data to obtain the
threshold. Second, the thresholds are normally not
stable over dierent years. Compared with a xed
threshold, using a exible Z-score normalized thresh-
old can make the model robust enough to be applied
in dierent years (B. Chen, Jin, and Brown 2019).
However, we still need to be cautious in applying
the thresholds to changed environmental conditions
in the future. Besides, MBWI thresholds need to be
adjusted for other irrigation districts with dierent
management practices, irrigation volumes, soil char-
acteristics, and cropping patterns.
Inaccuracies in land-use products can also lead to
uncertainties in the identication of irrigated land.
Identifying possible cultivated land is the rst step in
mapping irrigated land (Jin et al. 2016). Usually, the
extraction of potential cultivated land depends on the
existing land cover maps (Xie et al. 2019). However,
the classied maps were generally for country or even
global scales, which is seldom independently vali-
dated for a specied irrigation district (Gumma et al.
2020). Moreover, the temporal resolution of these
products is relatively low, and the updating speed of
these products generally cannot keep up with the
speed of actual land-use change (Pan et al. 2020).
According to the land-use maps for 2010 and 2020
(Figure 3), the distribution of lakes in WLBH has chan-
ged signicantly from 2010 to 2020. Since lakes are
possibly confused with the phenomenon of AI, land-
use maps in only two years cannot eliminate the
inuences of lakes in other years. Therefore, precision
land-use maps for the study region on an annual basis
are benecial for high precision in AI mapping.
For accuracy assessment, the mixed pixel problem
complicates the evaluation of the nal mapping accu-
racy. A threshold of 50% for AI or NAI area proportion
was chosen to extract MODIS validation pixels based
on sampling points and Sentinel-2 images with ner
scales, which is a common practice in binary
GISCIENCE & REMOTE SENSING 1615
classication with mixed pixels (Hao et al. 2022).
However, the most appropriate proportions for dier-
ent regions need to be further explored (Colditz et al.
2018). In future studies, AI classication results con-
sidering the percentage of cropland in each pixel are
useful to study mapping accuracies at dierent levels
of cropland abundance, which can also provide useful
insights to nd the most appropriate proportion to
identify a MODIS pixel as an AI pixel in the study
region.
Moreover, the data gap of Sentinel-2 may aect the
evaluation results on irrigation time. Studies have
shown that the number of images involved is crucial
when detecting objects with spatial-temporal hydro-
logical uctuations (Borro et al. 2014). The principles
we set for determining the irrigation time for samples
may not be applicable to all samples due to data gaps.
Therefore, the optimal number of images for valida-
tion sample selection should also be considered in the
future, which may aid in achieving a realistic and
more accurate assessment of the model.
6. Conclusions
Knowing exactly when and where irrigation is applied
is essential for eective irrigation water management.
Autumn irrigation has been adopted in many regions
of the world but is usually less investigated than
irrigation during the growing season. A new approach
was developed to map AI area and timing distribu-
tions using MODIS MBWI time series in the Hetao
Irrigation District (HID) of Northwest China. This
study addressed three key technical issues, i.e. selec-
tion of an appropriate WI for the freezing period,
temporal incomparability of thresholds, and diculty
in validating the identied irrigation results, especially
irrigation timing. The model is simple but eective
and suitable for rapidly updating AI lands in large
irrigation districts, especially for irrigated cropland
distributed in large blocks. The generated maps
from 2010 to 2020 for HID with a total land area of
1.189 million ha and an irrigated land area of
0.733 million ha delivered reasonable performance.
The overall accuracy of irrigated area is over 90%,
whereas the overall accuracies of irrigated dates
assessed with “MTLP” and “WP” are 76.4% and
91.7%, respectively. The proposed mapping metho-
dology and associated datasets provide the rst
detailed AI extent and timing distribution dataset for
HID to date and can help reveal new insights into
irrigated land use and water use change.
Based on the mapping results of 11 years of
data, although the AI area has increased or
decreased in some regions, there is no signicant
dierence in the annual spatial distributions of AI
extent compared to the signicant dierence in
the annual spatial distribution of AI time in HID.
Moreover, the area of AI tends to decline in the
whole HID, and the starting time of AI is gradually
delayed. These changes are related to many fac-
tors, including climate, irrigation fees, and changes
in cropping patterns. The results obtained in this
study can provide a reference for the rational for-
mulation of AI scheduling under water-saving irri-
gation conditions and can provide a basis for
reducing secondary salinization and promoting
agricultural production in the study area.
Acknowledgements
We are grateful to Dr. Khalil Ur Rahman from Tsinghua
University for proofreading the manuscript. We are also grate-
ful to the editors and anonymous reviewers, whose comments
and suggestions are helpful to improve the quality of the
manuscript.
Disclosure statement
No potential conict of interest was reported by the author(s).
Funding
This work was supported by the National Natural Science
Foundation of China [52279038, 51839006, and 51779119];
Research Program of the State Key Laboratory of Hydroscience
and Engineering, Tsinghua University [2020-KY-01].
Data availability
The data that support the ndings of this study are available on
request from the corresponding author.
References
Bazzi, H., N. Baghdadi, D. Ienco, M. El Hajj, M. Zribi,
H. Belhouchette, M. J. Escorihuela, and V. Demarez. 2019.
“Mapping Irrigated Areas Using Sentinel-1 Time Series in
Catalonia, Spain.” Remote Sensing 11 (15): 1836. doi:10.
3390/rs11151836.
1616 X. QIAN ET AL.
Borro, M., N. Morandeira, M. Salvia, P. Minotti, P. Perna, and
P. Kandus. 2014. “Mapping Shallow Lakes in A Large South
American Floodplain: A Frequency Approach on
Multitemporal Landsat TM/ETM Data.” Journal of Hydrology
512: 39–52. doi:10.1016/j.jhydrol.2014.02.057.
Boschetti, M., F. Nutini, G. Manfron, P. Brivio, and A. Nelson.
2014. “Comparative Analysis of Normalised Dierence
Spectral Indices Derived from MODIS for Detecting Surface
Water in Flooded Rice Cropping Systems.” PLoS One 9 (2):
e88741. doi:10.1371/journal.pone.0088741.
Chen, B., Y. Jin, and P. Brown. 2019. “Automatic Mapping of
Planting Year for Tree Crops with Landsat Satellite Time
Series Stacks.” ISPRS Journal of Photogrammetry and
Remote Sensing 151: 176–188. doi:10.1016/j.isprsjprs.2019.
03.012.
Chen, Y., D. Lu, L. Luo, Y. Pokhrel, K. Deb, J. Huang, and Y. Ran.
2018. “Detecting Irrigation Extent, Frequency, and Timing in
a Heterogeneous Arid Agricultural Region Using MODIS
Time Series, Landsat Imagery, and Ancillary Data.” Remote
Sensing of Environment 204 (October 2017): 197–211. doi:10.
1016/j.rse.2017.10.030.
Chen, Y., X. Song, S. Wang, J. Huang, and L. R. Mansaray. 2016.
“Impacts of Spatial Heterogeneity on Crop Area Mapping in
Canada Using MODIS Data.” ISPRS Journal of
Photogrammetry and Remote Sensing 119: 451–461. doi:10.
1016/j.isprsjprs.2016.07.007.
Colditz, R. R., C. Troche Souza, B. Vazquez, A. J. Wickel, and
R. Ressl. 2018. “Analysis of Optimal Thresholds for
Identication of Open Water Using MODIS-Derived Spectral
Indices for Two Coastal Wetland Systems in Mexico.”
International Journal of Applied Earth Observation and
Geoinformation 70: 13–24. doi:10.1016/j.jag.2018.03.008.
Dao, P. D., N. T. Mong, and H.-P. Chan. 2019. “Landsat-MODIS
Image Fusion and Object-Based Image Analysis for
Observing Flood Inundation in a Heterogeneous Vegetated
Scene.” GIScience & Remote Sensing 56 (8): 1148–1169.
doi:10.1080/15481603.2019.1627062.
Dari, J., L. Brocca, P. Quintana-Seguí, S. Casadei,
M. J. Escorihuela, V. Stefan, and R. Morbidelli. 2022. “Double-
Scale Analysis on the Detectability of Irrigation Signals from
Remote Sensing Soil Moisture over an Area with Complex
Topography in Central Italy.” Advances in Water Resources
161: 104130. doi:10.1016/j.advwatres.2022.104130.
Dari, J., P. Quintana-Seguí, M. J. Escorihuela, V. Stefan, L. Brocca,
and R. Morbidelli. 2021. “Detecting and Mapping Irrigated
Areas in a Mediterranean Environment by Using Remote
Sensing Soil Moisture and a Land Surface Model.” Journal
of Hydrology 596: 126129. doi:10.1016/j.jhydrol.2021.
126129.
De Carvalho Júnior, O. A., R. F. Guimarães, C. R. Silva, and
R. A. T. Gomes. 2015. “Standardized Time-Series and
Interannual Phenological Deviation: New Techniques for
Burned-Area Detection Using Long-Term MODIS-NBR
Dataset.” Remote Sensing 7 (6): 6950–6985. doi:10.3390/
rs70606950.
Deines, J. M., A. D. Kendall, M. A. Crowley, J. Rapp, J. A. Cardille,
and D. W. Hyndman. 2019. “Mapping Three Decades of
Annual Irrigation across the US High Plains Aquifer Using
Landsat and Google Earth Engine.” Remote Sensing of
Environment 233: 111400. doi:10.1016/j.rse.2019.111400.
DeVries, B., C. Huang, J. Armston, W. Huang, J. W. Jones, and
M. W. Lang. 2020. “Rapid and Robust Monitoring of Flood
Events Using Sentinel-1 and Landsat Data on the Google
Earth Engine.” Remote Sensing of Environment 240: 111664.
doi:10.1016/j.rse.2020.111664.
Durgun, Y. Ö., A. Gobin, G. Duveiller, and B. Tychon. 2020.
“A Study on Trade-Os between Spatial Resolution and
Temporal Sampling Density for Wheat Yield Estimation
Using Both Thermal and Calendar Time.” International
Journal of Applied Earth Observation and Geoinformation
86: 101988. doi:10.1016/j.jag.2019.101988.
Elwan, E., M. Le Page, L. Jarlan, N. Baghdadi, L. Brocca,
S. Modanesi, J. Dari, P. Quintana Seguí, and M. Zribi. 2022.
“Irrigation Mapping on Two Contrasted Climatic Contexts
Using Sentinel-1 and Sentinel-2 Data.” Water 14 (5): 804.
doi:10.3390/w14050804.
Feyisa, G. L., H. Meilby, R. Fensholt, and S. R. Proud. 2014.
“Automated Water Extraction Index: A New Technique for
Surface Water Mapping Using Landsat Imagery.” Remote
Sensing of Environment 140: 23–35. doi:10.1016/j.rse.2013.
08.029.
Fisher, A., N. Flood, and T. Danaher. 2016. “Comparing Landsat
Water Index Methods for Automated Water Classication in
Eastern Australia.” Remote Sensing of Environment 175:
167–182. doi:10.1016/j.rse.2015.12.055.
Gao, Q., M. Zribi, M. J. Escorihuela, N. Baghdadi, and P. Q. Segui.
2018. “Irrigation Mapping Using Sentinel-1 Time Series at
Field Scale.” Remote Sensing 10 (9): 1495. doi:10.3390/
rs10091495.
Gumma, M. K., P. S. Thenkabail, P. G. Teluguntla, A. Oliphant,
J. Xiong, C. Giri, V. Pyla, S. Dixit, and A. M. Whitbread. 2020.
“Agricultural Cropland Extent and Areas of South Asia
Derived Using Landsat Satellite 30-m Time-Series Big-Data
Using Random Forest Machine Learning Algorithms on the
Google Earth Engine Cloud.” GIScience & Remote Sensing
57 (3): 302–322. doi:10.1080/15481603.2019.1690780.
Hao, X., G. Huang, Z. Zheng, X. Sun, W. Ji, H. Zhao, J. Wang, H. Li,
and X. Wang. 2022. “Development and Validation of a New
MODIS Snow-Cover-Extent Product over China.” Hydrology
and Earth System Sciences 26 (8): 1937–1952. doi:10.5194/
hess-26-1937-2022.
Hu, S. S., C. X. Huang, B. Yang, and Z. F. Tan. 2020.
“Reconstruction of MODIS Vegetation Index Time Series by
Adaptive Weighted Savitzky-Golay Filter.” Science of Surveying
and Mapping 45 (4) 105–116. In Chinese with English abstract.
Jeong, S., S. Kang, K. Jang, H. Lee, S. Hong, and D. Ko. 2012.
“Development of Variable Threshold Models for Detection of
Irrigated Paddy Rice Fields and Irrigation Timing in
Heterogeneous Land Cover.” Agricultural Water
Management 115: 83–91. doi:10.1016/j.agwat.2012.08.012.
GISCIENCE & REMOTE SENSING 1617
Jing, M., H. M. Zhang, J. Yang, Y. M. Fan, F. G. Huang, and
L. Chen. 2020. “Study of Deep Water Saving and Control in
Ning-Meng Irrigation Area.” Yellow River 42 (9) 155–160. In
Chinese with English abstract.
Jin, N., B. Tao, W. Ren, M. Feng, S. Rui, L. He, W. Zhuang, and
Q. Yu. 2016. “Mapping Irrigated and Rainfed Wheat Areas
Using Multi-Temporal Satellite Data.” Remote Sensing 8 (3):
207. doi:10.3390/rs8030207.
Ji, L., L. Zhang, and B. Wylie. 2009. “Analysis of Dynamic
Thresholds for the Normalized Dierence Water Index.”
Photogrammetric Engineering & Remote Sensing 75 (11):
1307–1317. doi:10.14358/PERS.75.11.1307.
Karthikeyan, L., I. Chawla, and A. K. Mishra. 2020. “A Review of
Remote Sensing Applications in Agriculture for Food
Security: Crop Growth and Yield, Irrigation, and Crop
Losses.” Journal of Hydrology 586 (March): 124905. doi:10.
1016/j.jhydrol.2020.124905.
Klein, A. G., and A. C. Barnett. 2003. “Validation of Daily MODIS
Snow Cover Maps of the Upper Rio Grande River Basin for
the 2000–2001 Snow Year.” Remote Sensing of Environment
86 (2): 162–176. doi:10.1016/S0034-4257(03)00097-X.
Kreyszig, E. 2010. Advanced Engineering Mathematics. Hoboken,
New Jersey: John Wiley & Sons.
Li, R., H. Shi, G. N. Flerchinger, T. Akae, and C. Wang. 2012.
“Simulation of Freezing and Thawing Soils in Inner Mongolia
Hetao Irrigation District, China.” Geoderma 173–174: 28–33.
doi:10.1016/j.geoderma.2012.01.009.
Liu, W., J. Dong, K. Xiang, S. Wang, W. Han, and W. Yuan. 2018.
“A Sub-Pixel Method for Estimating Planting Fraction of
Paddy Rice in Northeast China.” Remote Sensing of
Environment 205 (September 2017): 305–314. doi:10.1016/j.
rse.2017.12.001.
Liu, J., S. Sun, P. Wu, Y. Wang, and X. Zhao. 2015. “Evaluation of
Crop Production, Trade, and Consumption from the
Perspective of Water Resources: A Case Study of the Hetao
Irrigation District, China, for 1960–2010.” Science of the Total
Environment 505: 1174–1181. doi:10.1016/j.scitotenv.2014.
10.088.
Lu, X., R. Li, H. Shi, J. Liang, Q. Miao, and L. Fan. 2019.
“Successive Simulations of Soil Water-Heat-Salt Transport
in One Whole Year of Agriculture after Dierent Mulching
Treatments and Autumn Irrigation.” Geoderma 344: 99–107.
doi:10.1016/j.geoderma.2019.03.006.
Macfarlane, C., A. H. Grigg, and M. I. Daws. 2017.
“A Standardised Landsat Time Series (1973–2016) of Forest
Leaf Area Index Using Pseudoinvariant Features and
Spectral Vegetation Index Isolines and A Catchment
Hydrology Application.” Remote Sensing Applications:
Society and Environment 6 (January): 1–14. doi:10.1016/j.
rsase.2017.01.006.
Mao, W., J. Yang, Y. Zhu, M. Ye, and J. Wu. 2017. “Loosely
Coupled SaltMod for Simulating Groundwater and Salt
Dynamics under Well-Canal Conjunctive Irrigation in
Semi-Arid Areas.” Agricultural Water Management 192:
209–220. doi:10.1016/j.agwat.2017.07.012.
Massari, C., S. Modanesi, J. Dari, A. Gruber, G. J. M. De Lannoy,
M. Girotto, P. Quintana-Seguí, et al. 2021. “A Review of
Irrigation Information Retrievals from Space and Their
Utility for Users.” Remote Sensing 13 (20): 4112. doi:10.
3390/rs13204112.
Myers, D. T., D. L. Ficklin, S. M. Robeson, R. P. Neupane, A. Botero-
Acosta, and P. M. Avellaneda. 2021. “Choosing an Arbitrary
Calibration Period for Hydrologic Models: How Much Does It
Inuence Water Balance Simulations?” Hydrological Processes
35 (2): e14045. doi:10.1002/hyp.14045.
Negri, C., E. Chiaradia, M. Rienzner, A. Mayer, C. Gandol,
M. Romani, and A. Facchi. 2020. “On the Eects of Winter
Flooding on the Hydrological Balance of Rice Areas in
Northern Italy.” Journal of Hydrology 590: 125401. doi:10.
1016/j.jhydrol.2020.125401.
Ozdogan, M., Y. Yang, G. Allez, and C. Cervantes. 2010. “Remote
Sensing of Irrigated Agriculture: Opportunities and
Challenges.” Remote Sensing 2 (9): 2274–2304. doi:10.3390/
rs2092274.
Pan, H., X. Tong, X. Xu, X. Luo, Y. Jin, H. Xie, and B. Li. 2020.
“Updating of Land Cover Maps and Change Analysis Using
GlobeLand30 Product: A Case Study in Shanghai
Metropolitan Area, China.” Remote Sensing 12 (19): 3147.
doi:10.3390/rs12193147.
Pervez, M. S., and J. F. Brown. 2010. “Mapping Irrigated Lands at
250-m Scale by Merging MODIS Data and National
Agricultural Statistics.” Remote Sensing 2 (10): 2388–2412.
doi:10.3390/rs2102388.
Pervez, M. S., M. Budde, and J. Rowland. 2014. “Mapping
Irrigated Areas in Afghanistan over the past Decade Using
MODIS NDVI.” Remote Sensing of Environment 149: 155–165.
doi:10.1016/j.rse.2014.04.008.
Sab, N. S. C. 2014. “Study on Stability of Reectance
Characteristics of Natural Features for Calibrating Remote
Sensing Data.” PhD diss., Universiti Teknologi Malaysia.
Salmon, J. M., M. Friedl, S. Frolking, D. Wisser, and E. Douglas.
2015. “Global Rain-Fed, Irrigated, and Paddy Croplands:
A New High Resolution Map Derived from Remote Sensing,
Crop Inventories and Climate Data.” International Journal of
Applied Earth Observation and Geoinformation 38: 321–334.
doi:10.1016/j.jag.2015.01.014.
Santana, N. C., O. A. De Carvalho Júnior, R. A. T. Gomes, and
R. F. Guimarães. 2018. “Burned-Area Detection in Amazonian
Environments Using Standardized Time Series per Pixel in
MODIS Data.” Remote Sensing 10 (12): 1904. doi:10.3390/
rs10121904.
Satterwhite, M. B., H. J. Mitchell, T. Hemmer, and J. D. Leckie. 2003.
“Field Spectral Signatures of Snow, Ice, and Water.” Algorithms
and Technologies for Multispectral, Hyperspectral, and
Ultraspectral Imagery IX 5093: 528. doi:10.1117/12.488363.
Seaton, D., T. Dube, and D. Mazvimavi. 2020. “Use of
Multi-Temporal Satellite Data for Monitoring Pool Surface
Areas Occurring in Non-Perennial Rivers in Semi-Arid
Environments of the Western Cape, South Africa.” ISPRS
Journal of Photogrammetry and Remote Sensing 167:
375–384. doi:10.1016/j.isprsjprs.2020.07.018.
Shazadeh-Moghadam, H., M. Khazaei, S. K. Alavipanah, and
Q. Weng. 2021. “Google Earth Engine for Large-Scale Land
Use and Land Cover Mapping: An Object-Based
1618 X. QIAN ET AL.
Classication Approach Using Spectral, Textural and
Topographical Factors.” GIScience & Remote Sensing 58 (6):
914–928. doi:10.1080/15481603.2021.1947623.
Shao, Y., R. S. Lunetta, B. Wheeler, J. S. Iiames, and
J. B. Campbell. 2016. “An Evaluation of Time-Series
Smoothing Algorithms for Land-Cover Classications Using
MODIS-NDVI Multi-Temporal Data.” Remote Sensing of
Environment 174: 258–265. doi:10.1016/j.rse.2015.12.023.
Shi, H. B., S. Q. Yang, R. P. Li, X. Y. Li, W. P. Li, J. W. Yan, Q. F. Miao,
and Z. Li. 2020. “Water-Saving Irrigation and Utilization
Eciency of Water and Fertilizer in Hetao Irrigation District
of Inner Mongolia: Prospect for Future Research.” Journal of
Irrigation and Drainage 39 (11): 1–12. In Chinese with English
abstract. doi:10.13522/j.cnki.ggps.2020155
Thompson, J. A., and D. J. Paull. 2017. “Assessing Spatial and
Temporal Patterns in Land Surface Phenology for the
Australian Alps (2000–2014).” Remote Sensing of
Environment 199: 1–13. doi:10.1016/j.rse.2017.06.032.
Tsela, P., K. Wessels, J. Botai, S. Archibald, D. Swanepoel,
K. Steenkamp, and P. Frost. 2014. “Validation of the Two
Standard MODIS Satellite Burned-Area Products and an
Empirically-Derived Merged Product in South Africa.”
Remote Sensing 6 (2): 1275–1293. doi:10.3390/rs6021275.
Velpuri, N. M., P. S. Thenkabail, M. K. Gumma, C. Biradar,
V. Dheeravath, P. Noojipady, and L. Yuanjie. 2009.
“Inuence of Resolution in Irrigated Area Mapping and
Area Estimation.” Photogrammetric Engineering and Remote
Sensing 75 (12): 1383–1395. doi:10.14358/PERS.75.12.1383.
Vermote, E. F., J.-C. Roger, and J. P. Ray. 2015. ““MODIS Surface
Reectance User’s Guide.” Collection 6 MayRetrieved
fromhttp://modis-sr.ltdri.org/guide/MOD09_UserGuide_v1.
4.pdf
Wang, L. X., M. Y. Cai, X. L. Bai, S. X. JI, B. Jia, X. W. Feng, and
X. L. Chang. 2021. “Estimation of Irrigation Area in Hetao
Irrigation District Based on MODIS Analysis.” Journal of Inner
Mongolia Normal University(Natural Science Edition) 50 (1):
44–52. In Chinese with English abstract. doi:10.3969/j.1001-
8735.2021.01.007
Wang, X., S. Xie, X. Zhang, C. Chen, H. Guo, J. Du, and Z. Duan.
2018. “A Robust Multi-Band Water Index (MBWI) for Automated
Extraction of Surface Water from Landsat 8 OLI Imagery.”
International Journal of Applied Earth Observation and
Geoinformation 68: 73–91. doi:10.1016/j.jag.2018.01.018.
Wang, R., C. Zeng, P. Li, and H. Shen. 2011. “Terra MODIS Band 5
Stripe Noise Detection and Correction Using MAP-Based
Algorithm.” In 2011 International Conference on Remote
Sensing, Environment and Transportation Engineering,
Nanjing, 8612–8615. doi:10.1109/RSETE.2011.5964181.
Wardlow, B. D., and K. Callahan. 2014. “A Multi-Scale Accuracy
Assessment of the MODIS Irrigated Agriculture Data-Set
(Mirad) for the State of Nebraska, USA.” GIScience & Remote
Sensing 51 (5): 575–592. doi:10.1080/15481603.2014.952546.
Wen, Y., S. Shang, and K. U. Rahman. 2019. “Pre-Constrained
Machine Learning Method for Multi-Year Mapping of Three
Major Crops in a Large Irrigation District.” Remote Sensing
11 (3): 242. doi:10.3390/rs11030242.
Wen, Y., S. Shang, K. U. Rahman, Y. Xia, and D. Ren. 2020. “A
Semi-Distributed Drainage Model for Monthly Drainage
Water and Salinity Simulation in A Large Irrigation District
in Arid Region.” Agricultural Water Management 230: 105962.
doi:10.1016/j.agwat.2019.105962.
Wen, Y., H. Wan, S. Shang, and K. U. Rahman. 2022. “A Monthly
Distributed Agro-Hydrological Model for Irrigation District in
Arid Region with Shallow Groundwater Table.” Journal of
Hydrology 609: 127746. doi:10.1016/j.jhydrol.2022.127746.
Xie, Y., and T. J. Lark. 2021. “Mapping Annual Irrigation from
Landsat Imagery and Environmental Variables across the
Conterminous United States.” Remote Sensing of Environment
260 (April): 112445. doi:10.1016/j.rse.2021.112445.
Xie, Y., T. J. Lark, J. F. Brown, and H. K. Gibbs. 2019. “Mapping
Irrigated Cropland Extent across the Conterminous United
States at 30 m Resolution Using a Semi-Automatic Training
Approach on Google Earth Engine.” ISPRS Journal of
Photogrammetry and Remote Sensing 155: 136–149. doi:10.
1016/j.isprsjprs.2019.07.005.
Xiong, L., X. Xu, B. Engel, Q. Huang, Z. Huo, Y. Xiong, W. Han, and
G. Huang. 2021. “Modeling Agro-Hydrological Processes and
Analyzing Water Use in a Super-Large Irrigation District (Hetao)
of Arid Upper Yellow River Basin.” Journal of Hydrology 603:
127014. doi:10.1016/j.jhydrol.2021.127014.
Xu, X., G. Huang, Z. Qu, and L. S. Pereira. 2010. “Assessing the
Groundwater Dynamics and Impacts of Water Saving in the
Hetao Irrigation District, Yellow River Basin.” Agricultural Water
Management 98 (2): 301–313. doi:10.1016/j.agwat.2010.08.025.
Yang, L. Q., Y. F. Liu, and K. Liu. 2017. “The Causes of the
Increase of Spring Irrigation Area in Yichang Irrigation
District.” Inner Mongolia Water Resour 4: 66–67. In Chinese.
Yeom, J.-M., S. Jeong, R. C. Deo, and J. Ko. 2021. “Mapping Rice
Area and Yield in Northeastern Asia by Incorporating a Crop
Model with Dense Vegetation Index Proles from
a Geostationary Satellite.” GIScience & Remote Sensing
58 (1): 1–27. doi:10.1080/15481603.2020.1853352.
Yu, B., and S. Shang. 2017. “Multi-Year Mapping of Maize and
Sunower in Hetao Irrigation District of China with High
Spatial and Temporal Resolution Vegetation Index Series.”
Remote Sensing 9 (8): 855. doi:10.3390/rs9080855.
Zeng, Q. Z., M. S. Cao, X. Z. Feng, F. X. Liang, X. Z. Chen, and
W. K. Sheng. 1984. “Study on Spectral Reection
Characteristics of Snow, Ice and Water of Northwest
China.” Scientia Sinica, Series B 27 (6): 647–656.
Zhang, C., Y. Chen, and D. Lu. 2015. “Detecting Fractional
Land-Cover Change in Arid and Semiarid Urban
Landscapes with Multitemporal Landsat Thematic Mapper
Imagery.” GIScience & Remote Sensing 52 (6): 700–722. doi:10.
1080/15481603.2015.1071965.
Zhan, Z. M., Q. M. Qin, A. Ghulan, and D. D. Wang. 2006. “NIR-
red Spectral Space Based New Method for Soil Moisture
Monitoring.” Scientia Sinica Terrae 36 (11) 46–52. In Chinese.
GISCIENCE & REMOTE SENSING 1619
Appendix A1. Proportions of valid observations from dierent remote sensing data sources
Proportions of valid observations from the three remote sensing data sources, i.e. Sentinel-2, MOD09GA, and MCD43A4, for
each day in the 2019 study period are shown in Figure 13. More than 60% of the dates contain more than 70% of MOD09GA data
that is usable. MCD43A4 has a lower percentage of available data since mid-October and has almost no data from November 9th,
which may result in a high degree of incomplete information in classication maps. Sentinel-2 data can be used for validation due
to its high spatial resolution (10m) and high percentage of data available on certain dates.
Appendix A2. Appropriate water index for mapping irrigated areas
The simplicity of the two-band indices makes them easy to use. Previous studies have shown that when using MODIS data for
calculation, normalized dierence water indices using bands 4 and 7 (NDWI47), 1 and 7 (NDWI17), 4 and 6 (NDWI46), 4 and 5
(NDWI45), or 1 and 6 (NDWI16) perform better (Boschetti et al. 2014; Ji, Zhang, and Wylie 2009). Among them, some studies have
found that NDWI46 has the best performance in water body recognition. It has a low magnitude of error and a stable threshold, and
performs well on mixed pixels (Colditz et al. 2018). Meanwhile, due to the repetitive stripes, we avoid using band 5 (R. Wang et al.
2011). Finally, NDWI46 and NDWI17 are selected that are more sensitive than other two-band WIs in low water percentage
conditions (Boschetti et al. 2014).
Multi-band indices considering more band information may be more suitable to distinguish water bodies when there are
many surfaces that can be easily confused with a water surface. Therefore, we also chose several multi-band indices
proposed in recent years, including MBWI (X. Wang et al. 2018), WI2015 (Fisher, Flood, and Danaher 2016), AWEIsh and
AWEInsh (Feyisa et al. 2014). In addition, a soil moisture index, SMMRS proposed by Zhan et al. (2006) is also compared with
these surface water indices. Six WIs and one soil moisture index used for comparison in this study are presented in Table A1,
which can be broadly divided into two types, i.e. two-band and multi-band indices.
Table A1. List of the seven indices compared in this study. The input is surface reectance (ρ) for bands b1 (ρb1) to b7 (ρb7)
derived from the MODIS multi-spectral bands.
In order to show the dierence of indices clearly in the graph, we normalize the indices which are not in the range of [−1, 1] to
[0, 1] by Equation (A1).
Figure 13. Time series of the proportion of pixels with no missing value in any of the bands used to calculate water index after
preliminary quality control during the study period in 2019 (DOY258–DOY334). DOY represents day of the Year.
Index Equation
NDWI46 ρb4ρb6ð Þ=ðρb4þρb6Þ
NDWI17 ρb1ρb7ð Þ=ðρb1þρb7Þ
MBWI 2 ρb4ρb1ρb2ρb6ρb7Þ
AWELnsh 4 ρb4ρb6ð Þ=ð0:25 ρb2þ2:75 ρb7Þ
AWELsh ρb3þ2:5ρb41:5 ðρb2þρb6Þ 0:25 ρb7
WI2015 1:7204 þ171 ρb4þ3ρb170 ρb245 ρb671 ρb7
SMMRS 1ρb1þMρb2ð Þ=ffiffiffiffiffiffiffiffiffiffiffiffiffi
1þM2
p
Note: M is the slope of the soil line, which is extracted using the near-infrared (NIR) and red bands reflectance of the study area (Zhan
et al. 2006).
1620 X. QIAN ET AL.
x¼xxmin
xmax xmin
(A1)
where x and x* are the original and normalized indices, and x
min
and x
max
are minimum and maximum values of the x series,
respectively.
By comparing the time proles of all sampling points, multi-band water indices are found more sensitive to changes in water
content. They usually have a more signicant increase after autumn irrigation, exhibiting an improvement than the two-band
normalized-dierence water indices (Figure 14). However, the inclusion of more bands also leads to more uctuations in the multi-
band water indices.
In addition, dierent water indices are increased with varying magnitudes in late November (around DOY 325 for the
representative irrigated pixel in Figure 14), possibly due to snow and/or ice. Among them, the increase of MBWI is small and
usually lower than the increase caused by irrigation, while it even does not increase when other indices have abrupt changes in
some cases. This is mainly because the number of absorption bands of ice and snow is similar to that of water, but their maximum
absorption is dierent from that of water (Zeng et al. 1984; Satterwhite et al. 2003). The weights of MBWI in the visible and near-
infrared bands reduce the inuence of ice and snow to some extent. In view of this, we nally chose MBWI to extract irrigation
information.
Appendix A3. Comparison of average feature values of dierent references
The reference should be selected to eliminate the eects of viewing angle, solar zenith angle, data pre-processing, and other
factors on the feature values. It should provide an objective baseline to measure feature values that can be related to autumn
irrigation in a given year.
Persistent water bodies are relatively stable in terms of reectance characteristics and experience little variations over the same
period of multiple years (Sab 2014). Therefore, they can be considered pseudo-invariant features to be used as the normalization
reference (Macfarlane, Grigg, and Daws 2017). However, not all sub-irrigation districts have enough water pixels that can be used as
references. Due to the fact that autumn-irrigated cropland pixels are covered by water during AI, they can be used as a proxy for
water bodies. By calculating the average of the feature values of all input croplands and grasslands pixels in the irrigation district, it
is found that the uctuation is also consistent with the uctuation of the water body pixels (Figure 15). It shows that in large-scale
irrigation districts when the irrigation ratio is high every year, the average value can reect the inherent uctuation of the index and
can be used as a reference to remove the inter-annual uctuations. Furthermore, since the proportion of AI areas in some other
irrigation districts may be small, pixels whose MMAVE-MBWI value ranks between 70% and 100% of all cropland and grassland
pixels were also selected. Figure 15 shows the inter-annual uctuations in the average MMAVE-MBWI for the above three
references. Finally, we selected a simple average of all the input data, because it is accurate enough to reect the inter-annual
changes of feature values in the study area due to factors unrelated to the study objectives. This method of computing baseline
statistics ensures that Z-scores are inuenced by autumn irrigation only.
Figure 14. Time profile of seven different indices for one representative irrigated pixel in the study area in 2019. DOY represents day of
the Year.
GISCIENCE & REMOTE SENSING 1621
Appendix A4. Validation points extracted from Sentinel-2 images from 2016 to 2020
As shown in Figure 16, a total of 3789 autumn irrigated points are selected based on the threshold method and 2694 non-autumn
irrigated points by visual interpretation from 2016 to 2020.
Figure 16. Validation points extracted from Sentinel-2 images from 2016 to 2020, including Autumn irrigation (AI) points (left) and
Non-Autumn irrigation (NAI) points (right).
Figure 15. The average values of the feature variable (MMAVE-MBWI) for water bodies, cropland-grassland, and part of cropland-
grassland with higher MMAVE-MBWI in the whole irrigation district from 2010 to 2020.
1622 X. QIAN ET AL.
Appendix A5. Cumulative areas of autumn irrigation in years from 2010 to 2020
Based on the evolution processes of the autumn irrigation area in dierent years, the cumulative areas of autumn irrigation from
2010 to 2020 was shown in Figure 17.
Figure 17. Cumulative areas of autumn irrigation in years from 2010 to 2020.
GISCIENCE & REMOTE SENSING 1623
... In practice, information about irrigation (including AI) is generally obtained from statistical results of different levels of irrigation administrations, where detailed spatiotemporal information on irrigation application is normally unavailable. As an alternative, satellite remote sensing has proven to be a valuable data source to extract detailed irrigation information at large scales both during crop growth (Karthikeyan et al., 2020) and AI after crop harvest (Qian et al., 2022). Specifically, optical remote sensing is highly effective in determining irrigation extent by identifying differences in the spectral indices of irrigated and non-irrigated cropland (Ozdogan et al., 2010;Yao et al., 2022). ...
... Despite the notable progress in identifying irrigation during the cropgrowing season, identifying autumn irrigation remains rare (Qian et al., 2022). The primary challenge in identifying autumn irrigation lies in the need for high temporal resolution remote sensing data. ...
... The primary challenge in identifying autumn irrigation lies in the need for high temporal resolution remote sensing data. As the surface water ponding caused by irrigation typically lasts for only a few days, current identification methods primarily rely on MODIS data with daily temporal resolution Qian et al., 2022). However, the moderate spatial resolution of MODIS can lead to significant uncertainties in precise mapping of smaller irrigated lands (Pervez et al., 2014). ...
Article
Full-text available
Flood irrigation is widely applied in harvested croplands of arid regions and can be classified as autumn/winter irrigation (AI) depending on the time of application. Due to its high water consumption, real-time monitoring of the AI extent is crucial to improve its scheduling. We proposed a remote sensing-based long short-term memory (LSTM) model for near-real-time monitoring of AI extent at sub-pixel scale. The model loosely coupled MODIS data with Sentinel-2 data to solve the mixed pixel issue of MODIS data, and calibrated Sentinel-2 thresholds for AI identification by a random forest (RF) module to extract large-scale reference data with high temporal frequencies. The variable importance estimated by RF is used as a reference for feature screening in the LSTM model. As Sentinel-2 images are not available daily, LSTM models trained with incomplete sequences were validated using multiple validation approaches. The model was applied to the Hetao Irrigation District, the largest irrigation district in arid region of China, and delivered reasonable performance with coefficient of determination of over 0.82 and mean absolute error of around 10.7%. Classification of irrigation patterns using simulated time series of irrigation area fractions revealed eight different irrigation patterns from 2010 to 2020 in the study region. Results indicate that the maximum fraction of AI area upstream is closely related to the cropland distribution pattern. The AI patterns changed significantly over the 11 years, with a more pronounced reduction in irrigation duration for individual pixels in the downstream area. These changes are associated with two important land policies implemented in many regions of China, land consolidation and land transfer. The proposed model demonstrates the potential for near-real-time monitoring of autumn irrigation extent within large irrigation districts, which can aid in AI scheduling and provide insight into irrigation patterns and practices.
... Sin embargo, el S2 es el satélite de acceso abierto con mejor periodicidad y tamaño de píxel para observar eventos de estrés hídrico (Bellvert et al., 2023). También permite supervisar riegos y así controlar el uso del agua en zonas áridas o semiáridas (Qian et al., 2022) Lin et al. (2020) evaluaron el potencial hídrico del tallo del algodón y su relación por separado con bandas multiespectrales del S2, índices calculados a partir de estas, así como combinación de bandas e índices. Además, se observó el efecto de regresión lineal (RL) y el método de aprendizaje automático bosque aleatorio (RF por sus siglas en inglés). ...
Article
Full-text available
p> Background . Food production as the world population increases is an issue of global importance, resulting in uncertainty around agri-food security (AS). It is essential to apply new technology that monitors and makes crop management more efficient to achieve sustainable development (SD). One technology responsible for improving conventional farmers' practices is precision agriculture (PA) using tools such as big data analytics, artificial intelligence and geospatial analysis. In this sense, the monitoring of markers through satellite images requires knowledge for adequate interpretations. Some satellites have multispectral images, monitoring with these has demonstrated high precision in predicting crop yields. There are several satellites, Landsat 8 (L8) and Sentinel-2 (S2) are the most advanced and open access. Objective . To present the spectrophotometric fundamentals and current state of remote sensing with S2 multispectral images in relation to problems facing agriculture. It is concluded that remote sensing is a tool that contributes to the proper management of agricultural production, showing high reliability. Methodology . This review was carried out in accordance with the PRISMA statement (Preferred Reporting Items for Systematic reviews and Meta-Analyses), a systematic analysis of literature in the databases was carried out. MDPI, Springer Link, Science Direct, Taylor and Francis and Google Scholar were consulted. The key words were “Sentinel-2, R2, Salinity, Pest, Nitrogen, Drought”. Zotero (www.zotero.org) was used as a free access bibliography manager; articles that were not related to the research objectives were excluded. Results . Articles related to the objective of this review were identified in the databases; the study was structured based on the needs of precision agriculture (PA). The importance of adequately managing the water demand of crops, salinity stress, pest detection and the movement of nitrogen (N) in the soil-plant-atmosphere interaction. Implications . Remote perception of phenomena at the field level represents an area for improvement in crop management. Increasing resource efficiency is crucial to achieving agri-food security (AS) and thus part of the sustainable development goals (SDGs). Conclusion . The S2 has multispectral images that, when processed, allow obtaining information about the crop, i.e. monitoring stress, humidity in plants and optimization of plant nutrition. With PA, decisions are made for crop management, costs are reduced by reducing losses due to crop stress, and resource efficiency increases. </p
... Liu and Chen 2019). On the other hand, the information about crop type refers to specific crop varieties (Dahal, Wylie, and Howard 2018), while water management practices include irrigation availability (Qian et al. 2022) and the source of irrigation (Feyisa et al. 2020). ...
Article
Full-text available
Accurate acquisition of spatial and temporal distribution information for cropping systems is important for agricultural production and food security. The challenges of extracting information about cropping systems in regions with smallholder farms are considerable, given the varied crops, complex cropping patterns, and the fragmentation of cropland with frequent reclamation and abandonment. This study presents a specialized workflow to solve this problem for regions with smallholder farms, which utilizes field samples and Sentinel-2 data to extract cropping system information over multiple years. The workflow involves four steps: 1) processing Sentinel-2 data to simulate crop growth curves with the Savitzky‒Golay filter and computing feature variables for classification, including phenology indices, spectral bands, and time series of vegetation indices; 2) mapping annual croplands with one-class support vector machine; 3) mapping various cropping patterns, including single cropping, intercropping, double cropping, multiple harvest, and fallow by decision tree and K-means clustering; and 4) mapping crops with random forest where Jeffries-Matusita distance was used to select appropriate vegetation indices. The workflow was applied in the Hetao irrigation district in Inner Mongolia Autonomous Region, China from 2018 to 2021. The overall accuracies were 0.98, 0.96, and 0.97 for cropland, cropping patterns, and crop type mapping, respectively. The mapping results indicated that the study area has low cropping continuity and is dominated by single cropping patterns. Furthermore, the area of wheat cultivation has decreased, and vegetable cultivation has expanded. Overall, the proposed workflow facilitated the accurate acquisition of cropping system information in regions with smallholder farms and demonstrated the effectiveness of available Sentinel-2 imagery in classifying complex cropping patterns. The workflow is available on Google Earth Engine.
... However, on the one hand, it is not yet clear how to adopt this in practice and, on the other hand, how to apply large irrigation depths when drip irrigation would be adopted. Currently, the progressive use of drip systems in Hetao is apparently concurrent with the application of the autumn irrigation, which area has remained stable over the last decade (Qian et al., 2022). Moreover, subsurface drainage, now being installed, require new irrigation approaches to maximize its impact in improving salt leaching. ...
Article
Full-text available
Soil salinization problems are widespread in the Hetao plain, Inner Mongolia, resulting from arid climate conditions , a shallow saline water table, poor irrigation water management, and insufficient drainage. This study follows previous research aimed at evaluating crop water use and controlling the salinity build-up in the region, namely using weighing and static water table lysimeters to parameterize a water balance model aimed at the development of appropriate irrigation scheduling. Two sets of five static water table lysimeters, which fixed depths ranged from 1.25 to 2.25 m, were used over two maize crop seasons. The mechanistic HYDRUS-1D model was further used to daily predict measured data on soil water contents, boundary fluxes in the interface between the saturated and unsaturated zones, the electrical conductivity of the soil saturation paste extract (EC e), and the actual crop evapotranspiration (ET c act). The soil water balance helped quantify the combined effect of water and salinity stresses on root water uptake as well as groundwater fluxes into the rootzone. The salt balance showed that the salinity build-up was much related to irrigation and capillary fluxes, and that the autumn irrigation carried out during the non-growing season was essential for controlling soil salinity. The efficiency of the autumn irrigation much depended on groundwater depth and the amount of water applied for salt leaching, with the best results found for the lysimeters with water table depths at 2.0 and 2.25 m (85-100%) for irrigation depths ≥ 220 mm. The lysimeters with shallower water tables never showed a leaching efficiency higher than 88%. This research shows that the sustainability of irrigation in Hetao depends on finding adequate solutions for controlling the depth of the saline groundwater, minimizing capillary fluxes to the rootzone, and developing consequent approaches for autumn irrigation leaching.
... In recent years, sunflower and corn have become the dominant crops in the HID, where the sunflower planting area tends to increase (Wen et al., 2019). The sunflower is sowed in late May, which is about one month later than other major crops and has a significant influence on irrigation during crop growing season and Autumn / Spring irrigation during fallow period (Qian et al., 2022). Besides, groundwater extraction for domestic, industrial, and agricultural uses is another human activity that influences water balance in the HID, especially the groundwater balance. ...
Article
Water scarcity and soil salinization are two major problems threatening the sustainability of irrigation districts in arid regions with shallow groundwater table. To study water and salt balances in arid irrigation districts, we developed a monthly distributed water and salt balance model for arid irrigation district (DAHMID-S) with shallow groundwater table by integrating models of major hydrological processes and the associated salt transport processes, including irrigation, evapotranspiration, soil water and groundwater and salt balances, drainage of water and salt through ditches, and interior drainage of water and associated dry drainage of salt between irrigated and non-irrigated lands. The model was successfully applied to the Hetao Irrigation District (HID), the largest irrigation district in arid region of China. Results indicate that evapotranspiration is the dominant water consumption component in the HID and accounts for about 95% of the total irrigation and precipitation, while drainage through ditches accounts for 10.5% of the total irrigation. Only 49.4% of the introduced salt is discharged from HID through drainage ditches. Consequently, most parts of the HID are in the state of salt accumulation except the northern Wulate sub-irrigation district, and the problem of salt accumulation is more serious in the western and southern HID with more favorable water use conditions. Dry drainage plays a crucial role in salt balance of both irrigated and non-irrigated lands, which results in desalinization in most irrigated lands while salt accumulation in non-irrigated lands. Appropriate measures should be taken to solve water and salt-related problems to improve the sustainability of the HID. In the western HID, irrigation water diversion should be decreased, and the drainage system should be further improved to decrease the salt accumulation to a lower level. Meanwhile, the irrigation water diversion can be increased in the southeastern HID by planting more corn and wheat. Besides, Irrigation management should be improved to reduce the surface water return flow to increase the groundwater drainage capacity, especially in the southwestern HID.
Article
Full-text available
Based on MOD09GA/MYD09GA surface reflectance data, a new MODIS snow-cover-extent (SCE) product from 2000 to 2020 over China has been produced by the Northwest Institute of Eco-Environment and Resources (NIEER), Chinese Academy of Sciences. The NIEER MODIS SCE product contains two preliminary clear-sky SCE datasets – Terra-MODIS and Aqua-MODIS SCE datasets and a final daily cloud-gap-filled (CGF) SCE dataset. The first two datasets are generated mainly through optimizing snow-cover discriminating rules over land-cover types, and the latter dataset is produced after a series of gap-filling processes such as aggregating the two preliminary datasets, reducing cloud gaps with adjacent information in space and time, and eliminating all gaps with auxiliary data. The validation against 362 China Meteorological Administration (CMA) stations shows that during snow seasons the overall accuracy (OA) values of the three datasets are larger than 93 %, all of the omission error (OE) values are constrained within 9 %, and all of the commission error (CE) values are constrained within 10 %. Bias values of 0.98, 1.02, and 1.03 demonstrate on a whole that there is no significant overestimation nor a significant underestimation. Based on the same ground reference data, we found that the new product accuracies are obviously higher than standard MODIS snow products, especially for Aqua-MODIS and CGF SCE. For example, compared with the CE of 23.78 % that the MYD10A1 product shows, the CE of the new Aqua-MODIS SCE dataset is 6.78 %; the OA of the new CGF SCE dataset is up to 93.15 % versus 89.54 % of MOD10A1F product and 84.36 % of MYD10A1F product. Besides, as expected, snow discrimination in forest areas is also improved significantly. An isolated validation at four forest CMA stations demonstrates that the OA has increased by 3–10 percentage points, the OE has dropped by 1–8 percentage points, and the CE has dropped by 4–21 percentage points. Therefore, our product has virtually provided more reliable snow knowledge over China; thereby, it can better serve for hydrological, climatic, environmental, and other related studies there.
Article
Full-text available
This study aims to propose an operational approach to map irrigated areas based on the synergy of Sentinel-1 (S1) and Sentinel-2 (S2) data. An application is proposed at two study sites in Europe—in Spain and in Italy—with two climatic contexts (semiarid and humid, respectively), with the objective of proving the essential role of multi-site training for a robust application of the proposed methodologies. Several classifiers are proposed to separate irrigated and rainfed areas. They are based on statistical variables from Sentinel-1 and Sentinel-2 time series data at the agricultural field scale, as well as on the contrasted behavior between the field scale and the 5 km surroundings. The support vector machine (SVM) classification approach was tested with different options to evaluate the robustness of the proposed methodologies. The optimal number of metrics found is five. These metrics illustrate the importance of optical/radar synergy and the consideration of multi-scale spatial information. The highest accuracy of the classifications, approximately equal to 85%, is based on training dataset with mixed reference fields from the two study sites. In addition, the accuracy is consistent at the two study sites. These results confirm the potential of the proposed approaches towards the most general use on sites with different climatic and agricultural contexts.
Article
Full-text available
Irrigation represents one of the most impactful human interventions in the terrestrial water cycle. Knowing the distribution and extent of irrigated areas as well as the amount of water used for irrigation plays a central role in modeling irrigation water requirements and quantifying the impact of irrigation on regional climate, river discharge, and groundwater depletion. Obtaining high-quality global information about irrigation is challenging, especially in terms of quantification of the water actually used for irrigation. Here, we review existing Earth observation datasets, models, and algorithms used for irrigation mapping and quantification from the field to the global scale. The current observation capacities are confronted with the results of a survey on user requirements on satellite-observed irrigation for agricultural water resources’ management. Based on this information, we identify current shortcomings of irrigation monitoring capabilities from space and phrase guidelines for potential future satellite missions and observation strategies.
Article
Full-text available
Mapping the distribution and type of land use and land cover (LULC) is essential for watershed management. The Tigris-Euphrates basin is a transboundary region in the Middle East shared between six countries, but a recent fine-scale LULC map of the area is lacking. Using Landsat-8 time series, a 30-m resolution LULC map was produced for the Tigris-Euphrates basin. In total, 1184 Landsat scenes were processed within the Google Earth Engine (GEE). For the collection of ground truth data, differential manifestations of green cover were considered by dividing the study area into five climatic regions and the training samples were taken from each sub-region. To account for the temporal variation of LULC types, six two-month interval composite layers, including the spectral and thermal bands of Landsat-8, texture and spectral indices, as well as topographic factors were created for the target year 2019. Image segmentation and classification were performed using the simple non-iterative clustering (SNIC) and Random Forest (RF) algorithms, respectively. A computationally effective parallel processing approach was developed, which created a number of tiles and sub-tiles and a bulk command was converted into smaller parallel commands. The generated LULC map showed a satisfactory overall accuracy of 91.7%, with the highest User accuracy in water and wetland, and the lowest in rainfed crop and rangeland, and the highest Producer’s accuracy in water and barren areas, and the lowest in garden and rangeland. This study highlights the necessity of using multi-temporal data for LULC mapping, in particular, multi-temporal NDVI, for the separation of different green cover types in the arid and semi-arid environments.
Article
Agro-hydrological processes in arid irrigation districts mainly include precipitation, water diversion, irrigation, drainage, evapotranspiration (ET), and soil water and groundwater flow, which interact with each other and are controlled by complex natural and anthropogenic drivers. To better understand the agro-hydrological processes in arid irrigation districts with shallow groundwater table, we developed a novel monthly distributed agro-hydrological model for irrigation districts (DAHMID) based on the concepts of canal command area (CCA) and sub-drainage command area (SDCA). The DAHMID model is driven by meteorology, irrigation, and evapotranspiration (ET) estimated from remote sensing-based ET model, and considers soil and groundwater balances in both irrigated and non-irrigated lands and interior drainage between them. The model was applied to Hetao Irrigation District (HID), the largest irrigation district in arid region of China with a total irrigated area of 0.68 million ha. The DAHMID model was calibrated with groundwater table depth measurements in 13 CCAs of HID from 2008 to 2010, and validated from 2012 to 2013. Results depicted that the root mean square errors (RMSEs), normalized RMSEs (NRMSEs), Nash-Sutcliffe efficiency coefficients (NSEs), and coefficients of determination (r²) of groundwater table depth in both irrigated and non-irrigated lands for all CCAs were in the ranges of 0.19-0.34 m, 0.10-0.25, 0.30-0.82, and 0.68-0.91, respectively. The simulation results from 2008 to 2014 indicated that interior drainage from irrigated land to non-irrigated land is an important measure of drainage in HID, which is about 14.3% of total irrigation water diversion and 34.9% more than the drainage through ditches. The interior drainage process is basically similar to irrigation and ditch drainage processes, all reaching their peaks in May and October. Evapotranspiration is the major water consumption in HID, which is about 95% of total irrigation water diversion and precipitation in average. The net capillary rise of irrigated land is significantly less than that of non-irrigated land due to the impact of irrigation infiltration. The DAHMID model has less parameters and requires less inputs, and can be better applied to continuous simulation of agro-hydrological processes in irrigation districts in medium and long periods with satisfactory simulation accuracy.
Article
Despite a detailed knowledge of the spatial-temporal dynamics of irrigation being necessary to optimize the agricultural production without exacerbating the pressure exercised on the water resource, such information is still often lacking worldwide. In this study, a double-scale analysis on the detectability of the irrigation occurrence over an area in central Italy through remote sensing soil moisture is proposed; the period of interest is a 3-year time span from 2017 to 2019. The detectability of district- or sub-district-scale irrigation signals through remotely sensed soil moisture data is investigated at two different spatial resolutions: 1 km and plot scale. Three soil moisture products sampled at 1 km resolution are evaluated: a DISPATCH (DISaggregation based on Physical And Theoretical scale CHange) downscaled version of SMAP (Soil Moisture Active Passive) and two Sentinel-1-derived products, namely the 1 km version delivered by Copernicus and a plot-scale-born version developed by THEIA and aggregated at 1 km. The THEIA Sentinel-1 product aggregated at 100 m is used in the plot-scale analysis. Over the study area, the irrigation extent is determined by the fragmentation of the agricultural fields and the complex topography, making the adoption of plot-scale data necessary. Satisfactory results are obtained by comparing maps of irrigated areas at 100 m spatial resolution produced through the k-means clustering algorithm with ground-truth data, since the method fails only once out of seven in properly reproducing the irrigated or non-irrigated conditions occurred over four pilot agricultural fields.
Article
High water consumption in large irrigation districts has resulted in large river water diversion and serious water use conflicts in the Yellow River basin (YRB). The excessive water applied is also a cause of the dual problems of soil salinity and water logging in the upper YRB, which further negatively impacts crop yields. This study is aimed at modeling and analyzing the agro-ecohydrological processes and water use performance in large irrigation districts of the YRB. The Hetao Irrigation District (Hetao), a super-large irrigation district located in the upper YRB, was selected as a typical research area. A watershed hydrological model (SWAT-AG) with enhanced capacity for simulating regional agro-ecohydrological processes was adopted in this study. We simulated the agro-ecohydrological processes in Hetao and analyzed the regional characteristics of depth to groundwater table, soil moisture and salt content, evapotranspiration and crop yield. Water accounting was further carried out to interpret water cycle processes and evaluate water use performance in Hetao. Results indicated that Hetao was characterized by shallow depth to groundwater table, high soil moisture and high evapotranspiration. For farmland, there was a decreasing trend for both soil moisture and evapotranspiration from the upper to lower regions in Hetao, but this decreasing trend was not found for natural land. Further, results showed soil salt content has strong spatial variations with land use type, crops and shallow groundwater environments. Correspondingly, crop yields presented a decreasing trend from the upper to lower regions of Hetao, with large spatial variabilities closely related to the spatial distribution of soil water and salt conditions across Hetao. It was worth noting that the regional water use efficiency was relatively low in Hetao, especially with large proportions of non-beneficial water consumption to total water consumption. Overall, this study provided a thorough understanding of agro-ecohydrological processes and their regional characteristics, and the current situation of irrigation water use issues in the whole Hetao district, and further provided valuable insights in improving water use and management in large irrigation districts of the upper YRB.
Article
Identifying the location of irrigated croplands and how they change over time is critical for assessing and managing limited water resources to navigate such challenges as local to global water scarcity, increasing demands for food and energy production, and environmental sustainability. Although efforts have been made to map irrigated area for the U.S., multi-year nationwide maps at field-relevant resolutions are still unavailable; existing products suffer from coarse resolution, uncertain accuracy, and/or limited spatial coverage, especially in the eastern U.S. In this study, we present an approach to map the extent of irrigated croplands across the conterminous U.S. (CONUS) for each year in the period of 1997–2017. To scale nationwide, we developed novel methods to generate training datasets covering both the western and eastern U.S. For the more arid western U.S., we built upon the methods of Xie et al. (2019) and further developed a greenness-based normalization technique to estimate optimal thresholds of crop greenness in any year based on those in USDA NASS census years (i.e., 1997, 2002, 2007, 2012, and 2017). For the relatively humid eastern states, we collected data on the current status of center pivot irrigated and non-irrigated fields and extended these sample points through time using various indices and observational thresholds. We used the generated samples along with remote sensing features and environmental variables to train county-stratified random forest classifiers annually for pixel-level classification of irrigated extent each year and subsequently implemented a logic-based post-classification filtering. The produced Landsat-based Irrigation Dataset (LANID-US) accurately reconstructed NASS irrigation patterns at both the county and state level while also supplying new annual area estimates for intra-epoch years. Nationwide pixel-level locational assessment further demonstrated an overall accuracy above 90% across years. In the 21-year study period, we found several hotspots of irrigation change including significant gains in the U.S. Midwest, the Mississippi River Alluvial Plain, and the East Coast as well as irrigation declines in the central and southern High Plains Aquifer and the southern California Central Valley, Arizona, and Florida. The resulting 30 m resolution LANID-US products represent the finest resolution account of nationwide irrigation use and dynamics across the United States to date. The developed approach, training data, and products are further extendable to other years (either before 1997 or after 2017) for continuous monitoring of irrigated area over CONUS and are spatially applicable to other regions with similar climate and cropping landscapes.
Article
Although irrigation practices affect food production and water resource management, with ever more impacting effects under climate change and population increasing scenarios, detailed knowledge of irrigation is still lacking. In fact, explicit information on the spatial occurrence of irrigation and on the amounts of water used for this purpose is often not available, thus making irrigation the missing variable to comprehensively understand the hydrological cycle dynamics over agricultural areas. Nevertheless, remote sensing techniques can be used to delimit the irrigation extent. In this study, the capability of five remotely sensed soil moisture products to detect the irrigation signal over an area intensely equipped for irrigation in the North East of Spain is investigated; moreover, a method to map the actually irrigated areas based on the K-means clustering algorithm is proposed. The remote sensing soil moisture data sets used in this study are SMOS (Soil Moisture and Ocean Salinity) at 1 km, SMAP (Soil Moisture Active Passive) at 1 km and 9 km, Sentinel-1 at 1 km, and ASCAT (Advanced SCATterometer) at 12.5 km. The 1 km resolution versions of SMOS and SMAP are obtained by downscaling coarser SMOS and SMAP data through the DISPATCH (DISaggregation based on Physical And Theoretical scale CHange) algorithm. The analyses are supported by an additional data set of soil moisture at 1 km resolution simulated by the SURFEX-ISBA (SURFace EXternalisée - Interaction Sol Biosphère Atmosphère) land surface model. Among all the considered data sets, the L-band passive microwave downscaled products show the best performances in detecting the irrigation signal over the pilot area, especially SMAP at 1 km. The proposed maps of irrigated areas derived by exploiting soil moisture from SMAP at 1 km data set agree well (up to 78%) with the ground truth derived irrigated areas. Furthermore, the method is able to well distinguish the actually irrigated areas from rainfed agricultural areas, thus representing a useful tool to obtain reliable spatial information about the areas where irrigation actually occurs.
Article
The selection of calibration and validation time periods in hydrologic modeling is often done arbitrarily. Nonstationarity can lead to an optimal parameter set for one period which may not accurately simulate another. However, there is still much to be learned about the responses of hydrologic models to nonstationary conditions. We investigated how the selection of calibration and validation periods can influence water balance simulations. We calibrated Soil and Water Assessment Tool hydrologic models with observed streamflow for three United States watersheds (St. Joseph River of Indiana/Michigan, Escambia River of Florida/Alabama, and Cottonwood Creek of California), using time period splits for calibration/validation. We found that the choice of calibration period (with different patterns of observed streamflow, precipitation, and air temperature) influenced the parameter sets, leading to dissimilar simulations of water balance components. In the Cottonwood Creek watershed, simulations of 50‐year mean January streamflow varied by 32%, because of lower winter precipitation and air temperature in earlier calibration periods on calibrated parameters, which impaired the ability for models calibrated to earlier periods to simulate later periods. Peaks of actual evapotranspiration for this watershed also shifted from April to May due to different parameter values depending on the calibration period's winter air temperatures. In the St. Joseph and Escambia River watersheds, adjustments of the runoff curve number parameter could vary by 10.7% and 20.8%, respectively, while 50‐year mean monthly surface runoff simulations could vary by 23‐37% and 169‐209%, depending on the observed streamflow and precipitation of the chosen calibration period. It is imperative that calibration and validation time periods are chosen selectively instead of arbitrarily, for instance using change point detection methods, and that the calibration periods are appropriate for the goals of the study, considering possible broad effects of nonstationary time series on water balance simulations. It is also crucial that the hydrologic modeling community improves existing calibration and validation practices to better include nonstationary processes. This article is protected by copyright. All rights reserved.