Download as pdf or txt
Download as pdf or txt
You are on page 1of 226

ABSTRACT

PERIN, VINICIUS. Quantifying On-farm Reservoir Dynamics: An Earth Observation and


Hydrological Modeling Approach. (Under the direction of Dr. Mirela G. Tulbure).

On-farm reservoirs (OFRs) are artificial water bodies that enable farmers to store water

during the raining season to support irrigation activities during the crop growing season. Although

the OFRs have a critical role in irrigated food production, these water bodies impact the hydrology

of the watersheds where they occur by decreasing peak flow and overall streamflow. Nonetheless,

OFRs are poorly monitored or not monitored in many countries, due to the OFRs dynamic nature,

their occurrence in high numbers (i.e., thousands), and their location mostly on private properties.

Aiming to improve OFRs monitoring, and to quantify their impact on surface water availability,

this dissertation explores various remote sensing technologies, including novel approaches and

imagery datasets, in combination with hydrological modeling to further understand the OFRs

occurrence and surface area changes. The study region is in eastern Arkansas, the third most

irrigated region in the USA with a high occurrence of OFRs.

Chapter 2 explores the use of Landsat-based inundation datasets, the U.S. Geological

Survey Dynamic Surface Water Extent (DSWE) and the European Commission’s Joint Research

Centre (JRC) Global Monthly Water History, to assess the OFRs surface area changes. This study

aimed to compare the performance of both datasets when monitoring OFRs of different sizes. Our

results showed that both datasets allow us to estimate the seasonality of the OFRs surface area

changes, their frequency of inundation, and their maximum extent. Nonetheless, these datasets are

limited to a few observations a year, due to sensor related issues and the 16-day repeat cycle, and

the spatial resolution (30 m) of the datasets is unsuitable to monitor OFRs smaller than 5 ha.

Chapter 3 presents a multi sensor satellite imagery approach based on the Kalman filter to

improve OFRs monitoring. The algorithm was used to assimilate data from freely available (i.e.,
Sentinel 1 and 2) and commercial (i.e., PlanetScope and RapidEye) satellite imagery to monitor

the OFRs sub-weekly surface area changes. Given that all sensors tended to underestimate the

OFRs surface area changes, the Kalman filter approach also underestimated the OFRs surface area;

however, the mean percent error was smaller than 12%. The algorithm allows monitoring the

spatial and temporal variability of a network of OFRs, and it can be used to monitor water use

trends. Results of this work can be employed to improve water management recommendations to

increase water use efficiency.

Chapter 4 explores the use of Planet Fusion and Planet Basemap—the next generation of

daily high-resolution (3 m) analysis ready datasets—to monitor OFRs. These datasets offer an

unprecedented opportunity to improve OFRs monitoring. Although both datasets are based on

PlanetScope imagery, they are generated using different algorithms and data sources. In general,

Planet Basemap presented higher surface area variability and it was more susceptible to the

presence of clouds and haze when compared to Planet fusion, which had a smoother time series

with less variability and fewer abrupt changes in the time series. Both datasets can help improve

fresh water management by allowing better assessment of the OFRs dynamics.

Chapter 5 combines satellite based information (e.g., derived from the Kalman filter

approach) with the Soil Water Assessment Tool+ hydrological model to further understand the

impacts of the OFRs on surface water availability. The results indicate that the OFRs do not have

an equally distributed impact on flow and peak flow across the watershed. In addition, the impact

of the OFRs will change according to different OFRs and water channels. We expect that the

results from this work will be improved in future studies, and that it can support water agencies

with information to enhance surface water resources management.


© Copyright 2022 by Vinicius Perin

All Rights Reserved


Quantifying On-farm Reservoir Dynamics: An Earth Observation
and Hydrological Modeling Approach

by
Vinicius Perin

A dissertation submitted to the Graduate Faculty of


North Carolina State University
in partial fulfillment of the
requirements for the degree of
Doctor of Philosophy

Geospatial Analytics

Raleigh, North Carolina


2022

APPROVED BY:

_______________________________ _______________________________
Mirela G. Tulbure Joshua Gray
Committee Chair

_______________________________ _______________________________
Brian Reich A. Sankarasubramanian Arumugam
DEDICATION

To Professor Paulo Cesar Sentelhas (1964–2021), who inspired this work.

ii
BIOGRAPHY

Vini was born in Salto, Sao Paulo, Brazil. He received a B.S. in Agronomy by the

University of Sao Paulo, and during his undergraduate studies, He studied one year abroad at

Wageningen University and Research Centre, The Netherlands, where He was introduced to

Remote Sensing and GIS. After receiving his B.S., He moved to Manhattan, Kansas, to pursue a

M.S. in Agronomy and Environmental Science. During his M.S., He focused his coursework on

Remote Sensing, GIS, and geospatial modeling. In August 2019, Vini joined the Center for

Geospatial Analytics to pursue his Ph.D. under the guidance of Dr. Mirela G. Tulbure. Vini

completed his degree in Fall 2022 before joining Planet Labs as a Geospatial Software Engineer.

iii
ACKNOWLEDGMENTS

Extremely grateful to my fiancé Isabella Possignolo, to Dr. Mirela G. Tulbure, to Mollie

D. Gaines, to Dr. Joshua Gray, to Dr. Brian Reich, to Dr. A. Sankarasubramanian Arumugam, to

Dr. Mary A. Yeager, to Dr. Michele L. Reba, to Dennis Carman, to Rachel Kasten and to Dr.

Megan Skrip for all the support during the past three years. My sincere gratitude to the Center for

Geospatial Analytics, to the National Aeronautics and Space Administration, and Planet Labs for

funding and supporting this work. A huge thanks to my family and friends that supported me

throughout this journey.

iv
TABLE OF CONTENTS

LIST OF TABLES ...................................................................................................................... viii


LIST OF FIGURES ....................................................................................................................... x

CHAPTER 1: INTRODUCTION ............................................................................................... 1


References .................................................................................................................................... 11

CHAPTER 2: ON-FARM RESERVOIR MONITORING USING LANDSAT


INUNDATION DATASETS ..................................................................................................... 14
Title Page and Abstract ................................................................................................................ 14
1 Introduction ............................................................................................................................... 16
2 Material and Methods ............................................................................................................... 19
2.1 Study location ............................................................................................................... 19
2.2 On-farm reservoirs ground-truth dataset ....................................................................... 20
2.3 Processing satellite imagery time series ....................................................................... 22
2.3.1 Dynamic Surface Water Extent (v.1.0) ............................................................... 23
2.3.2 Joint Research Centre Monthly Water History (v.1.1) ........................................ 24
2.4 Climate Information ...................................................................................................... 25
2.5 Data Analysis ................................................................................................................ 25
2.5.1 On-farm reservoir size, missing information and reservoir geometry ................. 25
2.5.1 Generalized additive mixed model approach ....................................................... 26
3 Results and Discussion ............................................................................................................. 28
4 Conclusions ............................................................................................................................... 46
References .................................................................................................................................... 48

CHAPTER 3: A MULTI-SENSOR SATELLITE IMAGERY APPROACH TO MONITOR


ON-FARM RESERVOIRS ....................................................................................................... 55
Title Page and Abstract ................................................................................................................ 55
1 Introduction ............................................................................................................................... 58
1.1 Background on OFRs and remotely sensed data uses for OFR monitoring ................. 58
1.2 Algorithms to combine different satellite imagery for OFR monitoring ...................... 60
2 Data and Methods ..................................................................................................................... 62
2.1 Study region and OFRs dataset ..................................................................................... 62
2.2 Satellite imagery processing ......................................................................................... 64
2.2.1 PlanetScope CubeSat and RapidEye Ortho tiles ................................................. 66
2.2.2 Sentinel 1 and 2 .................................................................................................... 67
2.3 Surface water classification .......................................................................................... 68
2.4 Validation scheme and surface area uncertainties ........................................................ 69
2.5 The Kalman filter approach .......................................................................................... 72
2.6 Kalman filter setup and validation ................................................................................ 74
2.7 Kalman filter analysis and study cases ......................................................................... 76
3 Results ....................................................................................................................................... 78
3.1 Comparing the OFRs’ surface area with the validation dataset .................................... 78
3.2 Same day pairwise comparisons between PS with RE, S2, S1 ..................................... 81
3.3 Kalman filter algorithm performance analysis ............................................................. 83

v
3.4 Kalman filter applications ............................................................................................. 86
4 Discussion ................................................................................................................................. 98
4.1 OFRs surface area uncertainties.................................................................................... 98
4.2 Kalman filter as a data assimilation algorithm ........................................................... 100
4.3 Research impact and applications ............................................................................... 103
5 Future improvements .............................................................................................................. 104
6 Conclusion .............................................................................................................................. 105
References .................................................................................................................................. 107

Chapter 4: MONITORING SMALL WATER BODIES USING HIGH SPATIAL AND


TEMPORAL RESOLUTION ANALYSIS READY DATASETS ...................................... 116
Title Page and Abstract .............................................................................................................. 116
1 Introduction ............................................................................................................................. 118
2 Methods................................................................................................................................... 121
2.1 Study region ................................................................................................................ 121
2.2 Satellite imagery datasets ............................................................................................ 123
2.2.1 PlanetScope CubeSat Surface-Reflectance Ortho tiles ..................................... 123
2.2.2 Normalized Surface-Reflectance Basemap........................................................ 124
2.2.3 Planet Fusion Surface Reflectance..................................................................... 126
2.3 Data analysis ............................................................................................................... 127
2.4 Validation scheme ....................................................................................................... 126
3 Results ..................................................................................................................................... 131
3.1 Surface water validation using SkySat imagery ......................................................... 131
3.2 Number of surface area observations per dataset ....................................................... 132
3.3 Planet Fusion comparison with PlanetScope .............................................................. 131
3.4 Monthly comparisons between Basemap and Planet Fusion with PlanetScope ......... 134
3.5 OFR surface area time series ...................................................................................... 137
4 Discussion ............................................................................................................................... 139
5 Known issues and limitations ................................................................................................. 143
6 Conclusion .............................................................................................................................. 147
References .................................................................................................................................. 149

Chapter 5: ASSESSING THE IMPACTS OF ON-FARM RESERVOIRS ON MODELED


SURFACE HYDROLOGY ..................................................................................................... 155
Title Page and Abstract .............................................................................................................. 155
1 Introduction ............................................................................................................................. 158
1.1 The impact of OFRS on surface hydrology and different modeling techniques ........ 157
1.2 The Soil Water Assessment Tool to model the impacts of OFRs............................... 160
2 Methods................................................................................................................................... 162
2.1 Study region ................................................................................................................ 162
2.2 SWAT+ model setup .................................................................................................. 164
2.3 Model calibration and validation procedures .............................................................. 165
2.4 OFRs representation in SWAT+ ................................................................................. 166
2.5 Scenario analysis ......................................................................................................... 167
3 Results ..................................................................................................................................... 169
3.1 Model calibration and validation ................................................................................ 169

vi
3.2 Percent change in flow ................................................................................................ 172
3.3 Impact on peak flow.................................................................................................... 177
3.4 Simulated flow time series .......................................................................................... 178
4 Discussion ............................................................................................................................... 180
5 Future improvements .............................................................................................................. 185
6 Conclusion .............................................................................................................................. 186
References .................................................................................................................................. 188

CHAPTER 6: CONCLUSION................................................................................................ 195

APPENDICES .......................................................................................................................... 200


APPENDIX A: SUPPLEMENTAL MATERIAL ACCOMPANYING CHAPTER 2: ON-FARM
RESERVOIR MONITORING USING LANDSAT INUNDATION DATASET .................... 201

APPENDIX B: SUPPLEMENTAL MATERIAL ACCOMPANYING CHAPTER 5:


ASSESSING THE IMPACTS OF ON-FARM RESERVOIRS ON MODELED SURFACE
HYDROLOGY .......................................................................................................................... 209

vii
LIST OF TABLES

Chapter 2

Table 1 Hydrological Unit Code names and the number of OFRs in each HUC from the
ground-truth dataset .................................................................................................. 21

Table 2 Global generalized additive mixed model results using precipitation, maximum
temperature, and NDVI as fixed effects, and HUC as a random factor for the DSWE
and JRC .................................................................................................................... 37

Table A1 All generalized additive mixed models fitted employing SAratio as the response
variable for the DSWE and JRC. Precipitation, air temperature, and NDVI were
modelled as fixed effects, and month and year as random effects. The best model
was selected using the lowest Akaike Information Criterion (AIC), the highest
adjusted coefficient of determination (r2), and the least number of parameters. The
relative contribution (contri.) of each explanatory variable depicts the variable
contribution to r2. The selected models are highlighted with a bounding box ....... 203

Table A2 Global generalized additive mixed models fitted employing SAratio as the response
variable for the DSWE and JRC. Precipitation, air temperature, and NDVI were
modelled as fixed effects, and month and year as random effects. The best model
was selected using the lowest Akaike Information Criterion (AIC), the highest
adjusted coefficient of determination (r2), and the least number of parameters. The
relative contribution (contri.) of each explanatory variable depicts the variable
contribution to r2. The selected model is highlighted with a bounding box. .......... 204

Chapter 3

Table 1 Different sensors wavelengths bands (green and NIR, used to calculate NDWI), spatial
resolution, revisit time, and total number of image clips used between January 2017
and December 2020 including all OFRs .................................................................. 66

Table 2 SkySat image identification, image acquisition date, number of OFRs surface area
observations per image, cloud clear (indicates the percentage of cloud cover), ground
control ratio (GCR*, defines the image positional accuracy, and values closer to 1
means higher accuracy), and ground sampling distance (GSP) ............................... 71

Table 3 Variance used for PS, RE, S2 and S1, when implementing the Kalman filter for OFRs
of different size classes and Si ranges. The variance calculated by comparing each
sensors’ surface area and with the validation surface area ....................................... 75

Table 4 OFRs study cases surface area and Si ....................................................................... 77

viii
Chapter 4

Table 1 PlanetScope, Basemap, and Planet Fusion images spatial resolutions and different
wavelengths bands. ................................................................................................. 122

Table 2 Selected OFRs (see Fig. 8) to illustrate PlanetScope, Basemap, and Planet Fusion
surface area time series, and their size according to the OFRs dataset. ................. 129

Table 3 SkySat image identification, acquisition date, number of OFRs surface area
observations per image, percent clear (indicates the presence or absence of cloud
cover; higher values indicate fewer clouds), ground-control ratio (defines the image
positional accuracy; values closer to 1 mean higher accuracy), and ground-sampling
distance in meters.. ................................................................................................. 130

Table 4 Pairwise comparisons between the SkySat validation surface area and the surface area
obtained from PlanetScope, Basemap, and Planet Fusion for multiple observations in
time and divided into three size classes.................................................................. 132

Chapter 5

Table 1 USGS stations, drainage areas, and the periods used for flow calibration and
validation. ............................................................................................................... 165

Table 2 Monthly flow calibration parameters. .................................................................... 166

Table B1 SWAT+ parameters tested in sensitivity analysis. Parameters are ordered from most
to least important. ................................................................................................... 209

ix
LIST OF FIGURES

Chapter 2

Figure 1 Study region in eastern Arkansas, U.S. The study region encompasses the main HUCs
watersheds and the OFRs ground-truth dataset. Inset rectangles depict: (1) JRC (Pekel
et al., 2016), (2) DSWE (Jones et al., 2015, 2019), and (3) Potential irrigated crops
from CropScape, Cropland Data Layer (NASS–USDA, 2020a). The inset rectangles
also contain the OFRs. For HUC ID information please see Table 1. ..................... 20

Figure 2 The cumulative distribution of OFRs surface area extent (ha) in the ground-truth
dataset. OFRs with surface area higher than 50 ha (3.3% of total) are not shown on
this figure to avoid over plotting .............................................................................. 22

Figure 3 The distribution of minimum, maximum and mean air temperature, NDVI, and
precipitation for HUCs A and B considering the entire time series, and all OFRs
located on these HUCs. Mean air temperature is represented by the black solid line,
and minimum and maximum air temperature are represented by the red dashed lines.
The boxplots represent the NDVI monthly distribution. The violin plots represent the
precipitation monthly distribution. Please see Table 1 for HUC ID information .... 29

Figure 4 Comparison of each OFR ground-truth area with the highest surface area recorded
from both inundation datasets (DSWE and JRC) during the period of analysis (1995–
2018). Inset boxplots depict the distribution of MPE, and the p-value represents the
significance of the one-way ANOVA comparing MPE from DSWE and JRC for each
class size ................................................................................................................... 31

Figure 5 Distribution of the OFRs geometry index (SI) per reservoir class size. The boxplots
depict the SI distribution for all different class sizes ................................................ 33

Figure 6 Monthly NAindex and SAratio aggregated per OFR class size including the entire
period of analysis (1995–2018). Bars represented in brighter colors (yellows to greens)
indicate higher NAindex, and wider bars depict higher SAratio ................................... 34

Figure 7 Generalized additive mixed model adjusted r2 values at HUC level (i.e., spatial scale),
accounting for precipitation, maximum air temperature, and NDVI as fixed effects.
The left-hand side map depicts the results for the DSWE, and the right-hand side map
for the JRC. Please see Table 1 for HUC ID information ........................................ 40

Figure 8 Generalized additive mixed model fixed effects of precipitation, maximum


temperature and NDVI at HUC level (i.e., spatial scale). The left-hand side map
depicts the results for the DSWE, and the right-hand side map for the JRC ........... 42

Figure A1 The relationship between SAratio derived from the DSWE, between climate factors
and NDVI, and the distribution of all variables, and their correlation coefficient . 205

x
Figure A2 The relationship between SAratio derived from the JRC, between climate factors and
NDVI, and the distribution of all variables, and their correlation coefficient ........ 206

Figure A3 The distribution of temperature, NDVI, and precipitation for HUCs: C, D, E, F, G,


and H, and considering the entire time series (1995–2018), and all OFRs located on
these HUCs. The black and red dashed lines represent the average, maximum and
minimum temperature. The boxplots represent the NDVI monthly variation. The
violin plots represent the monthly variation of precipitation. Please see Table 1 for
HUC ID information .............................................................................................. 207

Figure A4 Example of the moving average filter with smoother of magnitude 2 for SAratio time
series-series for HUCs A and G, and both DSWE and JRC. Please see Table 1 for
HUC ID information .............................................................................................. 208

Chapter 3

Figure 1 Study region in eastern Arkansas, USA, and the sub-watersheds where the OFRs are
located. The frequency distribution of the OFRs’ surface area (ha) size classes and S i
ranges........................................................................................................................ 64

Figure 2 Workflow used to download and process the satellite imagery used to estimate the
OFRs’ surface area time series from PS, RE, S2, and S1 between January 2017 and
December 2020......................................................................................................... 65

Figure 3 Frequency distribution of the OFRs’ surface area validation dataset derived from
SkySat imagery and grouped into different size classes and Si ranges .................... 72

Figure 4 OFRs outlines (pink lines) overlaid on high-resolution Google maps satellite imagery.
.................................................................................................................................. 77

Figure 5 Linear relationships of the pairwise comparisons between the OFRs’ validation area
and the surface area obtained from PS, RE, S2, and S1 for observations grouped into
different size classes (0.1–5 ha, 5–10 ha, and 10–50 ha) ......................................... 79

Figure 6 Percent error calculated from the pairwise comparisons between the OFRs’ validation
area and the surface area obtained from the PS, RE, S2, and S1 for observations
grouped into different size classes (0.1–5 ha, 5–10 ha, and 10–50 ha) and Si ranges
(0.10–0.50, 0.50–0.75, and 0.75–1.0)....................................................................... 80

Figure 7 Surface area same day pairwise comparisons between PS and RE, S2, and S1 for OFRs
grouped into different size classes (0.1–5 ha, 5–10 ha, and 10–50 ha) .................... 82

Figure 8 Surface area percent error and MAPE for the same day pairwise comparisons between
PS and RE, S2, and S1 for OFRs grouped into different size classes (0.1–5 ha, 5–10
ha, and 10–50 ha) and Si ranges (0.10–0.50, 0.50–0.75, and 0.75–1.0) .................. 83

xi
Figure 9 Linear relationships of the pairwise comparisons between the OFRs’ validation area
and the surface area obtained from the Kalman filter divided into different size classes
(0.1–5 ha, 5–10 ha, and 10–50 ha), and the percent error and MAPE for observations
grouped into different size classes and Si ranges (0.10–0.50, 0.50–0.75, and 0.75–1.0).
.................................................................................................................................. 84

Figure 10 The distribution of r2 and MAPE when comparing the Kalman filter surface area
estimates with the independent PS subset—20% of the PS observations that were not
used in the Kalman filter. Boxplots are divided into different size classes accounting
for all OFRs in the OFRs dataset. The boxplots and histograms are colored according
to both Kalman filter scenarios, with and without S1 observations. ........................ 86

Figure 11 Distribution of the number of observations per month for PS, RE, S2, S1, and the
Kalman filter algorithm, considering all OFRs’ surface area between time series
between January 2017 and December 2020. ............................................................ 88

Figure 12 Sub-weekly surface area time series obtained from the Kalman filter for the study case
OFRs (see Table 4) between January 2017 and December 2020. Grey shaded area
represents +/- one and two standard deviations. The r2 and MAPE values were derived
from the Kalman filter comparisons with the independent PS subset ...................... 89

Figure 13 Mean weekly surface area variation (%) from the annual mean for OFRs study cases
represented by the size and color of the bars. The inter-annual variability is represented
by the error bars, which indicate +/- one standard deviation calculated from all surface
area variation values for a particular week of the year between 2017–2020 ........... 93

Figure 14 Kalman filter total weekly surface area variability considering all OFRs grouped into
different size classes (0.1–5 ha, 5–10 ha, and 10–50 ha) and Si ranges (0.10–0.50,
0.50–0.75, and 0.75–1.0). Y-axis for the three size classes is scaled differently due to
the different magnitude of the total surface area for the larger OFRs ...................... 97

Chapter 4

Figure 1 Study region in eastern Arkansas, USA, and the OFRs size distribution. The inset map
represents the OFRs shapefile overlaid on SkySat satellite imagery ..................... 122

Figure 2 Workflow used to estimate the OFRs’ surface area-time series from PlanetScope,
Basemap, and Planet Fusion between July 2020 and July 2021 ............................ 123

Figure 3 Pairwise comparisons between the SkySat validation surface area and the surface area
obtained from PlanetScope, Basemap, and Planet Fusion for multiple observations in
time ......................................................................................................................... 131

Figure 4 Number of surface area observations per month for PlanetScope and Basemap (A),
and for Planet Fusion real, mixed, and synthetic (C). The frequency distribution of the

xii
total number of observations per OFR per year for PlanetScope and Basemap (B), and
for Planet Fusion real, mixed, and synthetic (D) .................................................... 133

Figure 5 Same day pairwise comparisons between PlanetScope and Planet Fusion real, mixed,
and synthetic, for multiple observations in time and for all OFRs divided into three
size classes (0.1–5 ha, 5–10 ha, and 10–30 ha). Brighter colors indicate higher point
density .................................................................................................................... 135

Figure 6 Frequency distribution of r2 and MAPE calculated by comparing the OFRs time series
from Basemap and Planet Fusion with PlanetScope .............................................. 136

Figure 7 Monthly percent error variability and MAPE calculated from the same day pairwise
comparisons between Basemap and Planet Fusion with PlanetScope for the three size
classes (0.1–5 ha, 5–10 ha, and 10–30 ha) ............................................................. 137

Figure 8 OFRs (see Table 2) surface area time series derived from PlanetScope, Basemap and
Planet Fusion, and the OFRs shapefile overlaid on high-resolution Google Satellite
imagery ................................................................................................................... 140

Figure 9 OFRs A and B (see Table 2) PlanetScope, Basemap, and Planet Fusion false color
composites (Blue: Red, Green: Green, and Red: NIR), and the surface water
classification for 08/16/2020 (OFR A) and 08/30/2020 (OFR B).......................... 143

Figure 10 OFR E (see Table 2) PlanetScope, Basemap, and Planet Fusion false color composites
(Blue: Red, Green: Green, and Red: NIR), and the surface water classification for
10/16/2020 and 07/14/2020 .................................................................................... 143

Chapter 5

Figure 1 Study region located in eastern Arkansas, USA, the subbasins and surface water
streams and channels delineated with SWAT+, the model outlet, the United States
Geological Survey (USGS) stations (USGS, 2022) used for flow calibration and
validation, the digitized OFRs (Yaeger et al., 2017), and the Digital Elevation Model
(DEM) used in the modeling (Farr et al., 2007) ..................................................... 163

Figure 2 Flow calibration and validation time series for the three USGS stations A (07264000),
B (07263555) and C (07263580). See Fig. 1 and Table 1 for more information about
the USGS stations. The precipitation time series represents the monthly accumulated
precipitation at the watershed scale (i.e., for the entire study region) .................... 171

Figure 3 OFRs surface area distribution for the three surface area scenarios, lower, mean, and
upper ....................................................................................................................... 172

Figure 4 Monthly percent change in flow between the baseline scenario (vertical dotted blue
line) and the three surface area scenarios (lower, mean, and upper), and for the four

xiii
flow classes (1) 0.001–0.25 m3/s, (2) 0.25–0.50 m3/s, (3) 0.50–2.11 m3/s, and (4)
2.11–17.50 m3/s ...................................................................................................... 174

Figure 5 Kernel density estimation plots smoothed using a Gaussian kernel for the monthly
percent change in flow between the baseline scenario (vertical dotted blue line) and
the three surface area scenarios (lower, mean, and upper) for the four flow classes (1)
0.001–0.25 m3/s, (2) 0.25–0.50 m3/s, (3) 0.50–2.11 m3/s, and (4) 2.11–17.50 m3/s.
Note the different range of values of the y-axis for all four flow classes .............. 176

Figure 6 Percent change in peak flow between the baseline scenario (vertical dotted blue line)
and the three surface area scenarios (lower, mean, and upper) for the four flow classes
(1) 0.001–0.25 m3/s, (2) 0.25–0.50 m3/s, (3) 0.50–2.11 m3/s, and (4) 2.11–17.50
m3/s. ........................................................................................................................ 178

Figure 7 A subset of the time series of simulated flow for baseline and the three surface area
scenarios (lower, mean, and upper) between 1995 and 2005 for a selected channel
within the flow class 3 ............................................................................................ 180

xiv
CHAPTER 1: INTRODUCTION

Lakes and reservoirs cover a small fraction (<4%) of Earth’s surface; however, they store

most of the readily available fresh water for human consumption (Downing et al., 2006; Verpoorter

et al., 2014). These water bodies have essential roles in ecosystems and human and animal habitats

(Khazaei et al., 2022; Verpoorter et al., 2014). In particular, OFRs—artificial water bodies that

retain water from rainfall and run-off—enable farmers to store water during the raining season to

support irrigation activities during the crop growing season. Current estimates show that OFRs

span 6% of farmlands worldwide (Downing, 2010), and there are more than 2.6 million OFRs in

the USA alone (Downing, 2010; Renwick et al., 2005)—this number is expected to increase 1–2%

annually in the agricultural parts of the country (Downing et al., 2006).

Eastern Arkansas, USA, is the third most irrigated region in the country, and underlying

this region is the Mississippi Alluvial Aquifer—which partially explains the steady expansion of

irrigated agriculture in the region during the past decades (Lovelace et al., 2015; Marston et al.,

2015). The increase in irrigated land—common irrigated crops include rice and soybeans (NASS–

USDA, 2021)—coincides with critical declines in groundwater resources (Tacker et al., 2010;

Marston et al., 2015). One of the proposed solutions to alleviate the aquifer depletion was the

construction of OFRs, and for this reason, eastern Arkansas has seen a significant increase in the

number of these water bodies during the last 40 years (Shults et al., 2020; Yaeger et al., 2018).

Recently, a remote sensing inventory was carried out to map more than 750 OFRs constructed in

critical zones of the aquifer (Yaeger et al., 2017). The inventory was relevant to understand the

spatial distribution and multi-year changes of these water bodies—more than 400 OFRs were

constructed between 1995 and 2005, and the OFRs covered ~11,300 ha of agricultural land in 2015

(Yaeger et al., 2018, 2017).

1
While the OFRs play an important role in food production through irrigation activities,

OFRs can contribute to downstream water stress by decreasing streamflow and peak flow in the

watersheds where they are built. A recent review (Habets et al., 2018) gathered information from

more than 30 modeling studies conducted in multiple countries (e.g., USA, France, Brazil) to

assess the OFRs impacts on surface hydrology. The authors found that the modeled results vary

significantly between the studies, and the annual impact on streamflow, which ranged between 0.2

and 36% (Habets et al., 2018), was the most common variable reported by the analyzed studies.

Furthermore, the variability of the OFRs impacts increased when evaluating the low flow periods,

with mean decrease on streamflow ranging between 0.3 and 60%. Overall, the estimated mean

annual reduction in flow was 13.4% ± 8.0%, and the mean decrease in peak flow could be as high

as 45% (Habets et al., 2018).

With the increase pressure on surface water resources intensified by climate change and

population growth (Vörösmarty et al., 2010), it is pivotal to elucidate the OFRs dynamic changes

(e.g., filling and drying), and how these changes are related to the OFRs impacts on surface water

availability. Nonetheless, the OFRs are poorly monitored or not monitored in many countries,

mostly because they have a changing nature, they occur in high numbers, and the vast majority of

them are located on private properties (Berg et al., 2016). In addition, the use of incorrect

information on OFRs’ size, spatial distribution, and inter- and intra-annual surface area change for

water allocation plans can have detrimental consequences to the farmers relying on these OFRs

for their livelihood. Under this scenario, Earth Observation data (e.g., satellite imagery) allow us

to amend the monitoring of OFRs, and to further understand their impacts on the local and remote

environment where they occur, hence, to improve surface water resources management.

2
Even though the OFRs inventory (Yaeger et al., 2017) was carried out in eastern Arkansas,

this inventory represents a snapshot of the OFRs, and there is no research assessing the OFRs

dynamic surface area changes using satellite imagery in this region. When monitoring the OFRs

surface area changes, several studies have used the long-term (>25–30 years) Landsat inundation

analysis ready datasets to assess the spatial and temporal variability of OFRs (Arvor et al., 2018;

Jones et al., 2017; Ogilvie et al., 2018; Yao et al., 2019). However, these datasets—for example,

the Dynamic Surface Water Extent (DSWE) provided by the United States Geological Survey

(USGS) (Jones, 2019), and the Global Surface Water Extent provided by the European Joint

Research Centre (JRC) (Pekel et al., 2016)—have not been used in large scale assessment of OFRs

in the USA. In addition, the recent development and availability of freely accessible (e.g., Sentinel

1 [10 m] and Sentinel 2 [10 m]), and commercial (e.g., PlanetScope [3 m] and RapidEye [5 m])

satellite imagery have opened an array of opportunities to widespread monitoring OFRs at sub-

weekly scale. These imagery sources help addressing some of the key limitations of the Landsat-

based datasets, including the irregular intervals (e.g., at least 16-day revisit time) and its relatively

coarse spatial resolution (30 m)—given that OFRs area small water bodies (< 50 ha) the spatial

resolution of the imagery plays a key factor when assessing the OFRs surface area (Ogilvie et al.,

2020, 2018; Vanthof and Kelly, 2019). Nonetheless, there are no studies combining sensors of

different spatial resolution (i.e., PlanetScope, RapidEye, Sentinel 2 and Sentinel 1), and accessing

the uncertainties of each sensor when they are used to monitor OFRs surface area changes.

To assess the OFRs’ impacts on surface water resources, it is necessary to understand the

OFRs’ spatial and temporal variability (e.g., this can be done using satellite imagery), and the

impacts of individual OFRs as well as the interaction effects of multiple OFRs in a watershed (e.g.,

cumulative impacts), which can be achieved using hydrological models (Canter and Kamath, 1995;

3
Habets et al., 2018). Hydrological modeling is the most common approach to assess the OFRs’

impacts, and the models usually have three main components: the OFRs water balance, a

quantitative approach to simulate OFR inflows and outflows, and the spatial representation of the

OFRs network. However, there is a limited number of studies combining remote sensing derived

information (e.g., OFRs surface area change) with hydrological modeling (Ni and Parajuli, 2018;

Yongbo et al., 2014; Zhang et al., 2012), and there is no hydrological assessment of the how the

OFRs are impacting the surface hydrology in eastern Arkansas.

This dissertation explores various remote sensing technologies, including novel approaches

and imagery datasets, in combination with hydrological modeling to further understand the OFRs

occurrence and surface area changes in eastern Arkansas aiming to support surface water resources

management. This dissertation is presented as four manuscripts, each its own chapter, three of

which have been published in peer-reviewed journals, Agricultural Water Management (Chapter

2), Remote Sensing of Environment (Chapter 3), Remote Sensing (Chapter 4), and one that is

currently under clearance for submission to Water Resources Research (Chapter 4). These

manuscripts focus on independent but related research questions. The overarching goal of this

dissertation is twofold: i) to propose reproducible workflows to improve OFRs monitoring that

harnesses analysis ready datasets and satellite imagery from multiple sensors (e.g., freely available,

and commercial); and ii) to combine satellite based information with hydrological modeling to

assess the OFRs’ impacts on surface hydrology with the aim of supporting more efficient

management of OFRs, and mitigating their downstream impacts. Briefly, each chapter addresses

the following topics:

• Chapter 2: Explores the use of long-term (>25–30 years) Landsat inundation analysis ready

datasets (DSWE and JRC) to assess the OFRs surface area dynamics.

4
• Chapter 3: Proposes a new multi-sensor satellite imagery approach based on the Kalman

filter to assimilate data from optical (PlanetScope, RapidEye and Sentinel 2) and radar

(Sentinel 1) remote sensing.

• Chapter 4: Explores the use of the next generation high spatial (3m) and temporal (daily)

analysis ready datasets, Planet Fusion and Planet Basemap, to monitor OFRs while

highlighting the advantages and limitations of each dataset.

• Chapter 5: Proposes a new framework to combine satellite-based information with

hydrological modeling using the latest developments of the Soil and Water Assessment

Tool+ (SWAT+) hydrological model.

Chapter 2 explores the DSWE and JRC Landsat inundation datasets to assess the OFRs

spatial and temporal changes in surface area where there is insufficient OFRs monitoring. Both

datasets enabled us to describe the OFRs surface area seasonality, and to assess how climate

variables (e.g., precipitation and air temperature) impact the OFRs surface area changes. Overall,

the OFRs are filled in March, April, and May (e.g., when the region receives most of its

precipitation), and the OFRs are found to be at the lowest capacities ranges during the June, July

and August, when the highest temperatures are recorded, and most farmers are irrigating their

crops. When assessing the accuracy of both datasets, the DSWE had smaller mean percent error,

and higher agreement with the validation dataset when compared to the JRC. This study represents

the first assessment of these datasets to monitor OFRs in the USA. In addition, it highlighted

several constraints that should be accounted for before employing these Landsat-based datasets to

other study regions. These limitations include missing information due to cloud obscuration and

sensor-related issues, which have direct impact on the accuracy of the seasonal information, the

unsuitability of the datasets’ spatial resolution (30 m) to monitor OFRs smaller than 5 ha, given

5
that the surface area uncertainties can be higher than 20% for these OFRs sizes, and the 16-day

repeat cycle that may miss rapid filling and drying events that may occur on a shorter timescale.

To overcome the limitations of Landsat-based inundation datasets, and to improve the

OFRs monitoring, Chapter 3 introduced a multi-sensor satellite imagery approach that combines

data from multiple sensors that have higher spatial and temporal resolutions when compared to

Landsat, therefore, decreasing the latency of the satellite observations, and increasing the accuracy

of the OFRs’ surface area estimates.

Chapter 3 offers a novel approach based on multi-sensor satellite imagery and the Kalman

filter to assimilate information from optical (PlanetScope, RapidEye, Sentinel-2), and radar

(Sentinel-1) satellite imagery to monitor OFRs sub-weekly (i.e., every 3 days) surface area

changes. In addition, the uncertainties of each sensor are calculated by comparing the OFRs’

surface area derived from each sensor to an independent validation dataset—the validation dataset

was created using high spatial resolution (30 cm) SkySat satellite imagery. The algorithm enabled

the assimilation of optical and radar satellite data while accounting for the uncertainties in both

the satellite observations and their estimates. In addition, the algorithm is not sensor dependent,

and data from future upcoming satellites could be used in combination with these sensors to

improve the OFRs’ time series. The Kalman filter uses a Bayesian inference approach to estimate

the surface area uncertainties, which reflects the quality of the surface area estimates. This is an

improvement over other sensor-specific algorithms (Gao et al., 2006; Houborg and McCabe, 2018;

Zhu et al., 2010) that do not provide uncertainty estimates for the output values. Overall, all sensors

tended to underestimate the area of the OFRs when compared to the validation dataset, and the

uncertainties decreased as the OFRs increased in size, and for the OFRs with regular geometries

(e.g., fewer edges, and less fragmented shape). Likewise, the algorithm tended to underestimate

6
the OFRs surface area, with smaller uncertainties when only the optical sensors were included in

the Kalman filter—Sentinel 1 had the highest uncertainties between all sensors, with mean percent

error higher than 20%. Improving the OFRs’ surface area observations cadence, and estimating

the surface area uncertainties has the potential to improve water conservation plans by allowing

better assessment and management of OFRs. The algorithm allowed further understanding of the

spatial and temporal variability from a network of OFRs, which could enable water agencies to

monitor water use trends, and to act on management recommendations aiming to increase water

use efficiency.

The algorithm proposed in Chapter 3 required the download and processing of more than

400,000 satellite image clips from multiple platforms (e.g., Google Earth Engine, SentinelHub),

which can be a demanding task, and a limiting factor when implementing the proposed

methodology. In addition, although data from upcoming missions could be added to the existent

time series, one should be aware that this integration is contingent on having the uncertainties of

the new observations, therefore, requiring an updated validation dataset. An alternative to the

multi-sensor algorithm proposed in Chapter 3, is to employ the high spatial (3 m) and temporal

(daily) analysis ready datasets, Planet Fusion and Planet Basemap, which were further explored in

Chapter 4.

Chapter 4 demonstrates the use of the next generation analysis ready datasets, Planet

Fusion and Planet Basemap. Both datasets are based PlanetScope imagery, however, Planet Fusion

is generated using the CubeSat Enabled Spatiotemporal Enhancement Method (Houborg and

McCabe, 2018), and it leverages rigorously calibrated publicly available multispectral satellites

(e.g., Sentinel 2, Landsat, MODIS). Meanwhile, Planet Basemap is generated by mosaicking the

whole or part of highest quality PlanetScope imagery, which is selected based on a set of rules, for

7
instance, cloud cover, image acutance, and acquisition date. We assessed the usefulness of these

datasets to monitor the OFRs’ daily surface area changes, and by comparing the datasets while

describing their differences. Both datasets had small uncertainties (<10%) when compared to an

independent validation dataset (i.e., the same based on SkySat used in Chapter 3). Overall, Planet

Basemap presented higher surface area variability, and it was more susceptible to the presence of

cloud shadows and haze when compared to Planet Fusion, which had a smoother time series with

less variability and fewer abrupt changes in surface area throughout the year. On the other hand,

Planet Fusion is based on an algorithm that uses data from several satellites (e.g., Landsat 8,

Sentinel-2), and therefore, when this dataset is generated by Planet Labs, it requires extra image

processing steps, and higher computing power when compared to Planet Basemap. Although these

datasets do not require the users to process different types (e.g., optical and radar) of satellite

imagery with different spatial resolution from multiple platforms—which can be a demanding task

and a limiting factor when processing, downloading, moving data across multiple platforms—they

are proprietary, and they are not available free of costs. In addition, before employing these

datasets for monitoring purposes, it is recommended that users cautiously assess the OFRs time

series, as the OFRs surface area classifications may be impacted by the OFRs environmental

conditions (e.g., presence of vegetation inside the OFR). Despite their proprietary use, the on-

going partnerships between Planet Labs and government and research institutions, will make these

datasets more accessible to researchers across the globe. The amount of detailed information (i.e.,

daily and at 3 m spatial resolution) combined with minimal user processing steps, represent the

next generation of analysis ready datasets, from which users are able to quickly extract insights

(e.g., OFRs surface area seasonality), and insert into their platforms and workflows.

8
Chapters 2, 3, and 4 represent the efforts to fulfill the first part of the overarching goal of

this dissertation, which is to improve OFRs monitoring leveraging different techniques and

satellite imagery sources while using reproducible workflows that could be applied to other study

regions. Chapter 5 is dedicated to the second part of the overarching goal, and uses the OFRs

surface area information derived from Kalman filter in Chapter 3 in combination with the SWAT+

hydrological model to estimate the OFRs’ impacts on surface water availability.

Chapter 5 proposes a new framework to evaluate the OFRs’ impacts on surface hydrology

using the latest SWAT+ model developments while accounting for the OFRs surface area

variability derived from the Kalman filter time series. The results indicate that the OFRs do not

have an equally distributed impact on flow and peak flow across the watershed. In addition, the

impacts have an intra- and inter-annual variability that will change according to different OFRs

and water channels. In general, the largest impacts on flow occurred during the first part of the

year, between January and May—the period of the year when the region receives most of its

precipitation, hence, when the OFRs are storing their maximum amount of water, and when the

peak flows occur. The modeled results showed that the presence of the OFRs on the watershed

could decrease annual flow between 14 and 24%, and the mean reduction in peak flow may vary

between 43 and 60%. In this study, a new framework is proposed based on combining remote

sensing derived information with hydrological modeling to quantitatively analyze the impact of a

network of OFRs on flow and peak flow. In addition, the various potential reasons behind the

variability of the impacts are discussed as well as the assumptions and limitations of this

framework. Finally, it is expected that this methodology will be refined in future studies, and that

it can be used in other watersheds to support water agencies with information to improve surface

water resources management.

9
This dissertation provides tools for water authorities and managers to monitor OFRs

occurrence and their inter- and intra-annual surface area variability, both using analysis ready

datasets and freely available and commercial satellite imagery. Beyond monitoring the OFRs, a

new framework is proposed to quantitatively assess the OFRs impacts on surface hydrology.

Although this work is focused on eastern Arkansas, the developed methods could be applied to

other study regions. Ultimately, the efforts outlined in this dissertation have implications for

surface water management, as the methods can be used to assess regions under high pressure on

water availability, and identify the areas that could benefit from the constructions of new OFRs,

targeting better management of surface water resources and irrigation activities.

10
REFERENCES

Arvor, D., Daher, F.R.G., Briand, D., Dufour, S., Rollet, A.-J., Simões, M., Ferraz, R.P.D., 2018.
Monitoring thirty years of small water reservoirs proliferation in the southern Brazilian
Amazon with Landsat time series. ISPRS J. Photogramm. Remote Sens. 145, 225–237.
https://doi.org/10.1016/j.isprsjprs.2018.03.015

Berg, M.D., Popescu, S.C., Wilcox, B.P., Angerer, J.P., Rhodes, E.C., Mcalister, J., Fox, W.E.,
2016. Small farm ponds: Overlooked features with important impacts on watershed
sediment transport. J. Am. Water Resour. Assoc. 52, 67–76.
https://doi.org/10.1111/1752-1688.12369

Canter, L.W., Kamath, J., 1995. Questionnaire checklist for cumulative impacts. Environ. Impact
Assess. Rev. 15, 311–339. https://doi.org/10.1016/0195-9255(95)00010-C

Downing, J.A., 2010. Emerging global role of small lakes and ponds: little things mean a lot.
Limnetica 29, 9–24. https://doi.org/10.23818/limn.29.02

Downing, J.A., Prairie, Y.T., Cole, J.J., Duarte, C.M., Tranvik, L.J., Striegl, R.G., McDowell,
W.H., Kortelainen, P., Caraco, N.F., Melack, J.M., Middelburg, J.J., 2006. The global
abundance and size distribution of lakes, ponds, and impoundments. Limnol. Oceanogr.
51, 2388–2397. https://doi.org/10.4319/lo.2006.51.5.2388

Gao, F., Masek, J., Schwaller, M., Hall, F., 2006. On the blending of the Landsat and MODIS
surface reflectance: predicting daily Landsat surface reflectance. IEEE Trans. Geosci.
Remote Sens. 44, 2207–2218. https://doi.org/10.1109/TGRS.2006.872081

Habets, F., Molénat, J., Carluer, N., Douez, O., Leenhardt, D., 2018. The cumulative impacts of
small reservoirs on hydrology: A review. Sci. Total Environ. 643, 850–867.
https://doi.org/10.1016/j.scitotenv.2018.06.188

Houborg, R., McCabe, M.F., 2018. A Cubesat enabled Spatio-Temporal Enhancement Method
(CESTEM) utilizing Planet, Landsat and MODIS data. Remote Sens. Environ. 209, 211–
226. https://doi.org/10.1016/j.rse.2018.02.067

Jones, J.W., 2019. Improved automated detection of subpixel-scale inundation-revised Dynamic


Surface Water Extent (DSWE) partial surface water tests. Remote Sens. 11.
https://doi.org/10.3390/rs11040374

Jones, S.K., Fremier, A.K., DeClerck, F.A., Smedley, D., Pieck, A.O., Mulligan, M., 2017. Big
data and multiple methods for mapping small reservoirs: Comparing accuracies for
applications in agricultural landscapes. Remote Sens. 9.
https://doi.org/10.3390/rs9121307

Khazaei, B., Read, L.K., Casali, M., Sampson, K.M., Yates, D.N., 2022. GLOBathy, the global
lakes bathymetry dataset. Sci. Data 9, 36. https://doi.org/10.1038/s41597-022-01132-9

11
L Tacker, P., D Vories, E., Carman, D., 2010. Water Supply Problems and Solutions in the
Grand Prairie Region of Arkansas, in: 5th National Decennial Irrigation Conference
Proceedings, 5-8 December 2010, Phoenix Convention Center, Phoenix, Arizona USA.
ASABE, St. Joseph, MI. https://doi.org/10.13031/2013.35853

Lovelace, J.K., Nielsen, M.G., Read, A.L., Murphy, C.J., Maupin, M.A., 2015. Estimated
groundwater withdrawals from principal aquifers in the United States, 2015. US
Geological Survey. https://doi.org/doi.org/10.3133/cir1464

Marston, L., Konar, M., Cai, X., Troy, T.J., 2015. Virtual groundwater transfers from
overexploited aquifers in the United States. Proc. Natl. Acad. Sci. U. S. A. 112, 8561–
8566. https://doi.org/10.1073/pnas.1500457112

National Agricultural Statistics Service–U.S. Department of Agriculture (NASS–USDA), 2017.


Census of agriculture. Washington, DC. https://www.nass.usda.gov/AgCensus/

Ni, X., Parajuli, P.B., 2018. Evaluation of the impacts of BMPs and tailwater recovery system on
surface and groundwater using satellite imagery and SWAT reservoir function. Agric.
Water Manag. 210, 78–87. https://doi.org/10.1016/j.agwat.2018.07.027

Ogilvie, A., Belaud, G., Massuel, S., Mulligan, M., Le Goulven, P., Malaterre, P.-O., Calvez, R.,
2018. Combining Landsat observations with hydrological modelling for improved surface
water monitoring of small lakes. J. Hydrol. 566, 109–121.
https://doi.org/10.1016/j.jhydrol.2018.08.076

Ogilvie, A., Poussin, J.-C., Bader, J.-C., Bayo, F., Bodian, A., Dacosta, H., Dia, D., Diop, L.,
Martin, D., Sambou, S., 2020. Combining Multi-Sensor Satellite Imagery to Improve
Long-Term Monitoring of Temporary Surface Water Bodies in the Senegal River
Floodplain. Remote Sens. 12, 3157. https://doi.org/10.3390/rs12193157

Pekel, J.F., Cottam, A., Gorelick, N., Belward, A.S., 2016. High-resolution mapping of global
surface water and its long-term changes. Nature 540, 418–422.
https://doi.org/10.1038/nature20584

Renwick, W.H., Smith, S.V., Bartley, J.D., Buddemeier, R.W., 2005. The role of impoundments
in the sediment budget of the conterminous United States. Geomorphology 71, 99–111.
https://doi.org/10.1016/j.geomorph.2004.01.010

Shults, D.D., Nowlin, W.J., Yaeger, M.A., Massey, J.H., Reba, M.L., 2020. A spatiotemporal
analysis quantifying the need for more on-farm reservoirs to reduce groundwater use in
the Cache and L’Anguille River Regions in Northeastern AR, in: ESRI User Conference.
Online.

Vanthof, V., Kelly, R., 2019. Water storage estimation in ungauged small reservoirs with the
TanDEM-X DEM and multi-source satellite observations. Remote Sens. Environ. 235,

12
111437. https://doi.org/10.1016/j.rse.2019.111437

Verpoorter, C., Kutser, T., Seekell, D.A., Tranvik, L.J., 2014. A global inventory of lakes based
on high-resolution satellite imagery. Geophys. Res. Lett. 41, 6396–6402.
https://doi.org/10.1002/2014GL060641

Vörösmarty, C.J., McIntyre, P.B., Gessner, M.O., Dudgeon, D., Prusevich, A., Green, P.,
Glidden, S., Bunn, S.E., Sullivan, C.A., Liermann, C.R., Davies, P.M., 2010. Global
threats to human water security and river biodiversity. Nature 467, 555–561.
https://doi.org/10.1038/nature09440

Yaeger, M.A., Massey, J.H., Reba, M.L., Adviento-Borbe, M.A.A., 2018. Trends in the
construction of on-farm irrigation reservoirs in response to aquifer decline in eastern
Arkansas: Implications for conjunctive water resource management. Agric. Water
Manag. 208, 373–383. https://doi.org/10.1016/j.agwat.2018.06.040

Yaeger, M.A., Reba, M.L., Massey, J.H., Adviento-Borbe, M.A.A., 2017. On-farm irrigation
reservoirs in two Arkansas critical groundwater regions: A comparative inventory. Appl.
Eng. Agric. 33, 869–878. https://doi.org/10.13031/aea.12352

Yao, F., Wang, J., Wang, C., Crétaux, J.-F., 2019. Constructing long-term high-frequency time
series of global lake and reservoir areas using Landsat imagery. Remote Sens. Environ.
232, 111210. https://doi.org/10.1016/j.rse.2019.111210

Yongbo, L., Wanhong, Y., Zhiqiang, Y., Ivana, L., Jim, Y., Jane, E., Kevin, T., 2014. Assessing
Effects of Small Dams on Stream Flow and Water Quality in an Agricultural Watershed.
J. Hydrol. Eng. 19, 05014015. https://doi.org/10.1061/(ASCE)HE.1943-5584.0001005

Zhang, C., Peng, Y., Chu, J., Shoemaker, C.A., Zhang, A., 2012. Integrated hydrological
modelling of small- and medium-sized water storages with application to the upper
Fengman Reservoir Basin of China. Hydrol Earth Syst Sci 15.
https://doi.org/10.5194/hess-16-4033-2012

Zhu, X., Chen, J., Gao, F., Chen, X., Masek, J.G., 2010. An enhanced spatial and temporal
adaptive reflectance fusion model for complex heterogeneous regions. Remote Sens.
Environ. 114, 2610–2623. https://doi.org/10.1016/j.rse.2010.05.032

13
CHAPTER 2: ON-FARM RESERVOIR MONITORING USING LANDSAT

INUNDATION DATASETS

Published in Agricultural Water Management January 2021


(https://doi.org/10.1016/j.agwat.2020.106694)

Authors
Vinicius Perin1, Mirela G. Tulbure1, Mollie D.Gaines1, Michele L.Reba2, Mary A.Yaeger3

Affiliations
1
Center for Geospatial Analytics, North Carolina State University, 2800 Faucette Drive, Raleigh,
NC, 27606, USA
2
USDA-Agricultural Research Service, Delta Water Management Research Unit, Jonesboro, AR,
72467, USA
3
Center for Applied Earth Science and Engineering Research, The University of Memphis, 3675
Alumni Drive, Memphis, TN, 38152, USA

Acknowledgements
Funding for this work came from a North Carolina State University start-up package to Tulbure.

Abstract
On-farm reservoirs (OFRs)—artificial water impoundments that retain water from rainfall and run-

off—enable farmers to store water during the wet season to be used for crop irrigation during the

dry season. However, monitoring the inter- and intra-annual change of these water bodies remains

a challenging task because they are typically small (< 10 ha) and occur in high numbers. Therefore,

we used two existing Landsat inundation datasets—the U.S. Geological Survey Dynamic Surface

Water Extent (DSWE) and the European Commission’s Joint Research Centre (JRC) Global

Monthly Water History—to assess surface water area change of OFRs located in eastern Arkansas,

the third most irrigated state in the U.S. that has seen a rapid increase of OFRs occurrence. We

14
used an existent OFRs dataset as ground-truth. We aimed (i) to compare the performance of the

DSWE and the JRC when characterizing OFRs of varied sizes and (ii) to assess the impact of

climate variables (i.e., precipitation and temperature) on surface water area of OFRs. We found

the highest mean percent errors (MPE) in size (~20%) for OFRs between 0 and 5 ha, the smallest

size class in our study. The DSWE had a smaller MPE and higher agreement with our ground-truth

dataset when compared to the JRC for OFRs smaller than 5 ha (p-value < 0.05). Both inundation

datasets enabled us to estimate the seasonality in surface area change of OFRs, with the highest

surface water extent between March–May, the months when the region receives most of the annual

precipitation. Our results showed that both DSWE and JRC can be used to enhance hydrological

assessments in poorly monitored basins that have a concentration of OFRs, and the methods can

be applied to other study regions if the inundation datasets are available.

Keywords: Landsat, Irrigation, Water management, On-farm reservoir

15
1. INTRODUCTION

On-farm reservoirs enable farmers to store water during the wet season and support water

use during the dry season. On-farm reservoirs are particularly important for crop irrigation.

Estimates show that these water bodies span 6% of farmlands worldwide (Downing, 2010), and

their numbers are projected to increase 1–2% annually in the agricultural parts of the U.S.

(Downing et al., 2006).

On-farm reservoirs can impact the ecosystems where they are built. While the impact of an

individual OFR is bound to the local environment, the cumulative impact of a network of OFRs

extends to larger spatial scales (Berg et al., 2016; Habets et al., 2018; Mime and Young, 1989;

Smith, 1998). Previous studies reported that a set of OFRs can decrease annual stream discharge

between 0.2% (Hughes and Mantel, 2010) and 36% (Habets et al., 2018; Meigh, 1995). On-farm

reservoirs are usually small (<10 ha) and occur in higher numbers than large (>1000 ha) water

impoundments (Downing, 2010). However, most studies are focused on understanding the

hydrological impact of large water bodies (Adams and Hughes, 1986; Barrow, 1987; Dudgeon,

2000; Fearnside, 2006, 2001; Maingi and Marsh, 2002; Martins and Meurer, 2009), and end up

overlooking the collective effect of OFRs (Arvor et al., 2018; Habets et al., 2014). On-farm

reservoirs are poorly monitored or not monitored in many countries, mostly because they have a

changing nature, they occur in high numbers, and the vast majority of them are located on private

properties (Berg et al., 2016). The use of incorrect information on OFRs’ size, spatial distribution,

and inter- and intra-annual area change for water allocation plans can have detrimental

consequences to food production and to the farmers relying on these OFRs for their livelihood.

Earth observation data allow us to amend the monitoring of OFRs, which can then further our

understanding of their impacts on the local environment where they occur.

16
Water inundation datasets provide insights on surface water distribution and capture the

anthropogenic impacts on freshwater resources (Pekel et al., 2016). Currently, two products

leverage the extended (>30 years) Landsat mission archive: the DSWE provided by the United

States Geological Survey (USGS) (Jones, 2019, 2015) and the Global Surface Water Extent

provided by the European Joint Research Centre (JRC) (Pekel et al., 2016). Studies have used both

datasets for applications in water management (Ahmad et al., 2020; Avisse et al., 2017; Ogilvie et

al., 2018; Soulard et al., 2020), and a recent study compared their performance on integrating

satellite data and streamflow volume measurements to fill gaps in monthly surface water

chronologies (Walker et al., 2020). Moreover, several studies harnessed the capability of the JRC

to derive water management insights for OFRs in water basins of India (Vanthof and Kelly, 2019),

Brazil (Zhang et al., 2016), and the Volta Basin in Africa which intersects several countries,

including: Benin, Burkina Faso, Côte d'Ivoire, Mali, Togo and Ghana, (Jones et al., 2017).

Nonetheless, both datasets remain untested in large-scale assessment of OFRs in the U.S. In

addition, although the DSWE is potentially more accurate than JRC when applied to study regions

in the U.S.—since the dataset is country-specific—there has been no comparison of both datasets

for assessing changes in OFRs. Therefore, measuring the level of uncertainty of these water

inundation products when compared to ground-truth, an independently derived dataset of OFRs,

helps to ensure their effective use for water planning policies.

Precipitation and air temperature play a key role in controlling the evaporation losses from

OFRs, especially in dry critical periods of the year (Fowe et al., 2015; Liebe et al., 2005), when

there is less water available from precipitation and higher net solar radiation, resulting in more

energy for evaporation losses (Wang et al., 2018). A recent study in semi-arid Brazil showed that

future scenarios of climate change will likely have negative impact on OFRs water availability;

17
the study highlighted that an increase of ~7–18% in evaporated water volumes by the end of the

next century would decrease OFRs water availability by 4–10% (Althoff et al., 2020). Given the

relevance of climate variables (i.e., precipitation and air temperature) on governing the OFRs

dynamics, there is a need to unravel their impact on OFRs water availability when aiming to

develop strategic policies for water management.

The Mississippi Alluvial Aquifer—located in eastern Arkansas—is one of the most

exploited aquifers in the U.S. for irrigation purposes (Lovelace et al., 2015; Marston et al., 2015).

In addition, this is the third most irrigated region in the country and common cultivated crops are

rice, soybeans, corn and cotton (NASS–USDA, 2017). The rapid expansion of irrigated agriculture

coincides with substantial declines in groundwater resources (Tacker et al., 2010; Marston et al.,

2015). A proposed solution to alleviate the aquifer depletion was the construction of OFRs. For

this reason, eastern Arkansas has seen a significant increase in the number of these water bodies

during the last 40 years (Shults et al., 2020; Yaeger et al., 2018). A remote sensing inventory

derived from digitized orthoimages manually mapped more than 750 OFRs constructed in critical

regions of the aquifer (Yaeger et al., 2018, 2017). The inventory was important for mapping the

spatial distribution and multi-year changes of these water bodies. The authors found that more than

400 OFRs were constructed between 1995 and 2005 in critical regions of the Mississippi Alluvial

Aquifer. In addition, these OFRs covered ~11,300 ha of agricultural land in 2015 (Yaeger et al.,

2018, 2017). The inventory was the first step towards a necessary surface water assessment in the

region. However, further exploration on inter- and intra-annual area change and seasonality of

OFRs as well as the impact of climate variables on OFR is needed to improve existing models

used to calculate the Mississippi Alluvial Aquifer water budget (Clark et al., 2011).

18
We created time series of OFRs inundation areas based on the DSWE and JRC datasets to

provide information on the OFRs area temporal variability, and to assess their inter- and intra-

annual surface area change and seasonality. We aimed to (i) compare the performance of both

datasets when characterizing OFRs of varied sizes using the remote sensing inventory as ground-

truth (Yaeger et al., 2017) and (ii) to assess the impact of climate variables (i.e., precipitation and

temperature) on surface water area of OFRs.

2. MATERIAL AND METHODS

2.1 Study location


The study region is located in eastern Arkansas, U.S.—a region with high water demand

due to crop irrigation. Future projections estimate that irrigated crop water demand will increase

from 33.4 Mm3 d-1 to 37.8 Mm3 d-1 in 2050 (Arkansas Drought Planning: Summary Report, 2016).

Annual USGS reports in monitored wells showed a steady decline in water levels varying between

-0.64 m year-1 to 0 m year-1 from 1998 to 2012 for counties located in the study region (Schrader,

2015). The spatial extent of this study comprises the major 8-digit hydrologic unit codes (HUCs)

(Table 1) that intersect critical ground water regions in eastern Arkansas (Fig. 1). The 30-year

annual average precipitation is ~1,300 mm yr-1. Annual precipitation is moderately seasonal and

distributed mostly between March and May, adding up to a total ~400 mm during this period. The

average temperature is 29 °C. On average, August is the driest month, receiving about ~80 mm yr-
1
of precipitation (PRISM Climate Group, 2011).

19
Figure 1–Study region in eastern Arkansas, U.S. The study region encompasses the main HUCs

watersheds and the OFRs ground-truth dataset. Inset rectangles depict: (1) JRC (Pekel et al., 2016),

(2) DSWE (Jones et al., 2015, 2019), and (3) Potential irrigated crops from CropScape, Cropland

Data Layer (NASS–USDA, 2020a). The inset rectangles also contain the OFRs. For HUC ID

information please see Table 1.

2.2 On-farm reservoirs ground-truth dataset


We used the remote sensing inventory carried out by Yaeger et al., (2017) as ground-truth.

This inventory estimated the number, surface area, and spatial distribution of the OFRs in the

critical groundwater areas in eastern Arkansas. The authors utilized county-level images provided

by the National Agricultural Imagery Program. The image mosaics were compressed and geo-

referenced and provided by the USDA-NRCS Geospatial Data Gateway

(https://datagateway.nrcs.usda.gov/). The authors used images acquired at 1-m spatial resolution

20
with accuracy of ~6m of ground control points. They manually digitized the OFRs by combining

2015 Google Earth satellite imagery with the National Agricultural Imagery Program archive. The

Google Earth Explorer tool allowed them to sharpen the image details when zooming in and

provided a validation for features appearing indistinct or pixelated in the 1-m mosaic imagery. The

Google Earth imagery archive was collected using different sensors (e.g., Landsat-8 and Digital

Globe satellites), with a spatial resolution varying from 15 cm to 15 m, depending on the satellite

source. To further identify OFRs used for irrigation, the authors identified key surrounding

elements of a surface water irrigation system such as tailwater recovery and irrigation regulating

ditches, supply/drainage pipes and/or pumping houses, and one or more adjacent crop fields

(Sullivan and Delp, 2012; Yaeger et al., 2017). Table 1 shows the number of OFRs in each HUC—

more than 500 OFRs are located in two of the eight HUCs that cover our study area.

Table 1–Hydrological Unit Code names and the number of OFRs in each HUC from the ground-

truth dataset.

HUC ID HUC Name Number of OFRs


A 08020402 Bayou Meto 295
B 08020303 Lower White 238
C 08020205 L'anguille 79
D 08020301 Lower White-Bayou Des Arc 78
E 08020302 Cache Arkansas 63
F 08020401 Lower Arkansas 12
G 08020304 Big Arkansas 8
H 11110207 Lower Arkansas -Maumelle 2

Yaeger et al., (2018) expanded the work carried out by Yaeger et al., (2017) by analyzing

the trends in reservoir construction over time. The authors superimposed the 2015 reservoir

contours with the high spatial resolution 1-m ortho-imagery and systematically checked whether

the OFRs were present during specific years (1996, 2000, 2006, 2009, 2010, 2013 and 2015). The

21
temporal analysis resulted in seven reservoir shapefiles, one for each year checked. The intervals

between the OFR shapefiles are not evenly distributed due to the historical catalog of the ortho-

imagery. Figure 2 shows the cumulative distribution of OFR surface water area (ha) and their

construction period. We used the OFRs shapefile as ground-truth data for our analyses (see section

2.5).

Figure 2–The cumulative distribution of OFRs surface area extent (ha) in the ground-truth dataset.

OFRs with surface area higher than 50 ha (3.3% of total) are not shown on this figure to avoid

over plotting.

2.3 Processing satellite imagery time series


We processed the two independent surface inundation datasets derived from the Landsat

imagery catalog: DSWE (v2.0) (Jones et al., 2015, 2019) and JRC Global Monthly Water History

(v1.1) (Pekel et al., 2016). We assembled the monthly image time series starting in January 1995—

one year prior to the first year of the assessment made by Yaeger et al., (2018)—until December

22
2018, last JRC image available on Google Earth Engine (GEE). We chose this time series period

(1995-2018) to coincide with the OFR mapping period of the ground-truth dataset.

Aiming to capture the monthly variability of the vegetated surface surrounding the OFRs,

we combined the Landsat 5 TM, Landsat 7 ETM+, and Landsat 8 OLI images available on GEE

to calculate mean and monthly Normalized Difference Vegetation Index (NDVI) using a 1 km

buffer of each OFR in our ground-truth dataset—we assumed that this buffer distance includes the

area in which the water from the OFR is potentially being used for irrigation. This assumption is

based on the idea that the water from the OFR is being used to irrigate closely located cropland,

and it was not validated with a field study (personal conversation with geospatial specialist Dr.

John Nowlin in eastern Arkansas). The NDVI from the possible irrigated land surrounding the

OFRs can helps us understand the annual monthly change in crop phenology (Seo et al., 2019;

Yang et al., 2013), and therefore, the NDVI may be useful to further understand the OFRs surface

water dynamics by representing anthropogenic factors.

2.3.1 Dynamic Surface Water Extent (v.1.0)


The Level-3 product DSWE (v.1.0) (Jones et al., 2015, 2019) is available online for the

contiguous United States through the USGS Earth Explorer platform

(https://earthexplorer.usgs.gov/), and it is the newest product from the Landsat Analysis Ready

Data series. The DSWE product is generated using the Landsat surface reflectance—the most

geometrically accurate and consistently processed imagery from Landsat 4–5 TM, Landsat 7

ETM+, and Landsat 8 OLI. We recreated analysis ready data equal-sized tiles by downloading the

DSWE images projected to Albers Equal Area Conic. We used the DSWE “Interpreted layer” to

retrieve only the highest confidence water pixel value (i.e., 1-Water: High confidence). The

Interpreted layer has seven diagnostic pixel values, including: 0-Not water, 1-Water High

23
confidence, 2-Water Moderate confidence, 3-Potential wetland, 4-Water or wetland-low

confidence, 9-Cloud, shadow, and snow, and 255-Not available, fill. To match the JRC monthly

water inundation time series, we composited multiple overlapping DSWE scenes. We used a pixel-

based approach to retain the 1-High confidence water classification for co-located pixels. For

example, the final class of a pixel is “water—high confidence” (class 1), if this pixel was classified

as “potential wetland” (class 3) in one image and classified as “water—high confidence” (class 1)

in another one. The final dataset consists of 288 DSWE monthly composites time series from

January 1995 until December 2018.

2.3.2 Joint Research Centre Monthly Water History (v.1.1)


The JRC Global Monthly Water History (v.1.1) (Pekel et al., 2016) has a global coverage

and it is available online for public download through the cloud-computing platform GEE

(https://code.earthengine.google.com/). The Monthly Water History is part of a global collection

of water inundation layers from March 1984 until December 2018. The authors created the

collection using the entire Landsat archive (three million satellite images) combining images from

the Landsat 4–5 TM, Landsat 7 ETM+, and Landsat 8 OLI. The monthly history provides

information on the intra-annual distribution of water bodies and characterizes their seasonality.

The purpose of creating a monthly time series was to show the water distribution over time and

space. This dataset can also be used to flag when an OFR was constructed at a particular location.

The Monthly Water History is available in GEE as an image collection of 380 images; each band

contains a water layer that has three diagnostic pixel values: 0 (No observations), 1 (Not water)

and 2 (Water detected), for a complete description of the water classification methods refer to

Pekel et al., (2016). We downloaded the Monthly Water History from GEE for our study region

24
from 1995 until 2018, and we extracted the surface water area for each reservoir using R v.3.5.1

(R Core Team, 2020).

2.4 Climate information


We downloaded monthly precipitation, minimum and maximum air temperature and vapor

pressure deficit (VPD) (from January 1995 to December 2018) from the interpolated grid maps

available from Oregon State University’s PRISM Climate Group (PRISM Climate Group, 2011)

using each OFR centroid (i.e., latitude and longitude).

2.5 Data analysis


2.5.1 On-farm reservoir size, missing information and reservoir geometry
We used the single time-step surface area (ha) (i.e., OFR size) from the ground-truth dataset

as the OFR capacity. To compare this single surface area with the DSWE and the JRC, we looped

through all time-steps (i.e., 1995–2018) for each OFR to extract the highest surface area recorded

from both inundation datasets—the highest surface area was chosen because it reflects the OFR

capacity according with the inundation datasets. We divided the OFRs into four different size

classes (0–5 ha, 5–10 ha, 10–25 ha, and >25 ha) according with the surface area mapped in the

ground-truth dataset, which helped us understand how the discrepancies between the inundation

datasets and the ground-truth occur for different OFR sizes. Then, we performed a simple linear

regression analysis to calculate the coefficient of determination (r2) and the MPE (Eq. 1) between

each inundation dataset and the ground-truth; hence, we had two linear regressions models–one

for each dataset–for each class size. In addition, we compared the differences in MPE between

DSWE and JRC using one-way analysis of variance (ANOVA).


𝑦𝑖 − 𝑥𝑖
Mean Percent Error = | | * 100 (1)
𝑥𝑖

25
where 𝑥𝑖 is the ground-truth and yi is the highest surface area mapped by each inundation dataset

including all time-steps for each OFR.

For each OFR and both inundation datasets, we created an index to measure the monthly

percentage of missing information (NAindex). This index was calculated by dividing the number of

absent surface area information to the total possible number of observations including all time-

steps—the total possible number of observations depends on the OFR construction year, for

instance, if an OFR was constructed in 2000, the total number of observations will be 216, 12

months per year and 18 years. Then, we calculated NAindex for each class size by averaging the

NAindex of all OFRs included in each class.

We calculated a dimensionless shape index (SI) to describe the complexity of the OFRs

surface water extent (Karran et al., 2017; Vanthof and Kelly, 2019). SI is the ratio of the reservoir

perimeter to the circumference of a circle with the same area—we obtained the perimeter and the

reservoir area from ground-truth dataset. On-farm reservoirs with SI = 1 are perfectly circular while

OFRs with SI > 1 are increasingly complex. The SI is an important metric because OFRs with

complex shapes are more difficult to map, therefore, leading to higher surface area mapping errors

(Karran et al., 2017).

2.5.2 Generalized additive mixed model approach to explain the reservoirs area change
Similar to other studies linking climate variables to satellite-mapped surface water extent

(e.g., Heimhuber et al., 2019, Tulbure and Broich, 2019), we explored the impact of precipitation,

air temperature, and NDVI on the dynamics of OFRs surface area by employing a generalized

additive mixed model approach using the gamm4 R package (Wood and Scheipl, 2014).

Generalized additive mixed models are multivariate regression models that additively combine

parametric (linear) and non-parametric (non-linear) model terms, and allow both fixed and random

26
effects. The wiggliness of splines is penalized when the models are fitted, and the fitting algorithm

aims to maximize the fit of the spline to the data while minimizing its wiggliness (Wood, 2017).

Fixed effects are the factors we expect will have an impact on the dependent variable—these

effects are referred to as explanatory variables. Random effects are categorical variables, and often

grouping factors—these are the effects that we know may influence the patterns in our response

variable.

We used the OFR monthly surface area ratio (SAratio) as the dependent variable. SAratio was

calculated by dividing the monthly OFR surface area extracted from the DSWE and JRC by the

OFR ground-truth area. The SAratio allows us to infer the OFR capacity level in a given month in

terms of area. For instance, a SAratio of 0.7 means that the OFR is at 70% of the maximum capacity.

We aggregated SAratio per HUC boundaries by averaging the SAratio of all OFRs in each HUC for

each time-step (i.e., month) in our period of analysis and separately for the DSWE and JRC

inundation datasets. This level of aggregation resulted in a total of 16 monthly SAratio time series

(i.e., eight HUCs and two inundation datasets) from 1995 to 2018 totaling 276 monthly

observations per HUC per inundation dataset. We examined multicollinearity between the

explanatory variables—precipitation, minimum, maximum and mean air temperature, VPD, and

NDVI—and we computed the Pearson correlation coefficient for each two variables. Following

this procedure, we removed minimum and mean air temperature, and minimum, maximum and

mean VPD. These variables were highly correlated with maximum temperature (correlation

coefficient > 0.90)—this is a common practice as collinearity among predictors can lead to

interpretation problems (see Fig. A1 and A2 for more details). In addition, we matched the

temporal resolution of the explanatory variables with the OFR SAratio time series.

27
Due to the inherent characteristics of the remote sensing inundation datasets (i.e., cloud

obscuration), the SAratio time series can contain abrupt changes that may lead to erroneous

interpretation of the SAratio. These sudden changes are normally attributed to the presence of clouds

and sensor related issues that disrupt the OFR surface area information. To account for these issues,

we applied a simple centered moving average filter with an average smoother of magnitude 2. This

filter allowed us to preserve the SAratio time series information and seasonality, while reducing the

impact of SAratio outliers (Fig. A4). We modelled the SAratio with the moving average using two

modeling approaches: 1) a global model accounting for all HUCs together (i.e., using HUC as a

random effect) for the DSWE and JRC, and 2) per HUC models. The later approach aimed to

measure the model performance within each HUC and to capture the spatial variability of the fixed

effects between different HUCs. More specifically, to account for the per HUC models variability

in slopes of precipitation, maximum temperature and NDVI, and the model performance across

the study region. To minimize the complexity of both modeling approaches, we fitted multiple

models adding each explanatory variable one at a time—we added precipitation, maximum air

temperature and NDVI as fixed effects, and month and year as random effects. We measured the

contribution of each variable accounting for the Akaike Information Criterion (AIC) and adjusted

r2 (Tables A1 and A2). We selected model 3 as the best performing model based on the lowest

AIC and highest adjusted r2, and the least number of parameters. Model 3 includes precipitation,

maximum temperature, and NDVI as fixed effects (Tables A1 and A2).

3. RESULTS AND DISCUSSION

Figure 3 depicts the distribution of minimum, maximum and mean air temperature, NDVI

and precipitation for two HUCs (A and B) in our study region (see Fig. A3 for all other HUCs),

considering the entire time series. In general, most of the precipitation arrives between March and

28
May with a decrease during the months of June to August (Fig. 3, violin plots). Contrastingly, air

temperature has a different behavior, and the highest records occur during summer, between May

and August, and the lowest records between November to February (Fig. 3, black solid line and

red dashed lines). The mean NDVI from the vegetated surface surrounding the OFRs starts to

increase around April and May, which are the months when the main crops are being planted in

the study region. The peak of NDVI occurs between August and September, and the NDVI

decreases during October and November, which typically represent the final crop phenological

stages before harvest.

Figure 3–The distribution of minimum, maximum and mean air temperature, NDVI, and

precipitation for HUCs A and B considering the entire time series, and all OFRs located on these

HUCs. Mean air temperature is represented by the black solid line, and minimum and maximum

air temperature are represented by the red dashed lines. The boxplots represent the NDVI monthly

distribution. The violin plots represent the precipitation monthly distribution. Please see Table 1

for HUC ID information.

29
We divided the OFRs into different class sizes and compared the highest surface area

recorded during the entire time series from both inundation datasets—the highest OFR surface

area represents the reservoir size—with the surface area from the ground-truth dataset (Fig. 4).

Both datasets underestimate the OFRs size for all classes, and the smallest OFRs (0–5 ha and 5–

10 ha) had the highest MPE. The JRC (20.4% and 12.5%) had higher MPE than the DSWE (12.4%

and 7.4%) for both classes (Fig. 4, inset boxplots). In addition, these two classes had the highest

variability in MPE, which decreased as the size of the OFRs increased. In general, the MPE for

both datasets is below 10% for OFRs larger than 10 ha, and the JRC had higher values when

compared to the DSWE for all classes; however, the datasets were statistically different only for

the first two classes (p-value < 0.05) (i.e., OFRs smaller than 10 ha) (Fig. 4, inset boxplots). We

found the highest agreement in size for OFRs with surface area higher than 25 ha (r2 0.99 and 0.98

for DSWE and JRC). Although the smallest classes (0–5 ha and 5–10 ha) presented the highest

MPE, these classes also had good agreement (r2 > 0.80) for both datasets. Overall, the JRC had

smaller r2 values than the DSWE for all classes.

30
Figure 4–Comparison of each OFR ground-truth area with the highest surface area recorded from

both inundation datasets (DSWE and JRC) during the period of analysis (1995–2018). Inset

boxplots depict the distribution of MPE, and the p-value represents the significance of the one-

way ANOVA comparing MPE from DSWE and JRC for each class size.

The magnitude of MPE (Fig. 4, inset boxplots) corroborate with previous studies (Ogilvie

et al., 2018; Jones et al., 2017) that used Landsat imagery to describe small water bodies. Ogilvie

et al., (2018) found that the commission error (i.e., producer error) of the JRC increased up to 50%

for OFRs with area < 5 ha. Jones et al., (2017) used the JRC to assess 272 OFRs—surface area

ranging from 0.09 to 72 ha—and reported that the dataset was unsuitable for monitoring OFRs

31
with surface area smaller than 5.1 ha—56% of OFRs in the Volta Basin—since the magnitude of

mean absolute percent error reached up to 71%.

There are multiple challenges when using Landsat imagery to classify small water bodies;

these challenges help explain the discrepancies found in the OFR size analysis (Fig. 4). Mueller et

al., (2016) developed a methodology to map water bodies at a continental scale (i.e., Australia)

using 25 years of Landsat imagery. They concluded that their product was underestimating the

extent of surface water where there was a mix between water and vegetation pixels—highlighting

a limitation of using Landsat inundation products to assess small wetland and small water bodies.

Similarly, Ogilvie et al., (2018) found that the accuracy of the JRC reduces with the presence of

vegetation and algae, and for shallow water bodies. The authors also highlighted that the geometry

of the OFRs influenced how these water bodies are represented in optical remote sensing. In this

study, OFRs with regular shapes tended to present lower MPE and higher agreement with ground

truthing datasets and OFRs with asymmetric shapes were more inclined to have higher

discrepancies.

Jones et al., (2017) reported analogous findings and demonstrated that the accuracy of the

JRC area estimates improved with increasing the OFR size and decreasing the geometric shape

complexity. We assessed the OFRs geometric shape accounting for their area and perimeter. Figure

5 shows the distribution of SI for all OFRs in each class size. On-farm reservoirs with SI > 1 have

increasingly complex geometric shape. The smallest OFRs (surface area < 15 ha) presented higher

SI values (average SI > 2) when compared to the larger the OFRs (surface area >= 15 ha). On-farm

reservoirs with surface area > 25 ha presented the smallest SI (median SI = 0.77), with upper and

lower limits of 5.12 and 0.17. Although the smallest OFRs had a more complex geometric shape

32
and higher MPE (Fig. 4, 5), we did not find a close relationship (r2 < 0.2 and r2 < 0.10) between

the reservoirs SI and the MPE for any of the classes.

Figure 5–Distribution of the OFRs geometry index (SI) per reservoir class size. The boxplots

depict the SI distribution for all different class sizes.

Figure 6 shows that all classes have high NAindex between January through April, and

November through December (NAindex > 0.30). The JRC has no information for any of the OFRs

during December and January for the entire time series. The size of the bars (i.e., SAratio) helps us

understand the behavior of each OFR class during the year, enabling us to infer seasonal trends in

surface water change. Both datasets show that the OFRs were closer to maximum area during

March, April and May (SAratio ~ 0.75–0.90), and at the lowest capacity (i.e., smaller bars) in June,

July and August (SAratio < 0.60–0.70).

33
Figure 6–Monthly NAindex and SAratio aggregated per OFR class size including the entire period

of analysis (1995–2018). Bars represented in brighter colors (yellows to greens) indicate higher

NAindex, and wider bars depict higher SAratio.

There are multiple factors impacting the OFRs surface area change. These factors include

climate variables (e.g., precipitation, air temperature) as well as anthropogenic variables (e.g.,

reservoir usage), and other environmental factors (e.g., soil type on which the OFRs are built,

distance and topographical distance to major water streams) (Arvor et al., 2018; Habets et al.,

2018; Zhang et al., 2016). In this study, we focused on exploring how much of the variance in

OFRs surface area change can be explained by combining precipitation, air temperature, and the

NDVI from vegetated surface surrounding the OFRs. The OFRs filling process occurs mostly

during the wet season of the year—when the study region receives most of the annual precipitation.

We analyzed precipitation time series for all OFRs, and we found that most of the precipitation

events occur in March, April, and May (Fig. 3, Fig. A1), which coincides with the months when

the OFRs are closer to maximum extent (i.e., highest SAratio values) (Fig. 6). Moreover, the driest

months occur during the summer (i.e., June, July, and August) (Fig. 3, Fig. S. 1), which is the

34
period when the highest temperature values are recorded, and the OFRs are found to be at the

lowest capacities range. This can be an indication of irrigation activities, as the vegetated surfaces

surrounding the OFRs starts to “green up” (i.e., increase in NDVI values) when the study region

receives most of its precipitation (before and early summer), and the highest NDVI values occur

during late and end of summer. The lowest SAratio values during the summer months can be

explained by two main factors: 1) the dry critical months are when the farmers are irrigating their

crops—the last U.S. Department of Agriculture census showed that 100% of rice was irrigated,

and 84% of soybean and 92% of cotton were irrigated at least once during the 2018 growing season

in eastern Arkansas (NASS–USDA, 2020b); and 2) the high records of air temperature can

contribute to increase the OFRs outflow through evaporation, which can be >= 25% of the OFRs

outflow during dry critical months (Friedrich et al., 2018; Guerra et al., 1990; Zhao and Gao,

2019). Previous studies showed that shallow OFRs are more prone to evaporation losses than large

water bodies, and evaporation was identified as a major factor contributing to reduce water storage,

especially in dry critical periods of the year (Fowe et al., 2015; Liebe et al., 2005). We did not

account for evaporation losses in our modeling scenarios because it is beyond the scope of this

study. As illustrated by Zhao and Gao (2019), estimating evaporation losses from reservoirs is not

a trivial task. The authors stressed that detailed information about the OFRs surface water area and

evaporation rate is required to accurately estimate evaporation losses. In addition, besides climate

variables (i.e., air temperature and precipitation), the OFRs hydrological dynamic can be

influenced by the watershed discharge—reduced discharge results in less water supply during the

OFR filling process (Ibrahim et al., 2015; Tinoco et al., 2016). Nonetheless, the OFRs used in this

study are not located on the main water streams, and therefore, the OFRs surface area changes are

35
independent of the discharge seasonality, which is why we did not include discharge in our

analysis.

The OFRs filling process during the wet season is directly related to the volume of water

from precipitation and runoff events, and the filling process depends on the OFRs surface area.

Our results indicate that the DSWE and JRC could be used to describe a set of OFRs behavior in

terms of area change along the months and years (Fig. 6). This relevant to poorly monitored basins

around the world that have a high concentration of OFRs, given that the JRC data is available

globally. The methods implemented in this study could be used in other regions of the world if the

JRC is available. For instance, the Jaguaribe basin (area of 70,000 km2 in the state of Ceará, Brazil)

is known for having a conglomerate of OFRs (Dubreuil and Girard, 1973; Formiga Johnsson and

Kemper, 2007; Gondim et al., 2012). Although efforts to monitor these water bodies date back to

the 1950’s (Dubreuil and Girard, 1973), most of the reservoirs are poorly or not monitored at all.

The use of inundation datasets could then enhance the knowledge of water authorities to create

and/or improve current water planning in the region. Nevertheless, we should also highlight that

the uncertainties carried by the DSWE and JRC datasets (Fig. 4) will contribute to under- or over-

estimation the OFRs volume capacity, if the datasets are used to develop a sustainable water plan

(Fowe et al., 2015).

We further explored how climate variables and NDVI contribute to the variation of the

OFRs SAratio using a generalized additive mixed model approach. The results from the global

models (i.e., including HUC as a random effect) for both the DSWE and JRC datasets are shown

on Table 2. Both models explained less than 15% of the variability in SAratio (r2 < 0.15), and there

were small differences for the fixed effects between the inundation datasets.

36
Table 2–Global generalized additive mixed model results using precipitation, maximum

temperature, and NDVI as fixed effects, and HUC as a random factor for the DSWE and JRC.

DSWE JRC
Intercept 0.615 0.644
Precipitation 0.017 0.013
Maximum Temperature -0.072 -0.069
NDVI 0.069 0.048
Adj. r2 0.098 0.122

The r2 values from the generalized additive models at HUC level are shown on Fig.7. In

general, the JRC presented higher r2 values (0.04–0.29) when compared to DSWE (0.03–0.27).

The r2 values varied between different HUCs, and the HUC with the lowest marginal r2 (<0.1)

(HUC H) has only two OFRs. For HUCs that have a considerable number of reservoirs (>50) and

still presented a low r2 (< 0.20) (e.g., HUCs C and E), the variability in size of OFRs within each

HUC may influence the performance of the model when explaining the variance in SAratio, with

different processes governing reservoirs of different size classes. For instance, HUC E has a total

of 79 reservoirs, and the reservoirs size range varies between 2.4 and 150 ha. The higher JRC r2

values when compared to the DSWE was unexpected, as the JRC has generally higher NAindex, and

no information for January and December for any of the OFRs. A possible explanation is that the

lower amount of information is contributing to less variability in the SAratio, and therefore, higher

r2 values.

The performance of our models (Table 2, Fig. 7) can be attributed to several factors ranging

from the uncertainties contained in the explanatory variables (e.g., climate variables from

interpolated grids) to the fact that we have not included anthropological factors (e.g., reservoir

usage) in our modeling scenarios—NDVI was the only variable relating to crop phenology and

thus potentially related to irrigation water use.

37
Although the PRISM dataset is widely used in studies that assess the long-term impact

(>20 years) of climate variables (e.g., precipitation, air temperature) this dataset is created through

the interpolation of ground data, and therefore, the time series extracted from the dataset carries

the uncertainty from interpolation and modeling techniques (Gao et al., 2017; Spangler et al.,

2019). We believe that our modeling scenarios could be improved by using time series of climate

variables from in-situ data collection. However, it is very unlikely that every farmer will have an

in-situ meteorological station to keep records of climate variables. Our models are also influenced

by the uncertainty in the NDVI time series, which are subject to the same challenges of using

Landsat-based inundation datasets (e.g., data gaps, cloud obscuration, mixed pixels due to pixel

size). In addition, we assumed that the water stored by the OFRs was being used in a radius of 1

km—this assumption was based on the idea that OFRs supply water to cropland that is closely

located. In spite of being a fair assumption, we do not know exactly where the water is being used.

For instance, farmers can have different irrigation schedules for fields adjacent to the OFRs, and

it is common to mix the water resources obtained from the OFR and from irrigation wells.

Furthermore, there are several reservoirs that are located close to green areas (e.g., forest patches),

and that may contribute to less variation in NDVI along the year.

There are other anthropogenic factors that may contribute to the SAratio dynamics, and we

are not accounting for them in our models. These variables could be: (i) intensity of irrigation

activities—we are able to identify the general trends of SAratio, however, depending on the period

of the year, farmers are using the OFRs more or less extensively and this has direct impact on the

OFRs SAratio; (ii) the OFR filling process—the OFRs are mainly filled with precipitation and

runoff events, therefore, the OFRs receive most of their water volume during the wet season.

Nonetheless, farmers may opt to fill the OFRs by pumping water from the inlet streams if they

38
have the infrastructure in place. This filling process impacts SAratio and it is very challenging to

predict, as it varies among different farmers and periods of the year.

The monthly time scale of the inundation datasets can also help explain the performance

of our modeling results. The OFRs’ SAratio may have events, surface area changes, occurring at a

sub-monthly scale that are not being captured by the inundation datasets, and the explanatory

variables. We believe that better modeling results could be obtained by employing satellite datasets

with higher temporal resolution, for example, the Sentinel 1 and 2 (10 day revisit period), and

PlanetScope CubeSats (~ daily revisit period). Lastly, potential modelling improvements could be

achieved employing satellite-based evapotranspiration products (Allen et al., 2015; Mhawej and

Faour, 2020), nonetheless, these products require exhaustive calibration and validation, which is

site specific, and extensive ground-knowledge on which and where the crops are planted—

unfortunately, this information it is not available, and therefore, these products were not included

on this study.

39
DSWE JRC

Figure 7–Generalized additive mixed model adjusted r2 values at HUC level (i.e., spatial scale),

accounting for precipitation, maximum air temperature, and NDVI as fixed effects. The left-hand

side map depicts the results for the DSWE, and the right-hand side map for the JRC. Please see

Table 1 for HUC ID information.

The spatial variability of the fixed effects at the HUC level are shown on Fig. 8.

Precipitation resulted in positive slopes for all HUCs and both inundation datasets, and the slopes

varied among HUCs—DSWE (0.08–0.5 x10-2) and JRC (0.01–0.3 x10-2). Maximum temperature

also varied across different HUCs; however, all slopes were negative DSWE (-0.02– -1.4 x10-2)

and JRC (-0.02– -0.9 x10-2). Similarly, the NDVI varied across different HUCs with positive and

negative slopes—DSWE (0.22–0.47 x10-2) and JRC (-0.02–1.4 x10-2). The results of the

generalized additive mixed models approach (Fig. 7 and 8) can be combined with crop water

requirements to plan the irrigation schedule. For example, crop water needs vary in space and time,

and are influenced by climate factors like precipitation and air temperature (Perin et al., 2019).

40
Therefore, the number and intensity of irrigation events also varies in space and time. The slope

of precipitation, maximum temperature, and NDVI, can be combined and used as a simple index

when farmers are planning their irrigation schedule. This index can be used to increase or decrease

the amount of water applied in each irrigation event depending on the region’s slope values. This

kind of index is commonly used in crop modeling and climate zoning (Battisti and Sentelhas, 2019;

Boote et al., 1998), in which the simulated crop yield is penalized or not based on the region’s

climate (e.g., crop yield is penalized under drought conditions). The slope of precipitation,

maximum temperature, and NDVI could then enhance tools which support irrigation management

decisions, helping farmers plan for regions and periods of the year that will require more or less

water.

41
Figure 8–Generalized additive mixed model fixed effects of precipitation, maximum temperature

and NDVI at HUC level (i.e., spatial scale). The left-hand side map depicts the results for the

DSWE, and the right-hand side map for the JRC.

42
The DSWE and JRC inundation datasets enable us to describe the general trends of the

OFRs surface water area change and to depict the effects of climate variables on surface area

change; nonetheless, we should bear in mind that hydrological monitoring of small water bodies

using Landsat inundation products has several constraints that need to be assessed when applying

the methods to other study regions. The restrictions of using the long-term Landsat archive to

monitor small water bodies range from data gaps (i.e., cloud obscuration), which have direct

impact on the accuracy of the seasonal information, to spatial resolution issues that prevent the

accurate classification of water bodies, to the 16-day repeat cycle that may miss rapid flood events

that occur on a shorter timescale. Our analysis points out several challenges of using Landsat

inundation products for water planning purposes. For example, these datasets may be unsuitable

for monitoring reservoirs with area between 0–5 ha, as both datasets can have errors higher than

20% for reservoir extent (Figure 4). Furthermore, for certain months of the year (e.g., December

and January) there is a higher occurrence of missing information, which is an important limitation

when aiming to use the datasets for continuous monitoring (Figure 6). The JRC has no information

for the months of December and January, and this could be partially explained by the algorithms

used to create this inundation dataset. The JRC was primarily designed to represent the occurrence

of open surface water, and not to detect surface water when there is spectral reflectance mixing of

the water and vegetation (DeVries et al., 2017; Jones, 2019)—which is a common characteristic

of the OFRs that are relatively shallow and small water bodies—therefore, possibly leading to

inadequately classifying OFRs. In addition, the absence of information is related to the well-known

Landsat 7 ETM+ SLC scan line correctors failure after 31 May 2003 (Hossain et al., 2015). These

limitations are overcome by combining multiple satellite sources to monitor surface water

dynamics. Vanthof and Kelly (2019) used the JRC in conjunction with higher spatial and temporal

43
satellite imagery and were able to increase the number of observations of OFRs, and reduce the

mapping error to 3% of the reservoir size, and therefore, enable the use of different satellite

imagery for water planning.

Regardless of the limitations of the Landsat surface water datasets addressed above, the

DSWE and JRC can be used as a first step to monitor water bodies at a large-scale, which is

important in regions where there is a conglomerate of these water bodies. The National Inventory

of Dams (NIDs) dataset (http://nid.usace.army.mil/) provides the location (i.e., longitude and

latitude) of more than 80,000 dams located in the U.S. Although this dataset does not have

information about OFRs, this dataset brings relevant information of the major dams in the U.S.,

including the construction year, size in area, and relevant hydrological information (i.e., depth,

storage capacity). Also, the NID dataset can be coupled with surface water extent information to

enhance knowledge on the water budget in regions with a concentration of dams. Briefly, if the

depth of the reservoir and its topographic information (e.g., digital elevation model) are available,

it is possible to calculate the storage-area-depth for these reservoirs (Yigzaw et al., 2018), which

can be combined with surface area to estimate volume storage change from two different time-

steps (Yao et al., 2018), and create a time series of water volume change (Vanthof and Kelly,

2019). In addition, the same approach detailed above, and the methods used in our study, could be

expanded to other countries by employing global water impoundments datasets, for instance, the

global georeferenced database of dams (Mulligan et al., 2020), and the geo-referenced database on

dams (Lehner et al., 2011).

Our study furthers the use of Landsat inundation datasets where there is insufficient OFR

monitoring. The biggest hurdle of replicating our study to other countries or regions is the lack of

an OFR ground-truth dataset. However, nowadays, there is daily high spatial resolution (~3m)

44
satellite imagery for the entire globe (Planet Team, 2017), which enable future investigators to

create a similar OFR dataset used in our study. A ground-truth dataset can be created by combining

multiple platforms and satellite imagery sources. Our results showed that the DSWE or JRC can

be used to understand the temporal variability of OFRs, and although there are mapping errors due

to the spatial resolution of the Landsat imagery (Fig. 4), a proxy of the OFR shape can be obtained

by calculating the OFR frequency of inundation using the entire Landsat catalog—this process is

widely used in the scientific community (Pekel et al., 2016; Soulard et al., 2020; Walker et al.,

2020). The OFR shape proxy can be converted to a vector and be used to retrieve high-resolution

satellite imagery from PlanetScope CubeSats Imagery (~3m spatial resolution)—Planet Labs has

an international Education and Research Program that allows scientists around the globe to use

PlanetScope CubeSats imagery free of costs.

This study is the first effort to apply long-term surface water extent datasets to large-scale

OFRs monitoring in the U.S. Therefore, we highlight that future investigations should weight on

several factors before deciding which dataset to use (i.e., DSWE or JRC). The first of these factors

(i) is imagery processing: we downloaded the level-3 analysis ready data DSWE scenes. Though

the dataset is already pre-processed, several steps are required to obtain the monthly water history,

which can be a demanding task, especially when downloading and processing the dataset for a

large area. The JRC monthly water history was obtained through GEE using a straight-forward

downloading approach. The GEE platform can be used to recreate the DSWE monthly water layer

(Ahmad et al., 2020; Walker et al., 2020), and therefore, attenuate some of the demanding tasks

(e.g., downloading, recreating tiles, etc.) related to the DSWE. The second factor (ii) is missing

information: although both datasets are based on the Landsat archive, the DSWE has less missing

information than the JRC which is essential to continuous large-scale monitoring. The JRC is a

45
global dataset, and does not specifically target the classification of water bodies that have a

frequent spectral reflectance mixing (i.e., OFRs). Therefore, the algorithms used to classify surface

water tend to be more stringent when classifying water pixels, resulting in higher occurrences of

absent information (Pekel et al., 2016). The third factor (iii) is maximum area percent error: both

datasets levy the same challenges related to spatial resolution when classifying small water bodies

using Landsat products (Jones et al., 2017; Ogilvie et al., 2018; Solander et al., 2016). However,

the DSWE captured the maximum extent of OFRs with less errors and higher agreement when

compared to the JRC. Lastly, we should have in mind that our study is limited to maximum

reservoir extent analysis and a complete assessment of these datasets requires field validation

and/or measurement of the surface water area using higher spatiotemporal resolution satellite

imagery (Vanthof and Kelly, 2019).

4. CONCLUSION

Landsat surface inundation datasets can be used to map OFRs maximum extent and surface

water area seasonality. The DSWE had smaller mean percent errors (p-value <0.05) and higher

agreement compared to the JRC when mapping reservoirs smaller than 10 ha. Both datasets enable

us to characterize the seasonality of the OFRs area change. In eastern Arkansas, the OFRs are

mostly filled in March, April, and May, which can be partially explained by the precipitation

regimen of the region. We applied a generalized additive mixed model approach to quantify the

impact of precipitation, maximum temperature, and the NDVI—from the land surrounding the

OFRs—on the OFRs surface area change, and we found that the impact of these variables varies

in space (i.e., for different HUCs). Our results showed that these datasets can be used to enhance

hydrological assessments in poorly monitored basins that have a concentration of artificial small

water bodies, and the methods employed on this study could be used to enhance knowledge about

46
the OFRs surface water dynamics in other regions and different countries, if the inundation datasets

(i.e., DSWE or JRC) are available. Our findings are a unique contribution on which inundation

dataset to employ when aiming to monitor small water bodies at a large-scale in the US. Lastly,

the two main drawbacks imposed by these datasets when implementing them for water planning

purposes and policymaking are: (i) inconsistent surface water area time series, due to clouds and

sensor related issues, and the 16-day repeat cycle of the Landsat imagery; and (ii) spatial resolution

of 30-m that may be unsuitable to monitor OFRs with surface area between 0–5 ha.

47
REFERENCES

Adams, W.M., Hughes, F.M.R., 1986. The environmental effects of dam construction in tropical
Africa: Impacts and planning procedures. Geoforum 17, 403–410.
https://doi.org/10.1016/0016-7185(86)90007-2

Ahmad, S.K., Hossain, F., Eldardiry, H., Pavelsky, T.M., 2020. A Fusion Approach for Water
Area Classification Using Visible, Near Infrared and Synthetic Aperture Radar for South
Asian Conditions. IEEE Trans. Geosci. Remote Sens. 58, 2471–2480.
https://doi.org/10.1109/TGRS.2019.2950705

Allen, R.G., Morton, C., Kamble, B., Kilic, A., Huntington, J., Thau, D., Gorelick, N., Erickson,
T., Moore, R., Trezza, R., 2015. EEFlux: A Landsat-based evapotranspiration mapping tool
on the Google Earth Engine, in: 2015 ASABE/IA Irrigation Symposium: Emerging
Technologies for Sustainable Irrigation-A Tribute to the Career of Terry Howell, Sr.
Conference Proceedings. American Society of Agricultural and Biological Engineers, pp.
1–11.

Althoff, D., Rodrigues, L.N., da Silva, D.D., 2020. Impacts of climate change on the evaporation
and availability of water in small reservoirs in the Brazilian savannah. Clim. Change 159,
215–232. https://doi.org/10.1007/s10584-020-02656-y

Arkansas Drought Planning: Summary Report, 2016. Little Rock, Arkansas.

Arvor, D., Daher, F.R.G., Briand, D., Dufour, S., Rollet, A.J., Simões, M., Ferraz, R.P.D., 2018.
Monitoring thirty years of small water reservoirs proliferation in the southern Brazilian
Amazon with Landsat time series. ISPRS J. Photogramm. Remote Sens. 145, 225–237.
https://doi.org/10.1016/j.isprsjprs.2018.03.015

Avisse, N., Tilmant, A., Müller, M.F., Zhang, H., 2017. Monitoring small reservoirs’ storage
with satellite remote sensing in inaccessible areas. Hydrol. Earth Syst. Sci. 21, 6445–6459.
https://doi.org/10.5194/hess-21-6445-2017

Barrow, C.J., 1987. The environmental impacts of the tucuri dam on the middle and lower
tocantins river basin, Brazil. Regul. Rivers Res. Manag. 1, 49–60.
https://doi.org/10.1002/rrr.3450010106

Battisti, R., Sentelhas, P.C., 2019. Characterizing Brazilian soybean-growing regions by water
deficit patterns. F. Crop. Res. 240, 95–105.
https://doi.org/https://doi.org/10.1016/j.fcr.2019.06.007

Berg, M.D., Popescu, S.C., Wilcox, B.P., Angerer, J.P., Rhodes, E.C., Mcalister, J., Fox, W.E.,
2016. Small farm ponds: Overlooked features with important impacts on watershed
sediment transport. J. Am. Water Resour. Assoc. 52, 67–76. https://doi.org/10.1111/1752-
1688.12369

48
Boote, K.J., Jones, J.W., Hoogenboom, G., Pickering, N.B., 1998. Simulation of crop growth:
CROPGRO model, in: Agricultural Systems Modeling and Simulation. Marcel Dekker New
York, NY, pp. 651–692.

Clark, B.R., Hart, R.M., Gurdak, J.J., 2011. Groundwater availability of the Mississippi
embayment: U.S. Geological Survey Professional Paper, 5th ed.

DeVries, B., Huang, C., Lang, M., Jones, J., Huang, W., Creed, I., Carroll, M., 2017. Automated
Quantification of Surface Water Inundation in Wetlands Using Optical Satellite Imagery.
Remote Sens. 9, 807. https://doi.org/10.3390/rs9080807

Downing, J.A., 2010. Emerging global role of small lakes and ponds: little things mean a lot.
Limnetica 29, 9–24. https://doi.org/remote

Dubreuil, P., Girard, G., 1973. Influence of a very large number of small reservoirs on the annual
flow regime of a tropical stream. Washingt. DC Am. Geophys. Union Geophys. Monogr.
Ser. 17, 295–299.

Dudgeon, D., 2000. Large-Scale Hydrological Changes in Tropical Asia: Prospects for Riverine
Biodiversity: The construction of large dams will have an impact on the biodiversity of
tropical Asian rivers and their associated wetlands. Bioscience 50, 793–806.
https://doi.org/10.1641/0006-3568(2000)050[0793:LSHCIT]2.0.CO;2

Fearnside, P.M., 2006. Dams in the Amazon: Belo Monte and Brazil’s Hydroelectric
Development of the Xingu River Basin. Environ. Manage. 38, 16.
https://doi.org/10.1007/s00267-005-0113-6

Fearnside, P.M., 2001. Environmental impacts of Brazil’s Tucuruí Dam: Unlearned lessons for
hydroelectric development in amazonia. Environ. Manage.
https://doi.org/10.1007/s002670010156

Formiga Johnsson, R.M., Kemper, K E, 2007. Brazil: Jaguaribe Basin, in: Kemper, Karin E,
Dinar, A., Blomquist, W. (Eds.), Integrated River Basin Management through
Decentralization. Springer Berlin Heidelberg, Berlin, Heidelberg, pp. 111–129.
https://doi.org/10.1007/978-3-540-28355-3_6

Fowe, T., Karambiri, H., Paturel, J.E., Poussin, J.C., Cecchi, P., 2015. Water balance of small
reservoirs in the volta basin: A case study of boura reservoir in burkina faso. Agric. Water
Manag. 152, 99–109. https://doi.org/10.1016/j.agwat.2015.01.006

Friedrich, K., Grossman, R.L., Huntington, J., Blanken, P.D., Lenters, J., Holman, K.D., Gochis,
D., Livneh, B., Prairie, J., Skeie, E., Healey, N.C., Dahm, K., Pearson, C., Finnessey, T.,
Hook, S.J., Kowalski, T., 2018. Reservoir Evaporation in the Western United States:
Current Science, Challenges, and Future Needs. Bull. Am. Meteorol. Soc. 99, 167–187.
https://doi.org/10.1175/BAMS-D-15-00224.1

49
Gao, J., Sheshukov, A.Y., Yen, H., White, M.J., 2017. Impacts of alternative climate information
on hydrologic processes with SWAT: A comparison of NCDC, PRISM and NEXRAD
datasets. CATENA 156, 353–364.
https://doi.org/https://doi.org/10.1016/j.catena.2017.04.010

Gondim, R.S., de Castro, M.A.H., Maia, A. de H.N., Evangelista, S.R.M., Fuck, S.C. de F.,
2012. Climate Change Impacts on Irrigation Water Needs in the jaguaribe River Basin1.
JAWRA J. Am. Water Resour. Assoc. 48, 355–365. https://doi.org/10.1111/j.1752-
1688.2011.00620.x

Guerra, L.C., Watson, P.G., Bhuiyan, S.I., 1990. Hydrological analysis of farm reservoirs in
rainfed rice areas. Agric. Water Manag. 17, 351–366.
https://doi.org/https://doi.org/10.1016/0378-3774(90)90085-D

Habets, F., Molénat, J., Carluer, N., Douez, O., Leenhardt, D., 2018. The cumulative impacts of
small reservoirs on hydrology: A review. Sci. Total Environ.
https://doi.org/10.1016/j.scitotenv.2018.06.188

Habets, F., Philippe, E., Martin, E., David, C.H., Leseur, F., 2014. Small farm dams: Impact on
river flows and sustainability in a context of climate change. Hydrol. Earth Syst. Sci. 18,
4207–4222. https://doi.org/10.5194/hess-18-4207-2014

Heimhuber, V., Tulbure, M.G., Broich, M., Xie, Z., Hurriyet, M., 2019. The role of GRACE
total water storage anomalies, streamflow and rainfall in stream salinity trends across
Australia’s Murray-Darling Basin during and post the Millennium Drought. Int. J. Appl.
Earth Obs. Geoinf. 83, 101927. https://doi.org/https://doi.org/10.1016/j.jag.2019.101927

Hossain, M.S., Bujang, J.S., Zakaria, M.H., Hashim, M., 2015. Assessment of the impact of
Landsat 7 Scan Line Corrector data gaps on Sungai Pulai Estuary seagrass mapping. Appl.
Geomatics 7, 189–202. https://doi.org/10.1007/s12518-015-0162-3

Hughes, D.A., Mantel, S.K., 2010. Estimation des incertitudes lors de la simulation des impacts
de petites retenues agricoles sur les régimes d’écoulement en Afrique du Sud. Hydrol. Sci.
J. 55, 578–592. https://doi.org/10.1080/02626667.2010.484903

Ibrahim, B., Wisser, D., Barry, B., Fowe, T., Aduna, A., 2015. Hydrological predictions for
small ungauged watersheds in the Sudanian zone of the Volta basin in West Africa. J.
Hydrol. Reg. Stud. 4, 386–397. https://doi.org/https://doi.org/10.1016/j.ejrh.2015.07.007

Jones, J.W., 2019. Improved automated detection of subpixel-scale inundation-revised Dynamic


Surface Water Extent (DSWE) partial surface water tests. Remote Sens. 11.
https://doi.org/10.3390/rs11040374

Jones, J.W., 2015. Efficient wetland surface water detection and monitoring via landsat:
Comparison with in situ data from the everglades depth estimation network. Remote Sens.
7, 12503–12538. https://doi.org/10.3390/rs70912503

50
Jones, S.K., Fremier, A.K., DeClerck, F.A., Smedley, D., Pieck, A.O., Mulligan, M., 2017. Big
data and multiple methods for mapping small reservoirs: Comparing accuracies for
applications in agricultural landscapes. Remote Sens. 9. https://doi.org/10.3390/rs9121307

Karran, D.J., Westbrook, C.J., Wheaton, J.M., Johnston, C.A., Bedard-Haughn, A., 2017. Rapid
surface-water volume estimations in beaver ponds. Hydrol. Earth Syst. Sci. 21, 1039–1050.
https://doi.org/10.5194/hess-21-1039-2017

L Tacker, P., D Vories, E., Carman, D., 2010. Water Supply Problems and Solutions in the
Grand Prairie Region of Arkansas, in: 5th National Decennial Irrigation Conference
Proceedings, 5-8 December 2010, Phoenix Convention Center, Phoenix, Arizona USA.
ASABE, St. Joseph, MI. https://doi.org/https://doi.org/10.13031/2013.35853

Liebe, J., van de Giesen, N., Andreini, M., 2005. Estimation of small reservoir storage capacities
in a semi-arid environment. Phys. Chem. Earth 30, 448–454.
https://doi.org/10.1016/j.pce.2005.06.011

Lovelace, J.K., Nielsen, M.G., Read, A.L., Murphy, C.J., Maupin, M.A., 2015. Estimated
groundwater withdrawals from principal aquifers in the United States, 2015. US Geological
Survey. https://doi.org/doi.org/10.3133/cir1464

Maingi, J.K., Marsh, S.E., 2002. Quantifying hydrologic impacts following dam construction
along the Tana River, Kenya. J. Arid Environ. 50, 53–79.
https://doi.org/10.1006/JARE.2000.0860

Marston, L., Konar, M., Cai, X., Troy, T.J., 2015. Virtual groundwater transfers from
overexploited aquifers in the United States. Proc. Natl. Acad. Sci. U. S. A. 112, 8561–8566.
https://doi.org/10.1073/pnas.1500457112

Martins, D.P., Meurer, M., 2009. Changes in a large regulated tropical river: The Paraná River
downstream from the Porto Primavera Dam, Brazil. Geomorphology 113, 230–238.
https://doi.org/10.1016/J.GEOMORPH.2009.03.015

Meigh, J., 1995. The impact of small farm reservoirs on urban water supplies in Botswana. Nat.
Resour. Forum 19, 71–83. https://doi.org/10.1111/j.1477-8947.1995.tb00594.x

Mhawej, M., Faour, G., 2020. Open-source Google Earth Engine 30-m evapotranspiration rates
retrieval: The SEBALIGEE system. Environ. Model. Softw. 133, 104845.
https://doi.org/https://doi.org/10.1016/j.envsoft.2020.104845

Mime, M.M., Young, D.W., 1989. The Impact Of Stockwatering Ponds (Stockponds) On Runoff
From Large Arizona Watersheds. JAWRA J. Am. Water Resour. Assoc. 25, 165–173.
https://doi.org/10.1111/j.1752-1688.1989.tb05678.x

Mueller, N., Lewis, A., Roberts, D., Ring, S., Melrose, R., Sixsmith, J., Lymburner, L.,
McIntyre, A., Tan, P., Curnow, S., Ip, A., 2016. Water observations from space: Mapping

51
surface water from 25years of Landsat imagery across Australia. Remote Sens. Environ.
174, 341–352. https://doi.org/10.1016/j.rse.2015.11.003

National Agricultural Statistics Service–U.S. Department of Agriculture (NASS–USDA), 2020a.


CropScape–Cropland data layer. Washington, DC. https://nassgeodata.gmu.edu/CropScape

National Agricultural Statistics Service–U.S. Department of Agriculture (NASS–USDA), 2020b.


Quick stats. Washington, DC. https://www.nass.usda.gov/Quick_Stats/

National Agricultural Statistics Service–U.S. Department of Agriculture (NASS–USDA), 2017.


Census of agriculture. Washington, DC. https://www.nass.usda.gov/AgCensus/

Ogilvie, A., Belaud, G., Massuel, S., Mulligan, M., Le Goulven, P., Calvez, R., 2018. Surface
water monitoring in small water bodies: Potential and limits of multi-sensor Landsat time
series. Hydrol. Earth Syst. Sci. 22, 4349–4380. https://doi.org/10.5194/hess-22-4349-2018

Pekel, J.F., Cottam, A., Gorelick, N., Belward, A.S., 2016. High-resolution mapping of global
surface water and its long-term changes. Nature 540, 418–422.
https://doi.org/10.1038/nature20584

Perin, V., Sentelhas, P.C., Dias, H.B., Santos, E.A., 2019. Sugarcane irrigation potential in
Northwestern São Paulo, Brazil, by integrating Agrometeorological and GIS tools. Agric.
Water Manag. 220, 50–58. https://doi.org/https://doi.org/10.1016/j.agwat.2019.04.012

PRISM Climate Group, 2011. PRISM Climate data. Oregon State Univ.

R Core Team, 2020. R: A language and environment for statistical computing.

Schrader, T.P., 2015. Water levels and water quality in the Mississippi River Valley alluvial
aquifer in eastern Arkansas, 2012, Scientific Investigations Report. Reston, VA.
https://doi.org/10.3133/sir20155059

Seo, B., Lee, J., Lee, K.-D., Hong, S., Kang, S., 2019. Improving remotely-sensed crop
monitoring by NDVI-based crop phenology estimators for corn and soybeans in Iowa and
Illinois, USA. F. Crop. Res. 238, 113–128.
https://doi.org/https://doi.org/10.1016/j.fcr.2019.03.015

Shults, D.D., Nowlin, W.J., Yaeger, M.A., Massey, J.H., Reba, M.L., 2020. A spatiotemporal
anlysis quantifying the need for more on-farm reservoirs to reduce groundwater use in the
Cache and L’Anguille River Regions in Northeaster AR, in: ESRI User Conference. Online.

Smith, D.G., 1998. Vibracoring: a new method for coring deep lakes. Palaeogeogr.
Palaeoclimatol. Palaeoecol. 140, 433–440. https://doi.org/10.1016/S0031-0182(98)00031-5

Solander, K.C., Reager, J.T., Famiglietti, J.S., 2016. How well will the Surface Water and Ocean
Topography (SWOT) mission observe global reservoirs? Water Resour. Res. 52, 2123–

52
2140. https://doi.org/10.1002/2015WR017952
Soulard, C.E., Walker, J.J., Petrakis, R.E., 2020. Implementation of a Surface Water Extent
Model in Cambodia using Cloud-Based Remote Sensing. Remote Sens. 12, 984.
https://doi.org/10.3390/rs12060984

Spangler, K.R., Weinberger, K.R., Wellenius, G.A., 2019. Suitability of gridded climate datasets
for use in environmental epidemiology. J. Expo. Sci. Environ. Epidemiol. 29, 777–789.
https://doi.org/http://dx.doi.org/10.1038/s41370-018-0105-2

Sullivan, M.E., Delp, W.M., 2012. Water conservation planning: How a systems approach to
irrigation promotes sustainable water use. NABC.

Planet Team, 2017. Planet application program interface: In space for life on Earth. San
Francisco. CA 2017, 40.

Tinoco, V., Willems, P., Wyseure, G., Cisneros, F., 2016. Evaluation of reservoir operation
strategies for irrigation in the Macul Basin, Ecuador. J. Hydrol. Reg. Stud. 5, 213–225.
https://doi.org/https://doi.org/10.1016/j.ejrh.2015.12.063

Tulbure, M.G., Broich, M., 2019. Spatiotemporal patterns and effects of climate and land use on
surface water extent dynamics in a dryland region with three decades of Landsat satellite
data. Sci. Total Environ. 658, 1574–1585.
https://doi.org/https://doi.org/10.1016/j.scitotenv.2018.11.390

Vanthof, V., Kelly, R., 2019. Water storage estimation in ungauged small reservoirs with the
TanDEM-X DEM and multi-source satellite observations. Remote Sens. Environ. 235.
https://doi.org/10.1016/j.rse.2019.111437

Walker, J.J., Soulard, C.E., Petrakis, R.E., 2020. Integrating stream gage data and Landsat
imagery to complete time series of surface water extents in Central Valley, California. Int. J.
Appl. Earth Obs. Geoinf. 84, 101973. https://doi.org/10.1016/j.jag.2019.101973

Wang, W., Lee, X., Xiao, W., Liu, S., Schultz, N., Wang, Y., Zhang, M., Zhao, L., 2018. Global
lake evaporation accelerated by changes in surface energy allocation in a warmer climate.
Nat. Geosci. 11, 410–414. https://doi.org/10.1038/s41561-018-0114-8

Wood, S., Scheipl, F., 2014. gamm4: Generalized additive mixed models using mgcv and lme4.
R package version 0.2-6.

Wood, S.N., 2017. Generalized additive models: an introduction with R. CRC press.

Yaeger, M.A., Massey, J.H., Reba, M.L., Adviento-Borbe, M.A.A., 2018. Trends in the
construction of on-farm irrigation reservoirs in response to aquifer decline in eastern
Arkansas: Implications for conjunctive water resource management. Agric. Water Manag.
208, 373–383. https://doi.org/10.1016/j.agwat.2018.06.040

53
Yaeger, M.A., Reba, M.L., Massey, J.H., Adviento-Borbe, M.A.A., 2017. On-farm irrigation
reservoirs in two Arkansas critical groundwater regions: A comparative inventory. Appl.
Eng. Agric. 33, 869–878. https://doi.org/10.13031/aea.12352

Yang, Z., Mueller, R., Crow, W., 2013. US national cropland soil moisture monitoring using
SMAP, in: 2013 IEEE International Geoscience and Remote Sensing Symposium -
IGARSS. pp. 3746–3749. https://doi.org/10.1109/IGARSS.2013.6723645

Yao, F., Wang, J., Yang, K., Wang, C., Walter, B.A., Crétaux, J.F., 2018. Lake storage variation
on the endorheic Tibetan Plateau and its attribution to climate change since the new
millennium. Environ. Res. Lett. 13. https://doi.org/10.1088/1748-9326/aab5d3

Yigzaw, W., Li, H.-Y., Demissie, Y., Hejazi, M.I., Leung, L.R., Voisin, N., Payn, R., 2018. A
New Global Storage-Area-Depth Data Set for Modeling Reservoirs in Land Surface and
Earth System Models. Water Resour. Res. 54, 10, 310–372, 386.
https://doi.org/10.1029/2017WR022040

Zhang, S., Foerster, S., Medeiros, P., de Araújo, J.C., Motagh, M., Waske, B., 2016. Bathymetric
survey of water reservoirs in north-eastern Brazil based on TanDEM-X satellite data. Sci.
Total Environ. 571, 575–593. https://doi.org/10.1016/j.scitotenv.2016.07.024

Zhao, G., Gao, H., 2019. Estimating reservoir evaporation losses for the United States: Fusing
remote sensing and modeling approaches. Remote Sens. Environ. 226, 109–124.
https://doi.org/https://doi.org/10.1016/j.rse.2019.03.015

54
CHAPTER 3: A MULTI-SENSOR SATELLITE IMAGERY APPROACH TO MONITOR

ON-FARM RESERVOIRS

Published in Remote Sensing of Environment November 2021


(https://doi.org/10.1016/j.rse.2021.112796)

Authors
Vinicius Perin1, Mirela G. Tulbure1, Mollie D.Gaines1, Michele L.Reba2, Mary A.Yaeger3

Affiliations
1
Center for Geospatial Analytics, North Carolina State University, 2800 Faucette Drive, Raleigh,
NC, 27606, USA
2
USDA-Agricultural Research Service, Delta Water Management Research Unit, Jonesboro, AR,
72467, USA
3
Center for Applied Earth Science and Engineering Research, The University of Memphis, 3675
Alumni Drive, Memphis, TN, 38152, USA

Acknowledgements
The first author is supported by NASA through the Future Investigators in NASA Earth and Space

Science and Technology fellowship. Part of the funding for this study came from MGT start up

package at North Carolina State University. This work was also supported by European Space

Agency in the framework of the Network of Resources Call, which sponsored the Sentinel Hub

Non-commercial Research license. The authors acknowledge the insightful revision provided by

Dr. Jida Wang and Dr. Brian Reich. The authors acknowledge the help on developing the Kalman

filter algorithm provided by Dr. Joshua Gray, Isabella Hinks, Allen Acken, and Pratap Adhikari.

Abstract
Fresh water stored by on-farm reservoirs (OFRs) is an important component of surface hydrology

and is critical for meeting global irrigation needs. Farmers use OFRs to store water during the wet

55
season and for crop irrigation during the dry season, yet their seasonal and inter-annual variability

and downstream impacts are not quantified. Therefore, OFRs' sub-weekly surface area changes

are critical to understanding their dynamics and mitigating their downstream impacts. However,

prior to the recent increase in satellite imagery availability and improvement in sensors' spatial

resolution, monitoring the OFRs' sub-weekly surface area changes across space and time was

challenging because OFRs occur in high numbers (i.e. thousands) and are small water bodies (<

50 ha). We propose a novel multi-sensor approach to monitor OFRs surface areas, developed based

on 736 OFRs in eastern Arkansas, USA, which leverages the use of PlanetScope (PS), RapidEye

(RE), Sentinel 2 (S2), and Sentinel 1 (S1). First, we estimate the uncertainties in surface area for

each sensor by comparing the surface area estimates to a validation dataset, and by comparing RE,

S2 and S1 to PS—the sensor with the highest spatial resolution (i.e. 3.125 m). Second, we use the

uncertainties of each sensor with a data assimilation algorithm based on the Kalman filter to obtain

sub-weekly surface area time series for all OFRs. Our results show the lowest uncertainties for PS,

followed by RE, S2 and S1. These uncertainties varied according to the OFRs' size and shape

complexities. The surface area estimates derived from the Kalman filter including only the optical

sensors resulted in high agreement (r2 > 0.95) and small uncertainties (4–8%) when compared to

the validation dataset. We found higher uncertainties (5–14%) when adding S1 to the Kalman

filter—this is related to the higher uncertainties found for S1 (~20%). The algorithm can assimilate

optical and radar satellite data to increase the OFRs' surface area time series cadence allowing us

to investigate sub-weekly surface area changes. The algorithm is not sensor-specific, and it

accounts for the uncertainties in both the sensors observations and the resulting surface areas,

which are key advantages when compared to other algorithms used to combine satellite data. By

improving the surface area observations cadence and providing the surface area uncertainties, the

56
approach presented in this study has the potential to enhance water conservation plans by allowing

better assessment and management of the OFRs.

Keywords: On-farm reservoirs, Water management, Data assimilation, Kalman filter, Multi-

sensor satellite imagery

57
1. INTRODUCTION

1.1 Background on OFRs and remotely sensed data used for OFR mapping
Fresh water stored by on-farm reservoirs (OFRs) is a fundamental component of surface

hydrology and is critical for meeting global irrigation needs (Döll et al., 2009; Downing, 2010;

Van Den Hoek et al., 2019). Farmers use OFRs to store water during the wet season for crop

irrigation during the dry season. There are more than 2.6 million OFRs in the USA alone (Downing

et al., 2006; Renwick et al., 2005), and many of these OFRs were constructed during the last 40

years (Berg et al., 2016; Downing, 2010). Despite their importance for irrigating crops, OFRs can

contribute to downstream water stress by decreasing stream discharge and peak flow in the

watersheds where they are built (Habets et al., 2018, Habets et al., 2014; Mime and Young, 1989),

thereby exacerbating water stress, already intensified by climate change and population growth

(Vörösmarty et al., 2010). The OFRs' surface areas vary frequently during the crop growing season

when farmers are irrigating, and farmers may opt to pump water from nearby streams to fill the

OFRs if more water is required to irrigate their crops. Therefore, understanding the OFRs' sub-

weekly surface area changes is critical to assess their seasonal and inter-annual variability, which

is key information when modeling the OFRs impacts on surface hydrology and mitigating their

downstream impacts.

Prior to the recent development and availability in private and public satellite imagery,

widespread monitoring of sub-weekly changes in OFRs' surface area across space and time was

challenging due to satellite observation latency (i.e. irregular intervals) and spatial resolution (i.e.

30 m resolution). The newest private images available include PlanetScope (PS, 3.125 m), and

RapidEye (RE, 5 m) that recently became freely available globally. Within the public domain, the

latest images include Sentinel 1 (S1, 10 m) and Sentinel 2 (S2, 10 m). A common approach to

58
assess the spatial and temporal variability of OFRs was to use long-term (>25–30 years) Landsat-

based (Arvor et al., 2018; Jones et al., 2017; Ogilvie et al., 2018a; Perin et al., 2021; Yao et al.,

2019) inundation datasets (Jones, 2019, Jones, 2015; Pekel et al., 2016). However, the Landsat

spatial resolution of 30 m is unsuitable to monitor OFRs with surface area 0.1–5 ha (surface area

monitoring error > 20% or higher, Perin et al., 2021). Furthermore, the surface area time series

was restricted to a few annual observations—due to clouds, sensor issues and the 16-day repeat

cycle.

The monitoring of OFRs can be improved by leveraging the latest developments and

availability in satellite imagery to reduce observation latency (Ogilvie et al., 2020, Ogilvie et al.,

2018a; Vanthof and Kelly, 2019). Recent research highlighted the use of PS (Cooley et al., 2017;

Hondula et al., 2021; Mishra et al., 2020; Vanthof and Kelly, 2019), RE (Casadei et al., 2019;

Cooley et al., 2017; Pickens et al., 2020; Pinhati et al., 2020), S1 (Ahmad et al., 2020; Bioresita et

al., 2019; López-Caloca et al., 2020; Vanthof and Kelly, 2019), and S2 (Ahmad et al., 2020;

Ogilvie et al., 2020; Pena-Regueiro et al., 2020; Vanthof and Kelly, 2019; Yang et al., 2020b;

Yang et al., 2017) to monitor water bodies. However, few, if any, studies have assessed a multi-

sensor approach that uses optical and radar sensors (e.g., PS, RE, S2, and S1) of different spatial

resolution, and assessed the surface area uncertainties of each sensor. The OFRs' surface areas,

alongside their uncertainties, are necessary to computing the OFRs' water volume changes and

thus have the potential to help policymakers and water authorities to quantify the OFRs' inflows

and outflows in space and time. In addition, the uncertainty of each sensor needs to be assessed

for OFRs of different sizes and shape complexities, as surface area classification accuracy will

vary according to the water body size and shape complexity (Bonnema and Hossain, 2019; Perin

et al., 2021; Solander et al., 2016; Vanthof and Kelly, 2019).

59
1.2 Algorithms to combine different satellite imagery for OFR mapping
Even though the use of a multi-sensor approach reduces the latency of OFRs' surface area

observations (Ogilvie et al., 2018a; Vanthof and Kelly, 2019), there are several challenges that

need to be overcome to obtain sub-weekly surface area time series. These challenges include filling

gaps when there are no satellite observations (e.g., due to sensor latency, cloud obscuration, sensor

issues), and accounting for discrepancies in surface area between different sensors—these

discrepancies are attributed to different sensors' radiometric and terrain calibrations, and spectral,

radiometric, and spatial resolutions. Therefore, it is necessary to employ an algorithm that

combines the data and the capabilities from multiple sensors, which generally entails data fusion

or data assimilation algorithms. Spatial and temporal image fusion algorithms include unmixing

linear spectral approaches, weighting-based approaches, Bayesian-based methods, learning-based

methods, and hybrid methods that combine aspects of multiple approaches (Zhu et al., 2018).

However, most of these fusion algorithms tend to be limited to specific sensors and do not provide

uncertainty estimates for the output values. Meanwhile, data assimilation algorithms use

information from multiple sensors by taking advantage of a model (e.g. linear or non-linear) to

describe a physical phenomenon evolution over time (e.g. vegetation phenology changes). An

overview of common assimilation algorithms is provided by Zhang and Moore (2014), and a well-

known data assimilation algorithm used in remote sensing is the Kalman filter (Kalman, 1960;

Welch and Bishop, 1995). The Kalman filter is a recursive inference algorithm that is not sensor-

dependent and allows the uncertainty estimation in both the input information and produced

estimates, both advantages over the sensor-specific fusion algorithms.

The Kalman filter estimates the states of a process (e.g. OFR surface area) by combining

observations, a state-transition model (e.g. linear), and the respective uncertainties of these

elements. The algorithm is implemented in two steps to estimate the state of a process by

60
minimizing the mean of the squared errors. First, the algorithm uses a state-transition model to

predict estimates of the current state variables, along with their uncertainties. Second, these

predictions are updated with new observations and their uncertainties using a weighted average,

with more weight assigned to estimates with less uncertainty (Welch and Bishop, 1995). The

Kalman filter allows for the uncertainties and the model estimates to be evolving randomly rather

than deterministically (Kalman, 1960; Welch and Bishop, 1995). This enables constructing time

series of dynamic processes with higher flexibility (Ye et al., 2021) when compared to models that

assume stationary and linearity (e.g. Verbesselt et al., 2012). Other advantages of the Kalman filter

when compared to other algorithms include its suitability to gap fill and smooth time series with

abrupt changes (Ogilvie et al., 2018b), to account for measurement uncertainties (Johnson et al.,

2021; Schwatke et al., 2015; Sedano et al., 2014; Ye et al., 2021; Zhou and Zhong, 2020), to not

rely on explicit parameter tuning making it well-suited to large-scale applications (Moreno-

Martínez et al., 2020); and its high-computational efficiency due to its recursive form (Clark et al.,

2008; Gillijns et al., 2006; Ye et al., 2021).

Recently, the Kalman filter has been applied to a variety of topics including studies in the

passive and active remote sensing spectrum. Several studies have explored the Kalman filter to

monitor Earth's surface processes using Landsat and MODIS imagery. For example, near-real time

forest monitoring using a stochastic continuous change detection (Ye et al., 2021), blending

MODIS and Landsat satellite imagery to produce synthetic Landsat imagery time series and their

uncertainty estimates (Zhou and Zhong, 2020), creating synthetic Normalized Difference

Vegetation Index from Landsat imagery to assess land surface phenology (Sedano et al., 2014),

and near-real time monitoring of insect induced forest defoliation using MODIS (Olsson et al.,

2016). In the active sensor spectrum, studies have used Synthetic Aperture Radar imagery to

61
monitor crop phenology (Vicente-Guijalba et al., 2014), and to assimilate data from multiple

satellite altimetry to monitor water impoundment levels (Schwatke et al., 2015). In addition, the

Kalman filter has been used in surface hydrology studies to gap fill, remove outliers, and smooth

daily water volume time series for reservoirs with surface areas ranging from 180 to 1600 ha

(Ogilvie et al., 2018b). It has also been used in multiple research projects for streamflow and

hydrological data assimilation (Clark et al., 2008; Vrugt et al., 2006; Weerts and El Serafy, 2006).

Building on existing studies, here we propose a novel use of the Kalman filter that

assimilates information from multiple sensors to monitor OFRs' sub-weekly surface area changes.

Specifically, the goals of this manuscript are twofold:

1) To derive OFR surface area using PS, RE, S2, and S1 sensors and to estimate the

uncertainties of each sensor by comparing (a) the OFRs' surface area derived from each sensor to

a validation dataset, and (b) the surface area obtained from RE, S2, and S1 to PS (i.e. the highest

spatial resolution sensor).

2) To describe a new data assimilation algorithm based on the Kalman filter to obtain OFRs'

sub-weekly surface area changes by accounting for the sensors’ uncertainty calculated in the

objective 1.

2. DATA AND METHODS

2.1 Study region and the OFRs dataset


Eastern Arkansas is the third most irrigated region in the USA, with common irrigated

crops including corn, rice, and soybeans (NASS–USDA, 2017). The rapid expansion of irrigated

agriculture during the last 40 years has made eastern Arkansas an area with a high concentration

of OFRs (Shults et al., 2020; Yaeger et al., 2018, 2017), and a dataset that digitally mapped 736

OFRs already exists for this region (Yaeger et al., 2017). The study region has a humid subtropical

62
climate, receiving about 1300 mm year-1 of precipitation, mostly distributed between March and

May, and November and January. All OFRs in the dataset have surface area smaller than 50 ha

and are distributed across eight sub-watersheds (Fig. 1). This dataset was manually mapped using

the high-resolution (1 m) National Agriculture Imagery Program archive in combination with 2015

Google Earth satellite imagery (Yaeger et al., 2018, 2017). The authors of the dataset used the

Google Earth Explorer to sharpen the image details when zooming in, and to provide a validation

for features appearing indistinct or pixelated in the 1-m mosaic imagery. The majority of the OFRs

have a surface area smaller than 20 ha and a shape index (Si) (Polsby and Popper, 1991) higher

than 0.50 (i.e., regular, and squared geometry shapes) (Fig. 1, bar-chart). The Si is calculated as

follows:
𝐴
𝑆𝑖 = 4𝜋 𝑃2 (1)

where Si is the shape index score, A is the OFR surface area in m2, and P is the OFR shape

perimeter in m.

Si is the ratio of the OFR’s area to the area of a circle whose circumference is equal to the

perimeter of the OFR. A Si score equal to 1 means that the OFR geometric shape represents a

perfect circle, and any other geometric shape has a smaller Si. For instance, OFRs with an Si score

equal to 0.785 have squared shapes, OFRs with an Si score equal to 0.589 have rectangular shapes,

and OFRs that have an Si smaller than 0.50 have a fragmented geometry (i.e., more edges and

higher shape complexity). We assigned the OFRs to three different size classes (0.1–5 ha, 5–10

ha, and 15–50 ha), and different Si ranges (0.10–0.50, 0.50–0.75, and 0.75–1.0) based on the

surface area mapped in the OFRs dataset. These classes were used to support the surface area

mapping analyses when accounting for different OFRs’ sizes and shape complexities (Fig. 1).

63
Figure 1–Study region in eastern Arkansas, USA, and the sub-watersheds where the OFRs are

located. The frequency distribution of the OFRs’ surface area (ha) size classes and Si ranges.

2.2 Satellite imagery processing


The workflow used to download and process all satellite imagery to estimate the OFRs’

surface area time series between January 2017 and December 2020—this time period was chosen

based on the data availability—is illustrated in Fig. 2. For PS, S2, RE (i.e. optical sensors), we

estimated the OFRs’ surface area using the Normalized Difference Water Index (NDWI) based on

the green and NIR wavelengths (McFeeters, 1996) (Table 1). For S2, we could have computed the

Modified NDWI (Du et al., 2016; Singh et al., 2015); however, to maintain a consistent

methodology, we decided to compute the same index for all optical sensors. For S1 (i.e. radar), we

estimated surface area from vertical transmit and vertical receive (VV) ascending single co-

64
polarization images. For each OFR time series, we clipped all images available from PS, RE, S2,

and S1 using the OFRs’ shapefile, and total number of image clips used per sensor is shown on

Table 1.

Figure 2–Workflow used to download and process the satellite imagery used to estimate the

OFRs’ surface area time series from PS, RE, S2, and S1 between January 2017 and December

2020.

65
Table 1–Different sensors wavelengths bands (green and NIR, used to calculate NDWI), spatial

resolution, revisit time, and total number of image clips used between January 2017 and December

2020 including all OFRs.

Spatial Revisit time Number of image


Sensor green (nm) NIR (nm) resolution (m) (days) clips used
PS 500–590 780–860 3.125 ~1 177208
RE 520–590 760–850 5 ~ 5.5 36629
S2 543–578 785–899 10 ~5 117432
S1 - - 10 ~6 137609

2.2.1 PlanetScope CubeSat and RapidEye Ortho tiles


We buffered the OFRs’ shapefile by 100 meters to search for and clip all Level 3A surface

reflectance imagery available through the Planet API for PS and RE. We downloaded surface

reflectance Ortho tiles which use a fixed UTM grid system defined in 25 km by 25 km tiles with

1 km overlap (Planet Labs Inc, 2021). We downloaded all images that contained < 10% of cloud-

cover using an image-based filter—this allowed us to download mostly cloud-free images.

However, because this threshold is image-based some useful observations (i.e. when the OFR is

not covered with clouds but the image is filtered out) were not downloaded, decreasing the total

number of observations per OFR. We did not use a lower image-based cloud-cover threshold

because that would considerably decrease the number of observations when searching for imagery.

In addition, because there were some images that still contained some clouds—due to the 10%

image-based threshold—we filtered out all Ortho tiles that had more than 5% of unusable data

(i.e., due to sensor problems, cloud-cover, etc.) employing the unusable data mask provided by

Planet Labs. To deal with potential cloud obscuration outliers for PS, we could have used the latest

cloud-mask (i.e. usable data mask 2) (Planet Labs Inc, 2021); however, this cloud-mask was not

available for the entire time series (2017–2020). For RE, there is no cloud-mask besides the

66
unusable data mask that is offered by Planet Labs for all RE Ortho tiles. The PS Ortho tiles are

resampled to 3.125 m and the RE Ortho tiles are resampled to 5 m, and both Ortho tiles are

projected using the WGS84 datum. In addition, the Ortho tile products are radiometrically, sensor,

and geometrically corrected and aligned to a cartographic map projection. The atmospheric

correction was done by Planet Labs using the 6S radiative transfer model with ancillary data from

MODIS (Kotchenova et al., 2006; Kotchenova and Vermote, 2007; Planet Labs Inc, 2021), and

the positional accuracy has been reported to be on the order of 10 m for both Ortho Tiles (Planet

Labs Inc, 2021).

2.2.2 Sentinel 1 and 2


We processed S1 images at 10 m spatial resolution using Google Earth Engine to obtain

the time series of OFRs’ surface area. We downloaded VV ascending single co-polarization

channel to ensure that images have the same transmit/receive. The S1 images are Level-1 Ground

Range Detected processed to backscatter coefficient in decibels. The backscatter coefficient

represents target backscattering area (i.e. radar cross-section) per unit ground area. The S1 images

were pre-processed using the Sentinel-1 Toolbox (European Space Agency, 2021a) to derive the

backscatter coefficient in each pixel. The pre-processing steps included 1) removing ground range

detected border noise, which filters out low intensity noise and invalid data at scene edges; 2)

removing thermal noise to reduce discontinuities between sub-swaths for scenes in multi-swath

acquisition modes; 3) radiometric calibration using the sensor metadata; and 4) terrain correction

(i.e. orthorectification) using digital elevation models. A common issue with radar products is the

presence of speckle noise caused by interference in the backscatter signal. To reduce speckle noise,

we applied a mean filter (i.e. 4 x 4 m window) to smooth the images; this is a common procedure

67
when pre-processing S1 images and has been previously used when detecting water bodies

(Ahmad et al., 2020; Vanthof and Kelly, 2019).

We used the 100 m buffered OFRs shapefile to search for and clip S2 Level 2A surface

reflectance images at 10 m spatial resolution available on Sentinel Hub API (Sentinel Hub, 2021).

The S2 data was processed using the Sentinel-2 Toolbox (i.e. Sen2Cor processor) (European

Space Agency, 2021b) using the associated Level 1C product (i.e. top of the atmosphere

reflectance) as input. We downloaded all available images with the S2 cloud-mask created using

the open-source s2cloudless Python Package (Sentinel Hub, 2020), and we only used images with

< 5% of cloud-cover and cloud probability > 90%. More accurate cloud masks could be obtained

using Fmask 4 (Sanchez et al., 2020); however, the Fmask 4 was not available through the Sentinel

Hub environment.

2.3 Surface water classification


We used an adaptive Otsu threshold (Otsu, 1979) to separate water from non-water pixels.

We applied the Otsu threshold to the NDWI and the S1 images to calculate the surface area for all

images in the time series. The Otsu threshold has been widely used to calculate the surface area of

inland water bodies (Cooley et al., 2017; Du et al., 2014; Li and Wang, 2015; Liu et al., 2016;

Sheng et al., 2016; Vanthof and Kelly, 2019; Wang et al., 2019). In addition, the Otsu threshold

optimizes the separability of the pixel values, and the efficacy of the threshold is contingent on the

bimodal distribution of the pixel values. A bimodal distribution is ensured by buffering the OFRs

shapefile to 100 meters and using this buffer to clip the satellite imagery thus including an area of

land immediately adjacent to water. After calculating the Otsu threshold and separating water from

non-water pixels, we clipped the images one more time using the OFRs’ shapefile buffered to 20

68
meters. This last step was done to minimize the impact of adjacent water bodies when estimating

the OFRs’ surface area.

2.4 Validation scheme and surface area uncertainties


To validate the OFRs’ surface area classification using the Otsu thresholding approach, we

used 20 orthorectified and multispectral SkySat Collect imagery product (blue: 450–515 nm,

green: 515–595 nm, red: 605–695 nm, and NIR: 740–900 nm) at sub-meter (0.67–0.95 nm) spatial

resolution (Table 2). For each image, we displayed a false color composite to enhance the presence

of water (Red: NIR, Green: green, and Blue: red), and we superimposed the OFRs dataset. Then,

we manually delineated the OFRs’ surface area, which resulted in 452 validation surface areas for

multiple observations in time from 228 different OFRs. In addition, we used SkySat imagery that

were partially covered by clouds (i.e. cloud cover 0–66%, Table 2); however, we only included

the surface area observations when the OFRs were entirely visible. To validate the surface area

obtained from PS, RE, S2 and S1, we assigned the surface area observations from the validation

dataset to three size classes 0.1–5 ha, 5–10 ha, and 10–50 ha, and to different Si ranges 0.10–0.50,

0.50–0.75, and 0.75–1.0. The distribution of the validation dataset into different size and Si ranges

is given in Fig. 3. For the validation scheme, we considered the current state (i.e. when the SkySat

imagery were collected) of the OFRs in terms of area (size class) and shape complexity (Si range).

For instance, the same OFR can be classified into two different size classes or Si ranges depending

on the validation surface area from different dates.

For PS, RE, S2, and S1, we pairwise compared the validation surface area with the surface

area obtained from each sensor. The validation surface area date may not correspond exactly to

the sensor’s dates; therefore, for these pairwise comparisons we used the sensors’ closest

observation in time, which had a maximum difference of three days before or after the validation

69
date. In addition, for each sensor, we distributed the pairwise comparisons into different size

classes and Si ranges according to the validation dataset (Fig. 3), and calculated the coefficient of

determination (r2), and the percent error (Eq. 2), and mean absolute percent error (Eq. 3).
𝑦 𝑖 − 𝑥𝑖
Percent error (%) = ( ) * 100 (2)
𝑥𝑖

1 𝑦 𝑖 − 𝑥𝑖
Mean absolute percent error (%) = 𝑛 𝛴| | ∗ 100 (3)
𝑥𝑖

Where 𝑥𝑖 is the validation surface area, and yi is the surface area derived from each sensor.

Moreover, we assessed the uncertainties in surface area introduced by the coarser spatial

resolution sensors—RE, S2, S1 compared to PS—by matching the OFRs’ surface area for images

collected on the same day. We chose PS as the closest observations to ground-truth due to its

spatial resolution (i.e. 3.125 m) as, in the case of OFRs, spatial resolution is the most important

limiting factor in their detection (Huang et al., 2018; Ogilvie et al., 2018a; Vanthof and Kelly,

2019). We assessed the uncertainties from RE, S2, and S1, accounting for different OFRs size

classes, and for different OFRs shape complexities. For all sensors, we filtered out outliers from

the time series of OFRs’ surface areas by calculating the standard deviation of the surface area

observations for each OFR across the entire time series and filtering out all observations that were

not within + three standard deviations from the mean surface area for a respective OFR.

70
Table 2–SkySat image identification, image acquisition date, number of OFRs surface area

observations per image, cloud clear (indicates the percentage of cloud cover), ground control ratio

(GCR*, defines the image positional accuracy, and values closer to 1 means higher accuracy), and

ground sampling distance (GSP).

Number of Cloud cover GSP


SkySat image Date observations (%) GCR (m)
20180924_193612_ssc8_u0001 9/24/2018 17 40 0.59 0.74
20180928_193847_ssc7_u0001 9/28/2018 17 50 0.71 0.76
20191013_164407_ssc4_u0001 10/13/2019 5 66 0.71 0.86
20191108_194239_ssc8_u0001 11/8/2019 49 11 0.69 0.90
20191217_193016_ssc8_u0001 12/17/2019 38 15 0.69 0.80
20200107_171221_ssc1_u0001 1/7/2020 41 3 0.76 0.72
20200306_193502_ssc10_u0001 3/6/2020 26 15 0.80 0.79
20200306_193502_ssc10_u0002 3/6/2020 3 15 0.81 0.76
20200329_171211_ssc1_u0001 3/29/2020 11 54 0.03 0.85
20200329_171211_ssc1_u0002 3/29/2020 24 54 0.31 0.83
20200416_193328_ssc11_u0001 4/16/2020 11 1 0.77 0.68
20200416_193328_ssc11_u0002 4/16/2020 13 1 0.60 0.68
20200416_193328_ssc11_u0003 4/16/2020 3 1 0.40 0.69
20200424_163836_ssc4_u0001 4/24/2020 8 8 0.72 0.80
20200510_171832_ss01_u0001 5/10/2020 42 13 0.83 0.95
20200929_193409_ssc10_u0001 9/29/2020 35 13 0.91 0.66
20201013_194518_ssc6_u0001 10/13/2020 44 1 0.97 0.73
20201102_193752_ssc9_u0001 11/2/2020 16 0 0.97 0.67
20201102_193752_ssc9_u0002 11/2/2020 10 0 0.96 0.67
20201210_194154_ssc11_u0001 12/10/2020 39 1 0.91 0.68
*SkySat Collect product image is an orthorectified composite of smaller SkySat scenes (Planet
Labs Inc, 2021), and the GCR represents the ratio of individual small scenes that are successfully
rectified.

71
Figure 3–Frequency distribution of the OFRs’ surface area validation dataset derived from SkySat

imagery and grouped into different size classes and Si ranges.

2.5 The Kalman filter approach


We propose a data assimilation algorithm based on the Kalman filter to assimilate data

from PS, RE, S2, and S1 to estimate the OFRs’ sub-weekly surface area time series. The Kalman

filter algorithm updates a state-transition model by weighting the surface area observation

accuracy and predicting the current state to the next time step (i.e. model evolution). The algorithm

works recursively based on a combination of sensor observations, a linear state-transition model,

and the uncertainties of these elements. We used the Kalman filter to obtain a sub-weekly (every

three days) surface area time series between the first day of 2017 and the last day 2020.

We applied the Kalman filter using a smoothing inference problem (i.e. retrospective

analysis), in which the algorithm incorporates future surface area observations into the time series

to estimate the surface area at step t. The surface area from t-1 and t+1 will have the most effect on

t, followed by t-2, t+2 and t-3, t+3, and so on. The smoothing functionality allows us to study the

OFRs’ surface area behavior retrospectively by harmonizing all available information to optimally

72
estimate the OFRs’ surface area at step t. The general Kalman filter formulation is described by

Eq. 4 (observation equation) and Eq. 5 (state or transition equation) for t = 1, 2, ...T:

𝑧𝑡 = 𝐹𝑡 𝑥𝑡 + 𝑣𝑡 , 𝑣𝑡 ~ 𝑁(0, 𝑉𝑡 ) (4)

𝑥𝑡 = 𝐺𝑡 𝑥𝑡−1 + 𝑤𝑡 , 𝑤𝑡 ~ 𝑁(0, 𝑊𝑡 ) (5)

Where 𝑥𝑡−1 and 𝑥𝑡 are model estimates in the previous and present states; 𝑧𝑡 is a p-dimensional

vector at a given state; Ft represents the observation matrix between the current observation and

the current state at t; Gt is a p x p state transition matrix, and it relates to the current and prior states

𝑥𝑡 and 𝑥𝑡−1 ; 𝑣𝑡 and 𝑤𝑡 are Gaussian random variables that represent the surface area observation

uncertainty and the process uncertainty, and 𝑉𝑡 and 𝑊𝑡 are variance matrices of the appropriate

dimension.

The Kalman filter state variables are described by the Markov chain properties, in which

the current state is solely dependent on the previous state and the transition probability (i.e.

probability of going from the last state to the current one). The state variable relates to past values

through the state equation (Eq. 5), and the observations in turn are related to the current state

variables through the observation equation (Eq. 4).

“Prediction” refers to the first step of the Kalman filter implementation. In this study, this

step is a dynamic linear state-transition model that propagates the surface area estimates of the

previous state and its uncertainty to provide surface area estimates of the current state of the model

(Eq. 6) and its uncertainty (Eq. 7).

𝑥𝑡 = G 𝑥̂𝑡−1 (6)

𝑃𝑡 = G 𝑃𝑡−1 + 𝑤𝑡−1 (7)

73
Where 𝑥̂𝑡−1 is the posterior surface area estimate in the previous state; 𝑥𝑡 is the prior estimate of

the current state; 𝑃𝑡−1 is the posterior uncertainty of the previous state, and 𝑃𝑡 is the uncertainty

for the current state.

“Update” is the second step of the filter implementation, and it relates to the update of the

prior surface area estimates (Eq. 8) and their uncertainties (Eq. 9) as new observations become

available. The update is a linear combination of the prior estimate (𝑥𝑡 ) and a weighted difference

between the observations and the prior estimate of the model state. The weights are defined as the

Kalman gain (Eq. 10), defined as the ratio of the uncertainty in the prior and the current

observation, which will give more weight to the model process or to the observation depending on

their uncertainty. For instance, a low Kalman gain means that the observation has large uncertainty

(i.e. high measurement noise, 𝑉𝑡 ), hence, more weight is given to the model process. On the other

hand, a high Kalman gain means a high model process uncertainty (𝑊𝑡 ), and more weight is given

to the new observation.

𝑥̂𝑡 = 𝑥𝑡 + 𝐾𝑡 (𝑧𝑡 − 𝐹𝑥𝑡 ) (8)

𝑃𝑡 = (1 − 𝐾𝑡 𝐹) 𝑃𝑡 (9)

−1
𝐾𝑡 = 𝑃𝑡 𝐹 𝑇 (𝐹𝑃𝑡 𝐹 𝑇 + 𝑉𝑡 ) (10)

Where 𝑥̂𝑡 is the posterior estimate of state, 𝑃𝑡 is the prior surface area uncertainty, and 𝐾𝑡 is the

Kalman gain at current state at step t.

2.6 Kalman filter setup and validation


We implemented the Kalman filter using the dlm: Bayesian and Likelihood Analysis of

Dynamic Linear Models (Petris, 2010) R package. To estimate the initial OFR surface area state

(i.e. first value of the time series), the initial state uncertainty, and the noise process variance (𝑤𝑡 ),

we used the maximum likelihood through the Quasi-Newton numerical searching algorithm

74
(Durbin and Koopman, 2012; Helske, 2016). To account for the observation uncertainty (𝑣𝑡 ) of

each sensor, we used a time-varying variable that changed according to the observation sensor. In

addition, the observation uncertainty varied based on the OFR size and Si range. The uncertainty

was calculated using the variance between each sensor surface area and the validation surface area

(Table 3). Given that S1 may underestimate the surface area of small water bodies by ~18% or

more (López-Caloca et al., 2020; Vanthof and Kelly, 2019), we tested the Kalman filter approach

with and without the S1 observations. Moreover, we used a fixed interval of three days to

implement the Kalman filter algorithm. For each OFR, we added the dates of the satellite imagery

observations to the time series, and the resulting time series for each OFR had a total of ~200–230

surface area estimates per year.

Table 3–Variance used for PS, RE, S2 and S1, when implementing the Kalman filter for OFRs of

different size classes and Si ranges. The variance calculated by comparing each sensors’ surface

area and with the validation surface area.

Class size Si range PS RE S2 S1


variance
0.1–5 0.10–0.50 0.13 0.27 0.84 1.14
0.1–5 0.50–0.75 3.73 2.96 6.08 20.18
0.1–5 0.75–1.0 0.07 0.09 0.28 0.91
5–10 0.10–0.50 1.30 0.87 1.36 3.40
5–10 0.50–0.75 0.05 0.06 0.19 0.86
5–10 0.75–1.0 1.03 1.54 1.23 4.21
10–50 0.10–0.50 1.40 1.57 1.89 8.28
10–15 0.50–0.75 1.07 0.49 1.64 4.43
10–15 0.75–1.0 3.13 2.96 6.08 20.18

We validated the Kalman filter surface area estimates using the SkySat validation dataset,

and the same approach described in Section 2.4. The dates from the validation dataset may or may

not match with a date from the Kalman filter time series, if the latter is true then we used the closest

surface area estimate in time—maximum difference is two days—to compare the validation and

75
Kalman filter surface areas. In addition, we randomly selected 20% of all PS observations, for each

OFR, to withhold from the Kalman filter, and used these observations to evaluate the Kalman filter

performance by calculating the r2 and MAPE between the independent PS observations and the

Kalman filter surface areas estimates. The general workflow to implement the Kalman filter

involved the following steps:

(1) Initialization: estimate the OFR initial surface area, the prior uncertainty (𝑃𝑡 ) associated

with the initial state, and the noise process variance (𝑤𝑡 ) using maximum likelihood.

(2) Prediction: use the estimated parameters from previous step and the state-transition

model (Gt) to predict the OFR surface area at the next time step 𝑥𝑡 | 𝑧𝑡−1 .

(3) Update: obtain a surface area observation and the associated belief about its accuracy

(i.e. measurement uncertainty, 𝑣𝑡 ). Calculate the residual between the estimated state and

measurement, and the scaling factor based on whether the observation or prediction is more

accurate (Eq. 10). Update the belief (𝑥̂𝑡 ) in the state based on the observation uncertainty

(Eq. 8).

(4) Repeat steps (2) and (3) from time 𝑡 = 0 to time 𝑡 = 𝑇.

2.7 Kalman filter analysis and study cases


We applied the Kalman filter approach to all OFRs in the OFRs dataset (Fig. 1), and we

assessed the total surface area variation between January 2017 and December 2020 including all

OFRs. In addition, we selected eight OFRs of different sizes and shape complexities to illustrate

the Kalman filter surface area time series. These OFRs’ outlines are overlaid on high-resolution

Google maps satellite imagery to show where the OFRs are located (Fig. 4). The OFRs’ surface

areas and Si’s are available in Table 4. Also, we estimated the mean weekly surface area variation

(%) from the annual mean for the eight OFRs study cases.

76
Figure 4–OFRs outlines (pink lines) overlaid on high-resolution Google maps satellite imagery.

77
Table 4–OFRs study cases surface area and Si.

OFR Surface area (ha) Si


A 4.13 0.65
B 4.93 0.83
C 13.30 0.78
D 11.66 0.82
E 9.96 0.08
F 10.36 0.19
G 2.94 0.28
H 6.67 0.11

3. RESULTS

3.1 Comparing the OFRs’ surface area with the validation dataset
There was a good agreement (r2 >= 0.77) when pairwise comparing the OFRs’ validation

surface area with the surface area calculated from PS, RE, S2 and S1 (Fig. 5). The MAPE for

surface area observations between 0.1–5 ha and 5–10 ha was similar for PS (6.15% and 7.25%)

and RE (7.73% and 8.39%), whereas higher MAPE values were found for S2 (11.27% and

11.78%). Meanwhile, S1 had the highest MAPE (> 20%) among all sensors for all size classes.

When comparing the different size classes for the same sensor, the largest surface area

observations (10–50 ha) had the highest agreement (r2 > 0.90) and lowest MAPE (< 6% for PS

and RE, 8.06% for S2, and 21.73% for S1). In addition, for most of the surface area validation

observations, PS, RE, S2 and S1 tended to underestimate the OFRs surface area—points located

above the 1:1 line in Fig. 5 scatter plots; this is most visible for surface area between 5–10 ha, and

for S2 and S1. The number of pairwise comparisons varied among different sensors and for

different size classes, and S1 had the highest number of pairwise comparisons followed by PS, S2,

and RE.

78
Figure 5–Linear relationships of the pairwise comparisons between the OFRs’ validation area and

the surface area obtained from PS, RE, S2, and S1 for observations grouped into different size

classes (0.1–5 ha, 5–10 ha, and 10–50 ha).

For PS, RE and S2, when dividing the validation pairwise comparisons into different S i

ranges, the observations with regular geometries (Si > 0.50) had lower percent error variation and

lower MAPE when compared to observations with more complex geometries (Si < 0.50) for all

size classes (Fig. 6). The lowest MAPE values were found for observations with Si between 0.75–

1.0 and area between 10–50 ha for RE (2.47%), PS (2.88%), and S2 (4.24%). Besides S1, which

had MAPE varying between 14–21% between all size classes and Si ranges, the highest MAPE

79
values were found for the smallest observations (0.1–5 ha) with Si between 0.10–0.50 for S2

(12.63%), RE (12.13%), and PS (9.10%) (Fig. 6).

As the surface area observations increased in size, we would expect a decrease in MAPE

(Fig. 5) and percent error variability (Fig. 6). Nonetheless, the observations between 5–10 ha had

higher MAPE when compared to observations between 0.1–5 ha (Fig. 5), and a similar percent

error variability for different Si ranges (Fig. 6). This can be explained by the higher number of

observations with complex geometries in the 5–10 ha size class, which had 43% of the observations

with Si between 0.10–0.50, compared to 20% from 0.1–5 ha (Fig. 3). In addition, S1 showed

greater accuracy for the observations with the most complex geometries (i.e. Si between 0.10–

0.50) for the 0.1–5 ha class size. This is most likely due to the fact that S1 tends to underestimate

the OFRs’ surface area, and therefore, the difference in accuracy between different Si ranges is not

evident for the smallest class size.

Figure 6–Percent error calculated from the pairwise comparisons between the OFRs’ validation

area and the surface area obtained from the PS, RE, S2, and S1 for observations grouped into

different size classes (0.1–5 ha, 5–10 ha, and 10–50 ha) and Si ranges (0.10–0.50, 0.50–0.75, and

0.75–1.0).

80
3.2 Same day pairwise comparisons between PS with RE, S2 and S1
The same day pairwise comparisons between PS with RE, S2 and S1 varied among

different OFRs, and RE, S2, and S1 tended to underestimate the OFRs’ surface area for all size

classes (Fig. 7)—for these analyses, we divided the surface area observations according to the

OFRs area obtained from the OFRs dataset. All pairwise comparisons showed high agreement

between PS and the different sensors (r2 >= 0.88), and the MAPE decreased as the OFRs increased

in size. PS–RE had the highest agreements (r2 > 0.96) and the lowest MAPE between 2.84–4.37%,

followed by PS–S2 with r2 > 0.94 and MAPE between 4.81–8.47%, and PS–S1 with r2 values

between 0.88–0.97 and MAPE between 18.28–21.77%, which had the highest density of points

located the farthest from the 1:1 line (i.e. outliers) (Fig. 7). The number of outliers tended to

increase as the number of pairwise comparisons increased, which impacted the PS–S2 relationship

for all size classes. For instance, for OFRs with surface area of 5–10 ha, PS–S2 (n = 10458) had

two times as many observations as PS–S1 (n = 4813), and almost four times as many as PS–RE (n

= 2751). In addition, the number outliers for the OFRs with surface area between 5–10 ha can be

attributed to the fact that this size class has a higher number of OFRs with complex geometries

when compared to the other classes (see frequency distribution of the OFRs’ Si range in Fig. 1).

81
Figure 7–Surface area same day pairwise comparisons between PS and RE, S2, and S1 for OFRs

grouped into different size classes (0.1–5 ha, 5–10 ha, and 10–50 ha).

PS–RE had the lowest percent error variability and the lowest MAPE (3.17–5.85%)—

including all size classes and for the different Si ranges—followed by PS–S2 (7.32–9.35%) and

PS–S1 (17.91–20.77%) (Fig. 8). The pairwise comparisons showed that RE, S2 and S1 tended to

underestimate the OFRs’ surface area when compared to PS. In addition, the percent error

variability decreased and the MAPE was smaller for OFRs with more regular shapes (higher Si

values), for all pairwise comparisons, and for larger OFRs (Fig. 8).

82
Figure 8–Surface area percent error and MAPE for the same day pairwise comparisons between

PS and RE, S2, and S1 for OFRs grouped into different size classes (0.1–5 ha, 5–10 ha, and 10–

50 ha) and Si ranges (0.10–0.50, 0.50–0.75, and 0.75–1.0).

3.3 Kalman filter algorithm performance analysis


We validated the Kalman filter algorithm by comparing the surface area estimates with the

SkySat validation dataset (see Section 2.4). In addition, we evaluated the Kalman filter algorithm

for two different scenarios, including only the optical sensors (PS, RE, and S2), and including all

sensors (PS, RE, S2, and S1). For both scenarios, there was a good agreement (r2 > 0.83) between

the Kalman filter surface area estimates and the validation surface area. On the other hand, the

Kalman filter had smaller MAPE for all size classes when only the optical sensors were included—

MAPE ranging from 5.68–6.28% compared to 7.49–10.59% when S1 was added to the Kalman

filter (Fig. 9, scatter plots). Overall, the percent error variability showed that the Kalman filter

tended to underestimate the OFRs surface area, and the MAPE was smaller for observations with

regular geometry (Si > 0.50) for all size classes and for both scenarios (i.e. with and without S1).

In addition, the percent error variability and MAPE were smaller when deploying the Kalman filter

without the S1 observations for all size classes and Si ranges (Fig. 9 boxplots). The highest MAPE

values (13.45% and 7.78%) for both scenarios, with and without S1, were found for Si range 0.10–

83
0.50 and area between 0.1–5 ha, and the lowest MAPE values (5.19% and 4.13%) were found for

Si range 0.75–1.0 and area between 10–50 ha (Fig. 9 boxplots).

Figure 9–Linear relationships of the pairwise comparisons between the OFRs’ validation area and

the surface area obtained from the Kalman filter divided into different size classes (0.1–5 ha, 5–

10 ha, and 10–50 ha), and the percent error and MAPE for observations grouped into different size

classes and Si ranges (0.10–0.50, 0.50–0.75, and 0.75–1.0).

To assess the performance of the Kalman filter considering all OFRs, we compared the

Kalman filter surface area estimates with an independent PS subset—20% of the PS observations

that were not used in the Kalman filter—and we calculated the r2 and MAPE distribution

accounting for different size classes (Fig. 10). For both Kalman filter scenarios, the mean r2 values

84
were similar for OFRs of different size classes, and higher r2 values were found when including

only the optical sensors, which had mean r2 between 0.54–0.65 and more than 440 (60%) OFRs

with r2 > 0.50. Meanwhile, when adding the S1 observations, the mean r2 values ranged between

0.43–0.57, and there were 290 (40%) OFRs with r2 > 0.50 (Fig. 10, r2 boxplots and histograms).

In addition, the Kalman filter without the S1 observations had mean MAPE ranging between 4.2–

5.6% for all OFRs size classes, and 590 (80%) of the OFRs had MAPE < 5%. When adding the

S1 observations, the mean MAPE ranged between 6.2–7.5%, and 480 (65%) of the OFRs had

MAPE < 5% (Fig. 10, MAPE boxplots and histograms).

85
Figure 10–The distribution of r2 and MAPE when comparing the Kalman filter surface area

estimates with the independent PS subset—20% of the PS observations that were not used in the

Kalman filter. Boxplots are divided into different size classes accounting for all OFRs in the OFRs

dataset. The boxplots and histograms are colored according to both Kalman filter scenarios, with

and without S1 observations.

3.4 Kalman filter applications


Adding the S1 observations to the Kalman filter resulted in lower agreement and higher

uncertainties (Fig. 9–10) when comparing the surface area estimates with the validation dataset

and the independent PS subset. Hence, we proceed with the Kalman filter application analyses

considering only the optical sensors.

86
The number of satellite observations varied along the year and between different years

influencing the number of observations in the Kalman filter time series (Fig. 11). In general, when

accounting for all OFRs’ surface area time series, PS had the highest number of observations (~7

month-1) followed by S2 (~4 month-1), S1 (~3 month-1), and RE (~2 month-1). Using the Kalman

filter as a data assimilation algorithm, the mean number of observations per OFR increased to ~

18 month-1. By improving the OFRs surface area cadence, the Kalman filter allowed us to

investigate sub-weekly surface area change patterns, for instance, the study cases (Table 4) OFRs’

surface area time series between January 2017 and December 2020 shown on Fig. 12.

The comparisons between the Kalman filter surface area with the independent PS subset

resulted in r2 >= 0.90 and MAPE < 3% for OFRs A–C, and for OFR D, r2 (0.71) and MAPE

(5.11%). Meanwhile the OFRs E, F, and H had r2 varying between 0.12–0.59 and MAPE between

6.62–15.75%, and the OFR G had r2 (0.32) and MAPE (5.45%). The OFRs with the lowest r2 and

the highest MAPE are the ones with higher shape complexity (i.e., E–H, Si < 0.28), and the OFRs

with higher variability in surface area observations between all sensors (e.g., E and H).

The highest uncertainties on the Kalman filter surface area time series occurred when there

were fewer satellite imagery observations (Fig. 12, confidence intervals). For instance, the

beginning (January–February) and end (November–December) of each year are the periods when

there is more cloud obscuration in the study region and, as a consequence, there are fewer satellite

imagery observations available. This pattern is due to the precipitation regime in the study region

(Perin et al., 2021). The uncertainty also varied according to the observation sensor, and it is higher

for S2, RE and PS (Fig. 5–8). In addition, the variability in surface area estimates between all

sensors will influence the Kalman filter. For instance, OFRs E–H, have more scattered surface

87
area observations when compared to OFRs A–D, affecting the performance of the algorithm (lower

r2 and higher MAPE).

Figure 11–Distribution of the number of observations per month for PS, RE, S2, S1, and the

Kalman filter algorithm, considering all OFRs’ surface area between time series between January

2017 and December 2020.

88
Figure 12–Sub-weekly surface area time series obtained from the Kalman filter for the study case

OFRs (see Table 4) between January 2017 and December 2020. Grey shaded area represents +/-

one and two standard deviations. The r2 and MAPE values were derived from the Kalman filter

comparisons with the independent PS subset.

89
90
91
The OFRs study cases’ mean weekly surface area variations from the annual mean—the

annual mean was calculated using all observations between 2017–2020—is represented by the size

and color of the bars in Fig. 13. In addition, the OFRs’ weekly inter-annual variability is

represented by the error bars (i.e. +/- 1 standard deviation considering all observations for a

particular week). In general, the OFRs had positive mean weekly surface area variation during the

first half of the year with the highest values occurring between February–May, and negative values

during the second half of the year with the lowest values occurring between July–October. OFRs

A–C had the highest mean weekly variation (i.e. -15.35% to 15.18%). Meanwhile, the other OFRs

had mean weekly variation > -10% and < 10%. The OFRs’ surface area inter-annual variability—

size and distribution of the error bars—was different among the study cases (e.g. OFRs E and H

had smaller error bars when compared to OFRs A and C), and it varied throughout the year.

92
Figure 13–Mean weekly surface area variation (%) from the annual mean for OFRs study cases

represented by the size and color of the bars. The inter-annual variability is represented by the

error bars, which indicate +/- one standard deviation calculated from all surface area variation

values for a particular week of the year between 2017–2020.

93
94
95
The OFRs’ surface area dynamics are impacted by the study region climate, as farmers are

irrigating during the hottest and driest months of the year (i.e. June–September) (Perin et al., 2021).

OFRs tend to have positive mean weekly surface area variations when the region receives most of

its precipitation between February–May, and negative variations when the highest temperatures

are recorded between June–September (Fig. 13). In addition, the inter-annual variability (i.e. size

of error bars in Fig. 13) tend to be larger when farmers are irrigating (i.e. June–September) and the

OFRs are at the highest usage rate, as farmers can have different irrigation schedule in different

years based on the crop water requirement.

The total weekly surface area for all OFRs—calculated by summing the OFRs’ weekly

surface area time series—between January 2017 and December 2020 followed the same pattern

observed for the study cases (i.e. the highest values during the first half of the year, and the lowest

values during the second half of the year) for the three size classes and different S i ranges (Fig.

14). The peak of the total surface area was ~ 8500 ha between May and June (i.e. including all size

classes and Si ranges), with the OFRs with surface area between 10–50 ha being responsible for

more than 75% of the total surface area, followed by 5–10 ha (~18%) and 0.1–5 ha (~ 7%). The

lowest total surface area was ~ 6100 ha between August and October. In general, the OFRs’ surface

area seasonality contributed to change the total surface area in ~ 2000 ha each year.

96
Figure 14–Kalman filter total weekly surface area variability considering all OFRs grouped into

different size classes (0.1–5 ha, 5–10 ha, and 10–50 ha) and Si ranges (0.10–0.50, 0.50–0.75, and

0.75–1.0). Y-axis for the three size classes is scaled differently due to the different magnitude of

the total surface area for the larger OFRs.

97
4. DISCUSSION

4.1 OFRs surface area uncertainties


The satellite imagery spatial resolution is the most important limiting factor when mapping

OFRs (Huang et al., 2018; Ogilvie et al., 2018a; Vanthof and Kelly, 2019). When comparing the

surface area classification from each sensor with the validation dataset, we found the smallest

uncertainties for PS (MAPE < 9.10%), RE (MAPE < 12.13%), S2 (MAPE < 12.67%), and S1

(MAPE < 23.98%), for all size classes and for different Si ranges (Fig. 5–6). This highlights the

importance of the recent development of high spatial resolution satellite imagery and data

availability to improve OFRs monitoring. In this regard, our study is one of a few to thoroughly

assess the PS imagery uncertainties using OFRs of different sizes and shape complexities for

multiple images across time. In addition, the same assessment was done for RE, S2, and S1, and

the same day pairwise comparisons encompass tens of thousands of data points to describe the

uncertainties in the surface area introduced by the coarser spatial resolution sensors—RE (MAPE

< 5.85%), S2 (MAPE < 9.35%), S1 (MAPE < 20.70%) compared to PS (Fig. 8).

We found that the uncertainties in surface area tend to decrease for larger OFRs and for

OFRs with regular shapes (Si > 0.50)—these OFRs presented the lowest percent error variation,

and the lowest MAPE (Fig. 5–8). Previous research that used these sensors to monitor small water

bodies, and assessed their uncertainties are either limited by the number of satellite imagery used

(e.g. normally a few images during a single year), and/or the number of water bodies used for

validation (e.g. <100) (Cooley et al., 2017; López-Caloca et al., 2020; Mishra et al., 2020; Ogilvie

et al., 2020; Vanhellemont, 2019; Vanthof and Kelly, 2019). Specifically for PS and RE, our

analysis showed that the uncertainties from these sensors are < 12%, and these findings are aligned

with previous research that mapped small water bodies using PS and RE in the Yokun Flats, a

98
large sub-Arctic wetland in north central Alaska (Cooley et al., 2017). In addition, our findings

related to S2 are aligned with previous research that found mean surface area error between ~ 4–

15% for a floodplain in Senegal (Ogilvie et al., 2020). Ultimately, our study is relevant to future

research that aims to use PS, RE, S2 and S1 to monitor small water bodies in other study regions,

as our findings could be used as a proxy to estimate surface area errors given that assessing the

uncertainties in surface area derived from satellite imagery is not always possible due to constraints

in imagery availability and lack of a ground-truth datasets.

Sentinel 1 measurements of water bodies are independent of cloud presence, and S1 has

proven to be a reliable sensor to monitor large water impoundments (>100 ha) (Bonnema and

Hossain, 2019) and to study surface water dynamics (Ahmad et al., 2020; Chini et al., 2020; Pham-

Duc et al., 2017). Yet, our analyses showed that using S1 to monitor OFRs—which tend to be

small water bodies <50 ha—should be carefully evaluated. S1 underestimates the OFRs’ surface

area by ~20% with large percent error variations when compared to the validation dataset (Fig. 5–

6) and when compared with PS (Fig. 7–8). These findings corroborate previous research that used

S1 to assess OFRs in India; the authors found that S1 underestimates OFRs’ surface area by ~18%

(Vanthof and Kelly, 2019). Another study showed that S1 underestimated the surface area of small

water bodies (1.5–3.3 ha) by 10–30% (López-Caloca et al., 2020). The reasons contributing to the

uncertainties in surface area derived from S1 are related to the large variability in S1 images

backscatter intensity, which depends on the OFRs’ surface water roughness. This is an issue

because water roughness can be strongly influenced by wind over space and time. In addition, the

efficacy of the Otsu threshold is contingent on the bimodal distribution of the S1 imagery pixel

values (i.e. water and non-water pixels); however, in the absence of vegetation surrounding the

OFRs the contrast between water and non-water pixels is less intense which makes the separability

99
of water/non-water challenging when using the Otsu threshold (Liebe et al., 2009; Vanthof and

Kelly, 2019).

A key component of the same day pairwise comparisons analyses is our filter to remove

outliers caused by cloud obscuration and sensor related issues. We flagged an outlier if the

observation was not within + three standard deviations of the mean surface area calculated using

the entire OFRs’ surface area time series for each sensor. This approach enables us to remove most

of the outliers; however, this filter is limited when applied to OFRs that have extremely large

variations in surface area in a short period of time. For instance, there are OFRs with 70–80%

surface area changes in less than a month, which leads to high standard deviations meaning that

some outliers are not flagged. In addition, some outliers are introduced due to the 5–10% cloud-

cover threshold used for PS, RE and S2. Even though we tried to account for problematic PS and

RE images using the unusable data mask offered by Planet Labs, this mask is known for its

uncertainties and limitations in detecting clouds (Mishra et al., 2020). For S2, we used the latest

cloud-mask product available (i.e. s2cloudless); nonetheless, the cloud-mask is not free of errors,

and it carries uncertainties of its own (Sentinel Hub, 2020).

4.2 Kalman filter as a data assimilation algorithm


We presented a novel multi-sensor approach to monitor OFRs sub-weekly surface area

changes that can assimilate optical and radar satellite data while accounting for the uncertainties

in both the satellite observations and the resulting estimates. The algorithm developed in this study

is not sensor-dependent, and data from upcoming satellites could be seamlessly added to the time

series. Beyond improving the OFRs observation cadence, the Kalman filter leverages a Bayesian

inference approach to estimate the surface area uncertainties. These uncertainties reflect the quality

of the surface area estimates, and highlights the advantage the Kalman filter when compared to

100
other sensor-specific algorithms (e.g., Gao et al., 2006; Houborg and McCabe, 2018; Zhu et al.,

2016, 2010) that do not provide uncertainty estimates for the output values.

Given that all sensors tended to underestimate the OFRs surface area, the Kalman filter,

likewise, underestimated the OFRs surface area (Fig. 9), and the uncertainties were smaller for

larger OFRs, and for OFRs with regular geometries (Si > 0.50). The surface area estimates derived

from the Kalman filter including only the optical sensors resulted in high agreement (r2 > 0.95)

and small uncertainties (MAPE 4–8%) when compared to the validation dataset. On the other hand,

higher uncertainties (MAPE 5–14%) were found when adding S1 to the Kalman filter—this is

related to the higher uncertainties found for S1 surface area observations (see section 3.1 and 3.2).

In addition, the comparisons between the Kalman filter results with the PS subset resulted in 440

(60%) OFRs with r2 > 0.50, and 590 (80%) OFRs with MAPE < 5%, compared to 290 (40%) and

480 (65%) OFRs when including S1 in the Kalman filter (Fig. 10). Even though the Kalman filter

with the S1 observations resulted in high uncertainties, the methods proposed in this study offer a

robust and flexible algorithm to combine optical and radar observations. This is relevant if more

accurate S1 surface area estimates are available, for instance, by employing different image

classification approaches (e.g. machine learning algorithms) (Ahmad et al., 2020; DeVries et al.,

2017; Hardy et al., 2019; Huang et al., 2018).

For OFRs that have satellite observations following a well-defined yearly seasonality (e.g.

Fig. 12, OFRs A–D) the Kalman filter was able to simulate the surface area changes accounting

for periods when the OFRs were at low and high capacities in terms of surface area. For these

OFRs, the satellite observations are not scattered with high surface area variability in different

periods of the year, as it is observed for OFRs E, F, and H (Fig. 12). For OFRs E–G, the Kalman

filter is affected by the variability between the sensors surface area estimates resulting in a less

101
evident seasonality and surface area changes. Furthermore, aside from accounting for the

uncertainty of each sensor, the Kalman filter approach allows for the combination of multiple

sensors by attenuating the impact of possible outliers in the time series (Fig. 12, see examples for

OFRs A, B, and E). The algorithm accounts for the sensor uncertainty from the newly added

observation, and it compares the observed surface area with the predicted surface area to smooth

out outliers—this a key application of the Kalman filter (Ogilvie et al., 2018b; Schwatke et al.,

2015). Different sensors can have divergent estimates of OFRs’ surface areas for multiple

observations close in time, and an individual sensor can introduce a negative bias on the OFRs

surface area time series (e.g., S2 observations for OFRs study cases in Fig. 12 E, G, and H). The

Kalman filter proposed in this study allows us to mitigate the impact of outliers and the bias of an

individual sensor, to assimilate data from multiple sensors into a unique surface area time series,

and to measure the uncertainty of each surface area estimate. This highlights the advantages of the

algorithm when compared to other studies (e.g. Vanthof and Kelly, 2019) that used a multi-sensor

approach to monitor OFRs without a data assimilation algorithm.

The Kalman filter performed poorly when there is high variability between the sensors’

surface area, and the factors contributing to surface area variability include: 1) OFRs divided into

several parts (e.g. Fig. 9, OFR H), as there are chances that different parts of the OFRs are

inundated in different periods of the year; 2) Old OFRs (e.g. > 30 years) tend to have a less defined

boundary, caused by erosion and low levee maintenance (e.g., Fig. 9, OFRs E–H). The erosion

process will likely impact the amount of soil that accumulates inside the OFRs thus providing

substrate for growing vegetation inside the OFRs, which are shallow water bodies ~3–5 meters

(e.g., Fig. 9 F, G); 3) The presence of vegetation inside the OFRs or their surroundings (e.g., Fig.

9, OFRs E–G) will contribute to increase the OFRs’ surface area variability. This is a known issue

102
that contributes to surface area inaccuracies when mapping water bodies with spectral mixing

between water and vegetation (DeVries et al., 2017; Hondula et al., 2021); and 4) The presence of

adjacent water bodies (i.e. not part of the OFR shapefile) may also impact the OFRs’ surface area

dynamic. During periods of precipitation and floods, the adjacent water bodies can merge with the

OFRs to form one large water body that can influence the surface area mapping.

4.3 Research impact and applications


The OFRs’ sub-weekly surface area time series (Fig. 11–12) is key to calculate the water

volume change, which can be done by combining the surface area with area-elevation equations—

these equations describe the OFRs’ bathymetry, and are derived using a digital elevation model

collected when the OFRs are empty (i.e. exposed bathymetry) (Avisse et al., 2017; Yao et al.,

2018; Zhang et al., 2016). The OFRs’ volume change time series (i.e. OFR inflows and outflows)

is valuable information to state and federal agencies in eastern Arkansas as it provides the spatial

and temporal variability of the water stored from a network of OFRs. This information enables the

monitoring of water use trends (Fig. 12–14), and allows the agencies to modify water irrigation

management recommendations aiming to increase water use efficiency. In addition, monitoring

the OFRs’ volume change across space and time enables these agencies to develop water status

reports during the most critical periods of the year (i.e. when farmers are irrigating, Fig. 13–14).

These reports can be used to update regional water management plans based on different regions

water availability status, for example, by assessing the total surface area variation accounting for

all OFRs in the study region (Fig. 14).

Our findings have important implications to surface hydrology modeling as volume change

is the most crucial information needed to estimate the OFRs’ cumulative impact on surface

hydrology (Fowler et al., 2015; Habets et al., 2018); however, volume change is rarely available

103
for individual OFRs. This knowledge gap led modelers to frequently assume that volume change

varies equally for all OFRs, including OFRs of different sizes and shape complexities in a given

watershed. Rather than relying on this faulty assumption, future studies could use the Kalman filter

approach presented in this study to estimate volume change utilizing multi-sensors and couple this

information into hydrological models (Evenson et al., 2018; Habets et al., 2014; Hughes and

Mantel, 2010; Perrin et al., 2012; Yang et al., 2020). This is relevant across the globe as there has

been a proliferation of OFRs in different parts of the world, for example, in France (Habets et al.,

2018, 2014) and Brazil (Althoff et al., 2020; Krol et al., 2011; Rodrigues et al., 2012), with still

little known about the impact of OFRs on surface hydrology (Arvor et al., 2018; Habets et al.,

2018; Krol et al., 2011).

5. FUTURE IMPROVEMENTS

We presented a robust and flexible algorithm based on the Kalman filter that can combine

optical and radar observations to monitor OFRs. The data from upcoming satellites could be

seamlessly integrated in the algorithm. Nonetheless, one should be aware that this integration is

contingent on having the uncertainties of the surface area observations, including OFRs of

different sizes and Si ranges, from the new satellites. This poses limitations when applying the

algorithm if an updated validation dataset is not available. Future work and methodology

improvements should focus on improving the Kalman filter estimates of the OFRs’ surface area

seasonality. Initially, we added a yearly seasonal component to the Kalman filter model; however,

the model did not see improvements (e.g. performing equally well or worse than the model without

the seasonal component), therefore we did not include seasonality in our final model. We believe

that improvements in estimating the OFRs’ seasonality may be achieved by accounting for the

surface area rate of change (e.g., monthly, or weekly), and employing the rate of change to estimate

104
the seasonality for each OFR individually. In addition, the system state-transition model used in

this study could be improved by accounting for climate variables (e.g., rate of evaporation,

precipitation) and anthropogenic drivers (e.g. OFR irrigation schedule). We used a dynamic linear

model as a system state-transition model that performed well for most of the OFRs (Fig. 9–10);

however, this approach simplifies the changes in the OFRs’ surface area behavior. For instance,

evaporation losses can reduce OFRs water storage, especially in dry critical periods of the year

when farmers are frequently irrigating (Ibrahim et al., 2015; Liebe et al., 2005), and evaporation

losses can be estimated across time and space using satellite imagery (Zhao and Gao, 2019).

Furthermore, surface phenology of the vegetation surrounding the OFRs can be an indication of

the anthropogenic drivers on the OFRs’ surface area changes—the water stored by the OFRs is

normally used to irrigate closely located cropland (~ 1 km radius)—and sub-weekly surface

phenology at 30 m spatial resolution can be obtained from the latest Harmonized Landsat Sentinel

product available for North America (Bolton et al., 2020).

6. CONCLUSION

We presented a novel multi-sensor approach based on the Kalman filter to monitor OFRs'

sub-weekly surface area changes that can assimilate optical and radar satellite data while

accounting for the uncertainties in both the satellite observations and the resulting estimates. We

used more than 450,000 satellite image clips to create sub-weekly surface area time series for 736

OFRs of different sizes and shape complexities. By improving the surface area observations

cadence and providing the surface area uncertainties, the approach presented in this study has the

potential to enhance water conservation plans by allowing better assessment and management of

OFRs. In addition, we demonstrated the use of the Kalman filter to monitor water use trends at

regional level—this information can be used for water irrigation management recommendations

105
aiming to increase water use efficiency. Lastly, with the impacts of the OFRs on surface hydrology

emerging as an issue across the globe, it is imperative that a multi-sensor approach is implemented

to remotely monitor the OFRs' surface area at a high resolution across space and time to improve

freshwater management.

106
REFERENCES

Ahmad, S.K., Hossain, F., Eldardiry, H., Pavelsky, T.M., 2020. A Fusion Approach for Water
Area Classification Using Visible, Near Infrared and Synthetic Aperture Radar for South
Asian Conditions. IEEE Trans. Geosci. Remote Sens. 58, 2471–2480.
https://doi.org/10.1109/TGRS.2019.2950705

Althoff, D., Rodrigues, L.N., da Silva, D.D., 2020. Impacts of climate change on the evaporation
and availability of water in small reservoirs in the Brazilian savannah. Clim. Change 159,
215–232. https://doi.org/10.1007/s10584-020-02656-y

Arvor, D., Daher, F.R.G., Briand, D., Dufour, S., Rollet, A.J., Simões, M., Ferraz, R.P.D., 2018.
Monitoring thirty years of small water reservoirs proliferation in the southern Brazilian
Amazon with Landsat time series. ISPRS J. Photogramm. Remote Sens. 145, 225–237.
https://doi.org/10.1016/j.isprsjprs.2018.03.015

Avisse, N., Tilmant, A., Müller, M.F., Zhang, H., 2017. Monitoring small reservoirs’ storage
with satellite remote sensing in inaccessible areas. Hydrol. Earth Syst. Sci. 21, 6445–6459.
https://doi.org/10.5194/hess-21-6445-2017

Berg, M.D., Popescu, S.C., Wilcox, B.P., Angerer, J.P., Rhodes, E.C., Mcalister, J., Fox, W.E.,
2016. Small farm ponds: Overlooked features with important impacts on watershed
sediment transport. J. Am. Water Resour. Assoc. 52, 67–76. https://doi.org/10.1111/1752-
1688.12369

Bioresita, F., Puissant, A., Stumpf, A., Malet, J.-P., 2019. Fusion of Sentinel-1 and Sentinel-2
image time series for permanent and temporary surface water mapping. Int. J. Remote Sens.
40, 9026–9049. https://doi.org/10.1080/01431161.2019.1624869

Bolton, D.K., Gray, J.M., Melaas, E.K., Moon, M., Eklundh, L., Friedl, M.A., 2020. Continental-
scale land surface phenology from harmonized Landsat 8 and Sentinel-2 imagery. Remote
Sens. Environ. 240, 111685. https://doi.org/10.1016/j.rse.2020.111685

Bonnema, M., Hossain, F., 2019. Assessing the Potential of the Surface Water and Ocean
Topography Mission for Reservoir Monitoring in the Mekong River Basin. Water Resour.
Res. 55, 444–461. https://doi.org/10.1029/2018WR023743

Casadei, S., Di Francesco, S., Giannone, F., Pierleoni, A., 2019. Small reservoirs for a
sustainable water resources management. Adv. Geosci. 49, 165–174.
https://doi.org/10.5194/adgeo-49-165-2019

Chini, M., Pelich, R., Hostache, R., Matgen, P., Bossung, C., Campanella, P., Rudari, R., Bally,
P., 2020. Systematic and Automatic Large-Scale Flood Monitoring System Using Sentinel-
1 SAR Data, in: IGARSS 2020 - 2020 IEEE International Geoscience and Remote Sensing
Symposium. pp. 3251–3254. https://doi.org/10.1109/IGARSS39084.2020.9323428

Clark, M.P., Rupp, D.E., Woods, R.A., Zheng, X., Ibbitt, R.P., Slater, A.G., Schmidt, J.,
Uddstrom, M.J., 2008. Hydrological data assimilation with the ensemble Kalman filter: Use

107
of streamflow observations to update states in a distributed hydrological model.
https://doi.org/10.1016/j.advwatres.2008.06.005

Cooley, S., Smith, L., Stepan, L., Mascaro, J., 2017. Tracking Dynamic Northern Surface Water
Changes with High-Frequency Planet CubeSat Imagery. Remote Sens. 9, 1306.
https://doi.org/10.3390/rs9121306

DeVries, B., Huang, C., Lang, M., Jones, J., Huang, W., Creed, I., Carroll, M., 2017. Automated
Quantification of Surface Water Inundation in Wetlands Using Optical Satellite Imagery.
Remote Sens. 9, 807. https://doi.org/10.3390/rs9080807

Döll, P., Fiedler, K., Zhang, J., 2009. Global-scale analysis of river flow alterations due to water
withdrawals and reservoirs. Hydrol. Earth Syst. Sci. 13, 2413–2432.
https://doi.org/10.5194/hess-13-2413-2009

Downing, J.A., 2010. Emerging global role of small lakes and ponds: little things mean a lot.
Limnetica 29, 9–24. https://doi.org/10.23818/limn.29.02

Downing, J.A., Prairie, Y.T., Cole, J.J., Duarte, C.M., Tranvik, L.J., Striegl, R.G., McDowell,
W.H., Kortelainen, P., Caraco, N.F., Melack, J.M., Middelburg, J.J., 2006. The global
abundance and size distribution of lakes, ponds, and impoundments. Limnol. Oceanogr. 51,
2388–2397. https://doi.org/10.4319/lo.2006.51.5.2388

Du, Y., Zhang, Y., Ling, F., Wang, Q., Li, W., Li, X., 2016. Water Bodies’ Mapping from
Sentinel-2 Imagery with Modified Normalized Difference Water Index at 10-m Spatial
Resolution Produced by Sharpening the SWIR Band. Remote Sens. 8, 354.
https://doi.org/10.3390/rs8040354

Du, Z., Li, W., Zhou, D., Tian, L., Ling, F., Wang, H., Gui, Y., Sun, B., 2014. Analysis of
Landsat-8 OLI imagery for land surface water mapping. Remote Sens. Lett. 5, 672–681.
https://doi.org/10.1080/2150704X.2014.960606

Durbin, J., Koopman, S.J., 2012. Time series analysis by state space methods. Oxford university
press.

European Space Agency, 2021a. The Sentinel-1 Toolbox. URL


https://sentinel.esa.int/web/sentinel/toolboxes/sentinel-1 (accessed 4.20.21).

European Space Agency, 2021b. Sentinel-2 MSI Technical Guide. URL


https://sentinel.esa.int/web/sentinel/technical-guides/sentinel-2-msi (accessed 4.20.21).

Evenson, G.R., Jones, C.N., McLaughlin, D.L., Golden, H.E., Lane, C.R., DeVries, B.,
Alexander, L.C., Lang, M.W., McCarty, G.W., Sharifi, A., 2018. A watershed-scale model
for depressional wetland-rich landscapes. J. Hydrol. X 1, 100002.
https://doi.org/10.1016/j.hydroa.2018.10.002

Fowler, K., Morden, R., Lowe, L., Nathan, R., 2015. Advances in assessing the impact of hillside
farm dams on streamflow. Australas. J. Water Resour. 19, 96–108.

108
https://doi.org/10.1080/13241583.2015.1116182

Gao, F., Masek, J., Schwaller, M., Hall, F., 2006. On the blending of the Landsat and MODIS
surface reflectance: predicting daily Landsat surface reflectance. IEEE Trans. Geosci.
Remote Sens. 44, 2207–2218. https://doi.org/10.1109/TGRS.2006.872081

Gillijns, S., Mendoza, O.B., Chandrasekar, J., Moor, B.L.R. De, Bernstein, D.S., Ridley, A.,
2006. What is the ensemble Kalman filter and how well does it work?, in: 2006 American
Control Conference. pp. 6 pp.-. https://doi.org/10.1109/ACC.2006.1657419

Habets, F., Molénat, J., Carluer, N., Douez, O., Leenhardt, D., 2018. The cumulative impacts of
small reservoirs on hydrology: A review. Sci. Total Environ.
https://doi.org/10.1016/j.scitotenv.2018.06.188

Habets, F., Philippe, E., Martin, E., David, C.H., Leseur, F., 2014. Small farm dams: Impact on
river flows and sustainability in a context of climate change. Hydrol. Earth Syst. Sci. 18,
4207–4222. https://doi.org/10.5194/hess-18-4207-2014

Hardy, A., Ettritch, G., Cross, D., Bunting, P., Liywalii, F., Sakala, J., Silumesii, A., Singini, D.,
Smith, M., Willis, T., Thomas, C., 2019. Automatic Detection of Open and Vegetated
Water Bodies Using Sentinel 1 to Map African Malaria Vector Mosquito Breeding Habitats.
Remote Sens. 11, 593. https://doi.org/10.3390/rs11050593

Helske, J., 2016. KFAS: Exponential family state space models in R. arXiv Prepr.
arXiv1612.01907.

Hondula, K.L., DeVries, B., Jones, C.N., Palmer, M.A., 2021. Effects of Using High Resolution
Satellite-Based Inundation Time Series to Estimate Methane Fluxes From Forested
Wetlands. Geophys. Res. Lett. 48, e2021GL092556. https://doi.org/10.1029/2021GL092556

Houborg, R., McCabe, M.F., 2018. A Cubesat enabled Spatio-Temporal Enhancement Method
(CESTEM) utilizing Planet, Landsat and MODIS data. Remote Sens. Environ. 209, 211–
226. https://doi.org/10.1016/j.rse.2018.02.067

Huang, C., Chen, Y., Zhang, S., Wu, J., 2018. Detecting, Extracting, and Monitoring Surface
Water From Space Using Optical Sensors: A Review. Rev. Geophys. 56, 333–360.
https://doi.org/https://doi.org/10.1029/2018RG000598

Huang, W., DeVries, B., Huang, C., Lang, M., Jones, J., Creed, I., Carroll, M., 2018. Automated
Extraction of Surface Water Extent from Sentinel-1 Data. Remote Sens. 10, 797.
https://doi.org/10.3390/rs10050797

Hughes, D.A., Mantel, S.K., 2010. Estimation des incertitudes lors de la simulation des impacts
de petites retenues agricoles sur les régimes d’écoulement en Afrique du Sud. Hydrol. Sci.
J. 55, 578–592. https://doi.org/10.1080/02626667.2010.484903

Ibrahim, B., Wisser, D., Barry, B., Fowe, T., Aduna, A., 2015. Hydrological predictions for
small ungauged watersheds in the Sudanian zone of the Volta basin in West Africa. J.

109
Hydrol. Reg. Stud. 4, 386–397. https://doi.org/10.1016/j.ejrh.2015.07.007

Johnson, M.C., Reich, B.J., Gray, J.M., 2021. Multisensor fusion of remotely sensed vegetation
indices using space-time dynamic linear models. J. R. Stat. Soc. Ser. C (Applied Stat. 70,
793–812. https://doi.org/https://doi.org/10.1111/rssc.12495

Jones, J.W., 2019. Improved automated detection of subpixel-scale inundation-revised Dynamic


Surface Water Extent (DSWE) partial surface water tests. Remote Sens. 11.
https://doi.org/10.3390/rs11040374

Jones, J.W., 2015. Efficient wetland surface water detection and monitoring via landsat:
Comparison with in situ data from the everglades depth estimation network. Remote Sens.
7, 12503–12538. https://doi.org/10.3390/rs70912503

Jones, S.K., Fremier, A.K., DeClerck, F.A., Smedley, D., Pieck, A.O., Mulligan, M., 2017. Big
data and multiple methods for mapping small reservoirs: Comparing accuracies for
applications in agricultural landscapes. Remote Sens. 9. https://doi.org/10.3390/rs9121307

Kalman, R.E., 1960. A new approach to linear filtering and prediction problems.

Kotchenova, S.Y., Vermote, E.F., 2007. Validation of a vector version of the 6S radiative
transfer code for atmospheric correction of satellite data Part II Homogeneous Lambertian
and anisotropic surfaces. Appl. Opt. 46, 4455. https://doi.org/10.1364/AO.46.004455

Kotchenova, S.Y., Vermote, E.F., Matarrese, R., Klemm, Jr., F.J., 2006. Validation of a vector
version of the 6S radiative transfer code for atmospheric correction of satellite data Part I:
Path radiance. Appl. Opt. 45, 6762. https://doi.org/10.1364/AO.45.006762

Krol, M.S., de Vries, M.J., van Oel, P.R., de Araújo, J.C., 2011. Sustainability of Small
Reservoirs and Large Scale Water Availability Under Current Conditions and Climate
Change. Water Resour. Manag. 25, 3017–3026. https://doi.org/10.1007/s11269-011-9787-0

Li, J., Wang, S., 2015. An automatic method for mapping inland surface waterbodies with
Radarsat-2 imagery. Int. J. Remote Sens. 36, 1367–1384.
https://doi.org/10.1080/01431161.2015.1009653

Liebe, J., van de Giesen, N., Andreini, M., 2005. Estimation of small reservoir storage capacities
in a semi-arid environment. Phys. Chem. Earth 30, 448–454.
https://doi.org/10.1016/j.pce.2005.06.011

Liebe, J.R., van de Giesen, N., Andreini, M., Walter, M.T., Steenhuis, T.S., 2009. Determining
watershed response in data poor environments with remotely sensed small reservoirs as
runoff gauges. Water Resour. Res. 45. https://doi.org/10.1029/2008WR007369

Liu, Z., Yao, Z., Wang, R., 2016. Assessing methods of identifying open water bodies using
Landsat 8 OLI imagery. Environ. Earth Sci. 75, 873. https://doi.org/10.1007/s12665-016-
5686-2

110
López-Caloca, A.A., Escalante-Ramírez, B., Henao, P., 2020. Mapping small and medium-sized
water reservoirs using Sentinel-1A: a case study in Chiapas, Mexico. J. Appl. Remote Sens.
14, 1–24. https://doi.org/10.1117/1.JRS.14.036503

McFeeters, S.K., 1996. The use of the Normalized Difference Water Index (NDWI) in the
delineation of open water features. Int. J. Remote Sens. 17, 1425–1432.
https://doi.org/10.1080/01431169608948714

Mime, M.M., Young, D.W., 1989. The Impact Of Stockwatering Ponds (Stockponds) On Runoff
From Large Arizona Watersheds. JAWRA J. Am. Water Resour. Assoc. 25, 165–173.
https://doi.org/10.1111/j.1752-1688.1989.tb05678.x

Mishra, V., Limaye, A.S., Muench, R.E., Cherrington, E.A., Markert, K.N., 2020. Evaluating the
performance of high-resolution satellite imagery in detecting ephemeral water bodies over
West Africa. Int. J. Appl. Earth Obs. Geoinf. 93, 102218.
https://doi.org/10.1016/j.jag.2020.102218

Moreno-Martínez, Á., Izquierdo-Verdiguier, E., Maneta, M.P., Camps-Valls, G., Robinson, N.,
Muñoz-Marí, J., Sedano, F., Clinton, N., Running, S.W., 2020. Multispectral high
resolution sensor fusion for smoothing and gap-filling in the cloud. Remote Sens. Environ.
247, 111901. https://doi.org/10.1016/j.rse.2020.111901

National Agricultural Statistics Service–U.S. Department of Agriculture (NASS–USDA), 2017.


Census of agriculture. Washington, DC. https://www.nass.usda.gov/AgCensus/

Ogilvie, A., Belaud, G., Massuel, S., Mulligan, M., Le Goulven, P., Calvez, R., 2018a. Surface
water monitoring in small water bodies: Potential and limits of multi-sensor Landsat time
series. Hydrol. Earth Syst. Sci. 22, 4349–4380. https://doi.org/10.5194/hess-22-4349-2018

Ogilvie, A., Belaud, G., Massuel, S., Mulligan, M., Le Goulven, P., Malaterre, P.-O., Calvez, R.,
2018b. Combining Landsat observations with hydrological modelling for improved surface
water monitoring of small lakes. J. Hydrol. 566, 109–121.
https://doi.org/10.1016/j.jhydrol.2018.08.076

Ogilvie, A., Poussin, J.-C., Bader, J.-C., Bayo, F., Bodian, A., Dacosta, H., Dia, D., Diop, L.,
Martin, D., Sambou, S., 2020. Combining Multi-Sensor Satellite Imagery to Improve Long-
Term Monitoring of Temporary Surface Water Bodies in the Senegal River Floodplain.
Remote Sens. 12, 3157. https://doi.org/10.3390/rs12193157

Olsson, P.-O., Lindström, J., Eklundh, L., 2016. Near real-time monitoring of insect induced
defoliation in subalpine birch forests with MODIS derived NDVI. Remote Sens. Environ.
181, 42–53. https://doi.org/10.1016/j.rse.2016.03.040

Otsu, N., 1979. A Threshold Selection Method from Gray-Level Histograms. IEEE Trans. Syst.
Man. Cybern. 9, 62–66. https://doi.org/10.1109/TSMC.1979.4310076

Pekel, J.F., Cottam, A., Gorelick, N., Belward, A.S., 2016. High-resolution mapping of global
surface water and its long-term changes. Nature 540, 418–422.

111
https://doi.org/10.1038/nature20584

Pena-Regueiro, J., Sebastiá-Frasquet, M.-T., Estornell, J., Aguilar-Maldonado, J.A., 2020.


Sentinel-2 Application to the Surface Characterization of Small Water Bodies in Wetlands.
Water 12, 1487. https://doi.org/10.3390/w12051487

Perin, V., Tulbure, M.G., Gaines, M.D., Reba, M.L., Yaeger, M.A., 2021. On-farm reservoir
monitoring using Landsat inundation datasets. Agric. Water Manag. 246, 106694.
https://doi.org/10.1016/j.agwat.2020.106694

Perrin, J., Ferrant, S., Massuel, S., Dewandel, B., Maréchal, J.C., Aulong, S., Ahmed, S., 2012.
Assessing water availability in a semi-arid watershed of southern India using a semi-
distributed model. J. Hydrol. 460–461, 143–155.
https://doi.org/10.1016/j.jhydrol.2012.07.002

Pham-Duc, B., Prigent, C., Aires, F., 2017. Surface Water Monitoring within Cambodia and the
Vietnamese Mekong Delta over a Year, with Sentinel-1 SAR Observations. Water 9, 366.
https://doi.org/10.3390/w9060366

Pickens, A.H., Hansen, M.C., Hancher, M., Stehman, S. V, Tyukavina, A., Potapov, P.,
Marroquin, B., Sherani, Z., 2020. Mapping and sampling to characterize global inland water
dynamics from 1999 to 2018 with full Landsat time series. Remote Sens. Environ. 243,
111792. https://doi.org/10.1016/j.rse.2020.111792

Pinhati, F.S.C., Rodrigues, L.N., Aires de Souza, S., 2020. Modelling the impact of on-farm
reservoirs on dry season water availability in an agricultural catchment area of the Brazilian
savannah. Agric. Water Manag. 241, 106296. https://doi.org/10.1016/j.agwat.2020.106296

Planet Labs Inc, 2021. Planet Imagery Product Specifications. URL


https://www.planet.com/products/planet-imagery/ (accessed 3.3.21).

Polsby, D.D., Popper, R.D., 1991. The third criterion: Compactness as a procedural safeguard
against partisan gerrymandering. Yale Law Policy Rev. 9, 301–353.

Renwick, W.H., Smith, S. V, Bartley, J.D., Buddemeier, R.W., 2005. The role of impoundments
in the sediment budget of the conterminous United States. Geomorphology 71, 99–111.
https://doi.org/10.1016/j.geomorph.2004.01.010

Rodrigues, L.N., Sano, E.E., Steenhuis, T.S., Passo, D.P., 2012. Estimation of Small Reservoir
Storage Capacities with Remote Sensing in the Brazilian Savannah Region. Water Resour.
Manag. 26, 873–882. https://doi.org/10.1007/s11269-011-9941-8

Sanchez, A.H., Picoli, M.C.A., Camara, G., Andrade, P.R., Chaves, M.E.D., Lechler, S., Soares,
A.R., Marujo, R.F.B., Simões, R.E.O., Ferreira, K.R., Queiroz, G.R., 2020. Comparison of
Cloud Cover Detection Algorithms on Sentinel–2 Images of the Amazon Tropical Forest.
Remote Sens. 12, 1284. https://doi.org/10.3390/rs12081284

Schwatke, C., Dettmering, D., Bosch, W., Seitz, F., 2015. DAHITI – an innovative approach for

112
estimating water level time series over inland waters using multi-mission satellite altimetry.
Hydrol. Earth Syst. Sci. 19, 4345–4364. https://doi.org/10.5194/hess-19-4345-2015

Sedano, F., Kempeneers, P., Hurtt, G., 2014. A Kalman Filter-Based Method to Generate
Continuous Time Series of Medium-Resolution NDVI Images. Remote Sens. 6, 12381–
12408. https://doi.org/10.3390/rs61212381

Sentinel Hub, 2021. Sentinel Hub API Documentation. URL https://docs.sentinel-


hub.com/api/latest/ (accessed 4.20.21).

Sentinel Hub, 2020. Technical Documentation Cloud Masks and Cloud Probabilities. Sentin.
Hub Doc. URL https://docs.sentinel-hub.com/api/latest/user-guides/cloud-masks/#cloud-
masks-and-cloud-probabilities (accessed 3.10.21).

Sheng, Y., Song, C., Wang, J., Lyons, E.A., Knox, B.R., Cox, J.S., Gao, F., 2016. Representative
lake water extent mapping at continental scales using multi-temporal Landsat-8 imagery.
Remote Sens. Environ. 185, 129–141.
https://doi.org/https://doi.org/10.1016/j.rse.2015.12.041

Shults, D.D., Nowlin, W.J., Yaeger, M.A., Massey, J.H., Reba, M.L., 2020. A spatiotemporal
anlysis quantifying the need for more on-farm reservoirs to reduce groundwater use in the
Cache and L’Anguille River Regions in Northeaster AR, in: ESRI User Conference. Online.

Singh, K.V., Setia, R., Sahoo, S., Prasad, A., Pateriya, B., 2015. Evaluation of NDWI and
MNDWI for assessment of waterlogging by integrating digital elevation model and
groundwater level. Geocarto Int. 30, 650–661.
https://doi.org/10.1080/10106049.2014.965757

Solander, K.C., Reager, J.T., Famiglietti, J.S., 2016. How well will the Surface Water and Ocean
Topography (SWOT) mission observe global reservoirs? Water Resour. Res. 52, 2123–
2140. https://doi.org/10.1002/2015WR017952

Van Den Hoek, J., Getirana, A., Jung, H., Okeowo, M., Lee, H., 2019. Monitoring Reservoir
Drought Dynamics with Landsat and Radar/Lidar Altimetry Time Series in Persistently
Cloudy Eastern Brazil. Remote Sens. 11, 827. https://doi.org/10.3390/rs11070827

Vanhellemont, Q., 2019. Daily metre-scale mapping of water turbidity using CubeSat imagery.
Opt. Express 27, A1372–A1399. https://doi.org/10.1364/OE.27.0A1372

Vanthof, V., Kelly, R., 2019. Water storage estimation in ungauged small reservoirs with the
TanDEM-X DEM and multi-source satellite observations. Remote Sens. Environ. 235.
https://doi.org/10.1016/j.rse.2019.111437

Verbesselt, J., Zeileis, A., Herold, M., 2012. Near real-time disturbance detection using satellite
image time series. Remote Sens. Environ. 123, 98–108.
https://doi.org/10.1016/j.rse.2012.02.022

Vicente-Guijalba, F., Martinez-Marin, T., Lopez-Sanchez, J.M., 2014. Crop Phenology

113
Estimation Using a Multitemporal Model and a Kalman Filtering Strategy. IEEE Geosci.
Remote Sens. Lett. 11, 1081–1085. https://doi.org/10.1109/LGRS.2013.2286214

Vörösmarty, C.J., McIntyre, P.B., Gessner, M.O., Dudgeon, D., Prusevich, A., Green, P.,
Glidden, S., Bunn, S.E., Sullivan, C.A., Liermann, C.R., Davies, P.M., 2010. Global threats
to human water security and river biodiversity. Nature 467, 555–561.
https://doi.org/10.1038/nature09440

Vrugt, J.A., Gupta, H. V, Nualláin, B., Bouten, W., 2006. Real-time data assimilation for
operational ensemble streamflow forecasting. J. Hydrometeorol. 7, 548–565.
https://doi.org/10.1175/JHM504.1

Wang, Z., Zhang, R., Zhang, Q., Zhu, Y., Huang, B., Lu, Z., 2019. An Automatic Thresholding
Method For Water Body Detection From SAR Image, in: 2019 IEEE International
Conference on Signal, Information and Data Processing (ICSIDP). pp. 1–4.
https://doi.org/10.1109/ICSIDP47821.2019.9172964

Weerts, A.H., El Serafy, G.Y.H., 2006. Particle filtering and ensemble Kalman filtering for state
updating with hydrological conceptual rainfall-runoff models. Water Resour. Res. 42.
https://doi.org/10.1029/2005WR004093

Welch, G., Bishop, G., 1995. An introduction to the Kalman filter.

Yaeger, M.A., Massey, J.H., Reba, M.L., Adviento-Borbe, M.A.A., 2018. Trends in the
construction of on-farm irrigation reservoirs in response to aquifer decline in eastern
Arkansas: Implications for conjunctive water resource management. Agric. Water Manag.
208, 373–383. https://doi.org/10.1016/j.agwat.2018.06.040

Yaeger, M.A., Reba, M.L., Massey, J.H., Adviento-Borbe, M.A.A., 2017. On-farm irrigation
reservoirs in two Arkansas critical groundwater regions: A comparative inventory. Appl.
Eng. Agric. 33, 869–878. https://doi.org/10.13031/aea.12352

Yang, Xiaoying, He, R., Ye, J., Tan, M.L., Ji, X., Tan, L., Wang, G., 2020. Integrating an hourly
weather generator with an hourly rainfall SWAT model for climate change impact
assessment in the Ru River Basin, China. Atmos. Res. 244, 105062.
https://doi.org/10.1016/j.atmosres.2020.105062

Yang, Xiucheng, Qin, Q., Yésou, H., Ledauphin, T., Koehl, M., Grussenmeyer, P., Zhu, Z.,
2020. Monthly estimation of the surface water extent in France at a 10-m resolution using
Sentinel-2 data. Remote Sens. Environ. 244, 111803.
https://doi.org/10.1016/j.rse.2020.111803

Yang, X., Zhao, S., Qin, X., Zhao, N., Liang, L., 2017. Mapping of Urban Surface Water Bodies
from Sentinel-2 MSI Imagery at 10 m Resolution via NDWI-Based Image Sharpening.
Remote Sens. 9, 596. https://doi.org/10.3390/rs9060596

Yao, F., Wang, J., Wang, C., Crétaux, J.-F., 2019. Constructing long-term high-frequency time
series of global lake and reservoir areas using Landsat imagery. Remote Sens. Environ. 232,

114
111210. https://doi.org/10.1016/j.rse.2019.111210

Yao, F., Wang, J., Yang, K., Wang, C., Walter, B.A., Crétaux, J.F., 2018. Lake storage variation
on the endorheic Tibetan Plateau and its attribution to climate change since the new
millennium. Environ. Res. Lett. 13. https://doi.org/10.1088/1748-9326/aab5d3

Ye, S., Rogan, J., Zhu, Z., Eastman, J.R., 2021. A near-real-time approach for monitoring forest
disturbance using Landsat time series: stochastic continuous change detection. Remote
Sens. Environ. 252, 112167. https://doi.org/10.1016/j.rse.2020.112167

Zhang, S., Foerster, S., Medeiros, P., de Araújo, J.C., Motagh, M., Waske, B., 2016. Bathymetric
survey of water reservoirs in north-eastern Brazil based on TanDEM-X satellite data. Sci.
Total Environ. 571, 575–593. https://doi.org/10.1016/j.scitotenv.2016.07.024

Zhang, Z., Moore, J.C., 2014. Mathematical and physical fundamentals of climate change.
Elsevier.

Zhao, G., Gao, H., 2019. Estimating reservoir evaporation losses for the United States: Fusing
remote sensing and modeling approaches. Remote Sens. Environ. 226, 109–124.
https://doi.org/10.1016/j.rse.2019.03.015

Zhou, F., Zhong, D., 2020. Kalman filter method for generating time series synthetic Landsat
images and their uncertainty from Landsat and MODIS observations. Remote Sens.
Environ. 239, 111628. https://doi.org/10.1016/j.rse.2019.111628

Zhu, X., Cai, F., Tian, J., Williams, T., 2018. Spatiotemporal Fusion of Multisource Remote
Sensing Data: Literature Survey, Taxonomy, Principles, Applications, and Future
Directions. Remote Sens. 10, 527. https://doi.org/10.3390/rs10040527

Zhu, X., Chen, J., Gao, F., Chen, X., Masek, J.G., 2010. An enhanced spatial and temporal
adaptive reflectance fusion model for complex heterogeneous regions. Remote Sens.
Environ. 114, 2610–2623. https://doi.org/10.1016/j.rse.2010.05.032

Zhu, X., Helmer, E.H., Gao, F., Liu, D., Chen, J., Lefsky, M.A., 2016. A flexible spatiotemporal
method for fusing satellite images with different resolutions. Remote Sens. Environ. 172,
165–177. https://doi.org/10.1016/j.rse.2015.11.016

115
CHAPTER 4: MONITORING SMALL WATER BODIES USING HIGH SPATIAL AND

TEMPORAL RESOLUTION ANALYSIS READY DATASETS

Published in Remote Sensing December 2021


(https://doi.org/10.3390/rs13245176)

Authors
Vinicius Perin1, Samapriya Roy2, Joe Kington2, Thomas Harris2, Mirela G. Tulbure1, Noah
Stone2, Torben Barsballe2, Michele Reba3 and Mary A. Yaeger4

Affiliations
1
Center for Geospatial Analytics, North Carolina State University, 2800 Faucette Drive, Raleigh,
NC, 27606, USA
2
Planet Labs Inc., San Francisco, CA 94107, USA
3
USDA-Agricultural Research Service, Delta Water Management Research Unit, Jonesboro, AR,
72467, USA
4
Center for Applied Earth Science and Engineering Research, The University of Memphis, 3675
Alumni Drive, Memphis, TN, 38152, USA

Acknowledgements
The first author is supported by NASA through the Future Investigators in NASA Earth

and Space Science and Technology fellowship. The first author acknowledges the support

provided by Planet Labs throughout his internship at the company, which included access to

SkySat and PlanetScope imagery and on-demand generation of Basemap and Planet Fusion

specifically for this study.

Abstract
Basemap and Planet Fusion—derived from PlanetScope imagery—represent the next generation

of analysis ready datasets that minimize the effects of the presence of clouds. These datasets have

116
high spatial (3 m) and temporal (daily) resolution, which provides an unprecedented opportunity

to improve the monitoring of on-farm reservoirs (OFRs)—small water bodies that store freshwater

and play important role in surface hydrology and global irrigation activities. In this study, we

assessed the usefulness of both datasets to monitor sub-weekly surface area changes of 340 OFRs

in eastern Arkansas, USA, and we evaluated the datasets main differences when used to monitor

OFRs. When comparing the OFRs surface area derived from Basemap and Planet Fusion to an

independent validation dataset, both datasets had high agreement (r2 ≥ 0.87), and small

uncertainties, with a mean absolute percent error (MAPE) between 7.05% and 10.08%. Pairwise

surface area comparisons between the two datasets and the PlanetScope imagery showed that 61%

of the OFRs had r2 ≥ 0.55, and 70% of the OFRs had MAPE <5%. In general, both datasets can

be employed to monitor OFRs sub-weekly surface area changes, and Basemap had higher surface

area variability and was more susceptible to the presence of cloud shadows and haze when

compared to Planet Fusion, which had a smoother time series with less variability and fewer abrupt

changes throughout the year. The uncertainties in surface area classification decreased as the OFRs

increased in size. In addition, the surface area time series can have high variability, depending on

the OFR environmental conditions (e.g., presence of vegetation inside the OFR). Our findings

suggest that both datasets can be used to monitor OFRs sub-weekly, seasonal, and inter-annual

surface area changes; therefore, these datasets can help improve freshwater management by

allowing better assessment and management of the OFRs.

Keywords: Analysis ready datasets, PlanetScope, Basemap, Planet Fusion, on-farm reservoirs,

water management

117
1. INTRODUCTION

Planet Labs currently operates more than 200 PlanetScope satellites in sun-synchronous

orbits and frequently launches new satellites that are designed to have a short operational lifetime

(<4 years). The PlanetScope satellite constellation enables near-daily monitoring with multi-

spectral imagery at high spatial resolution (3 m) (Planet Labs Team, 2021a). PlanetScope imagery

has been applied to a variety of studies to monitor phenomena that require both high spatial and

temporal resolution, for instance, to monitor small water bodies (Cooley et al., 2017; Mishra et al.,

2020; Vanthof and Kelly, 2019), estimate methane emissions from forested wetlands (Hondula et

al., 2021), assess river-ice and water velocity (Kääb et al., 2019), improve crop leaf-area-index

estimation with sensor data fusion (Houborg and McCabe, 2018b; Kimm et al., 2020; Sadeh et al.,

2021), and monitor near-real-time aboveground carbon emissions from tropical forests (Csillik et

al., 2020, 2019; Csillik and Asner, 2020).

A recent global analysis of PlanetScope’s temporal availability (Roy et al., 2021) showed

that the annual and monthly number of PlanetScope observations does not vary uniformly across

the globe. The authors attributed this finding to different PlanetScope orbits (i.e., altitude and

inclinations), due to different numbers of sensors in orbit, which vary when PlanetScope satellites

are decommissioned and replaced with new sensors, and due to images that cannot be geolocated

(Roy et al., 2021). In addition, it is well known that the number of observations from optical

wavelength satellite imagery will vary according to dynamic and global cloud obscuration. While

the PlanetScope cloud mask, Usable Data Mask 2 (UDM2), is available (Planet Labs Team, 2021a)

and allows for discernment of classes like cloud, cloud shadow, and heavy, haze among others, its

accuracy has not been thoroughly assessed (Cheng et al., 2020; Roy et al., 2021; Wang et al., 2021)

and it is not available for images prior to 2018 (Roy et al., 2021). Aiming to overcome these

118
limitations—irregular cadence and cloud obscuration—and to increase the applications of

PlanetScope imagery, Planet Labs has focused on developing the next generation of tiled analysis

ready datasets—Basemap (Planet Labs Team, 2019) and Planet Fusion (Planet Labs Team,

2021b)—which are less affected by the presence of clouds and are set for a fixed temporal cadence.

Basemap is generated by mosaicking the whole or part of the highest quality PlanetScope

imagery, which is selected based on cloud cover and image acutance (i.e., sharpness). For example,

for a given period of interest—Basemap can be processed using different image cadences, e.g.,

daily, weekly, biweekly—PlanetScope images are ranked based on these metrics such that cloud-

free images have higher scores than cloudy images (Kington et al., 2019; Planet Labs Team, 2019).

Basemap is designed to monitor changes over time and for analytics-driven use cases, and it has

been applied to several research projects, including monitoring of forest biomass (Csillik et al.,

2020, 2019; Csillik and Asner, 2020), to assess carbon emissions from drainage canals (Dadap et

al., 2021), and to monitor coral reef map probabilities (Knapp et al., 2018). Planet Fusion, on the

other hand, is based on the CubeSat-enabled spatiotemporal enhancement method (Houborg and

McCabe, 2018b), and it leverages the high spatial and temporal resolution provided by

PlanetScope scenes with rigorously calibrated publicly available multispectral satellites (i.e.,

Sentinel-2, Landsat, MODIS, and VIIRS) to provide daily and radiometrically consistent and gap-

filled surface-reflectance images that are free of clouds and shadows (Planet Labs Team, 2021b).

Planet Fusion is suitable to assess inter-day changes, for time series analysis, and monitoring of

disturbances of Earth’s surface. Recently, Planet Fusion has been applied to monitor crop

phenology, using the normalized difference vegetation index and leaf area index (Houborg and

McCabe, 2018a; Kong et al., 2021). Given that these datasets are cloud-free and processed to have

daily cadence at high spatial resolution—both Basemap and Planet Fusion have 3 m pixel size—

119
they provide an unprecedented opportunity to improve the monitoring of dynamic small water

bodies, for instance, on-farm reservoirs (OFRs) that are used by farmers to store water during the

wet season and for crop irrigation during the dry season. OFRs have a dynamic surface area time

series, especially during the crop-growing season, when farmers are irrigating their crops and may

pump water from nearby streams (Fowler et al., 2015; Habets et al., 2018; Perin et al., 2021).

There are more than 2.6 million OFRs in the USA alone, and these OFRs play a key role

in surface hydrology by storing fresh water and as an essential component of global irrigation

activities (Downing, 2010; Downing et al., 2006; Renwick et al., 2005). Nonetheless, OFRs can

contribute to downstream water stress by decreasing stream discharge and peak flow in the

watersheds where they are built (Habets et al., 2018, 2014; Mime and Young, 1989). Therefore,

monitoring OFRs sub-weekly surface area changes is critical to the assessment of their seasonal

and inter-annual variability, as well as to mitigation of their downstream impacts, with implications

concerning how OFRs are managed and where they are built. Previous research assessed the spatial

and temporal variability of OFRs by leveraging the long-term (≥25 years) Landsat-based

inundation datasets (Jones et al., 2017; Ogilvie et al., 2018b; Perin et al., 2021). Nonetheless, these

datasets are limited to a few annual observations—due to clouds, sensor issues, and the 16-day

repeat cycle—and Landsat’s spatial resolution (30m) limits the applications of these datasets to

monitor OFRs smaller than 5 ha (i.e., high surface area uncertainties ~ 20%). Aiming to overcome

these limitations, other studies (Ogilvie et al., 2018a; Perin et al., 2022; Vanthof and Kelly, 2019)

have applied a multi-sensor satellite imagery approach, including sensors of higher spatial and

temporal resolution (e.g., PlanetScope [3 m] and Sentinel-2 [10 m]) when compared to Landsat.

However, a multi-sensor approach requires processing of satellite imagery of different spatial

resolution from multiple platforms, which can be time-consuming and a limiting factor if it is

120
necessary to process, download, and move the satellite imagery across multiple platforms (Perin

et al., 2022). In this study, we propose a novel use of the analysis ready datasets Basemap and

Planet Fusion, and we aim (1) to assess the usefulness of both datasets to monitor OFRs sub-

weekly surface area changes and (2) to compare the two datasets and describe their differences

when used to monitor OFRs.

2. METHODS

2.1 Study region


Eastern Arkansas is one of the largest irrigated regions in the USA that has seen a rapid

increase in the number of OFRs during the last 40 years (Shults et al., 2020; Yaeger et al., 2018,

2017). The region has a humid subtropical climate with an average annual precipitation of 1300

mm, mostly distributed between March and May and November and January (Perin et al., 2021).

Recent studies (Yaeger et al., 2018, 2017) mapped the spatial distribution of 340 OFRs with

surface area <30 ha and distributed across three sub-watersheds in the study region (Fig. 1). The

OFR dataset was manually mapped using the high-resolution (1 m) National Agriculture Imagery

Program archive in combination with 2015 Google Earth satellite imagery. The authors of the OFR

dataset used Google Earth Explorer to sharpen the image details when zooming in and to provide

a validation for features appearing indistinct or pixelated in the 1-m mosaic imagery (Yaeger et

al., 2017). We assigned the OFRs to three size classes (0.1–5 ha, 5–10 ha, and 10–30 ha) based on

the surface area mapped in the OFR dataset. These classes were used to support the surface area

monitoring analyses when accounting for different OFR sizes (Fig. 1).

121
Figure 1–Study region in eastern Arkansas, USA, and the OFRs size distribution. The inset map

represents the OFRs shapefile overlaid on SkySat satellite imagery.

We downloaded PlanetScope images and processed daily Basemap and Planet Fusion

images between July 2020 and July 2021. This time frame was chosen based on the imagery

availability to generate both analysis ready datasets. The images spatial resolution and band-

wavelength ranges are presented in Table 1. In addition, the general workflow used to assess the

OFRs’ surface area time series is provided in Fig. 2.

Table 1–PlanetScope, Basemap, and Planet Fusion images spatial resolutions and different

wavelengths bands.

Source Pixel size (m) Blue (µm) Green (µm) Red (µm) NIR (µm)
PlanetScope 3 0.455–0.515 0.500–0.590 0.590–0.670 0.780–0.860
Basemap 3 0.450–0.510 0.530–0.590 0.640–0.670 0.850–0.860
Planet Fusion 3 0.450–0.510 0.530–0.590 0.640–0.670 0.850–0.880

122
Figure 2–Workflow used to estimate the OFRs’ surface area-time series from PlanetScope,

Basemap, and Planet Fusion between July 2020 and July 2021.

2.2. Satellite Imagery Datasets


2.2.1. PlanetScope CubeSat Surface-Reflectance Ortho Tiles
We used the OFRs’ shapefile to search for and clip Level 3A surface-reflectance imagery

available through Planet Orders API. The PlanetScope surface-reflectance ortho tiles use a fixed

UTM grid system in 25 km by 25 km tiles with 1 km overlap (Planet Labs Team, 2021a). We

filtered out all images with more than 10% cloud using an image-based cloud-cover filter—this

cloud-cover filter threshold allowed us to download mostly cloud-free images; however, because

it is an image-based filter rather than an OFR or area-of-interest-based cloud filter, some useful

observations (i.e., when the OFR is not covered with clouds but the image is filtered out) were not

downloaded, decreasing the total number of observations per OFR. In addition, to deal with

123
potential cloud-obscuration outliers, we used the PlanetScope UDM2 to filter out all image clips

that contained more than 5% unusable pixels (i.e., pixels covered by clouds, cloud shadow, with

light and heavy haze).

The PlanetScope ortho tiles were resampled to 3 m and projected using the WGS84 datum.

The ortho tiles were radiometrically, sensor, and geometrically corrected and aligned to a

cartographic map projection. These images were atmospherically corrected using the 6S radiative

transfer model with ancillary data from MODIS (Kotchenova et al., 2006; Kotchenova and

Vermote, 2007; Planet Labs Team, 2021a), and the positional accuracy has been reported to be

smaller than 10 m (Planet Labs Team, 2021a).

2.2.2. Normalized Surface-Reflectance Basemap


We processed daily Basemap images corrected to surface reflectance using PlanetScope

scenes, and a “best scene on top” algorithm (Kington et al., 2019; Planet Labs Team, 2019) that

selects the highest quality imagery from the PlanetScope catalog. This algorithm ranks the

PlanetScope scenes based on their quality by assessing the scenes’ acutance (i.e., sharpness), the

fraction of cloud cover, cloud shadow, haze, and presence of unusable pixels (e.g., no data).

Briefly, this algorithm is based on a linear regression model approach that uses the clear pixels

from the best-ranked scenes; we selected the best scenes first, then progressed successively until

the images were filled or no scenes remained (Kington et al., 2019). To obtain Basemap at a daily

cadence, we employed a 30-day rolling window that may use PlanetScope scenes collected up to

15 days before or after the target date; however, if no usable pixels (i.e., cloud-free) are available

in this time range, the image will contain no data. We did not observe any Basemap image with

no-data in our study period. The rolling window approach weights on the image recency, for

instance, a slightly hazy scene (e.g., ~<10% hazy pixels) on the day of the Basemap image, will

124
score higher than a very clear scene (i.e., no haze) from a few days before/after. In addition, due

to the daily cadence, there may be Basemap images with the same PlanetScope scene composition,

which leads to repetitive information when using the Basemap images to monitor OFRs.

Basemap images were generated employing a two-step process: normalization and

seamline removal. Normalization aims to radiometrically calibrate the Basemap images and to

minimize the scene-to-scene variability when mosaicking PlanetScope scenes. For this step, the

Framework for Operational Radiometric Correction for Environmental Monitoring (FORCE)

(Frantz, 2019) was used to generate a combined Landsat 8 and Sentinel-2 surface-reflectance

product to be used as the “gold” radiometric reference during normalization. FORCE infers surface

reflectance from Landsat 8 and Sentinel-2 by employing the 5S (simulation of the satellite signal

in the solar spectrum) approach (Tanre et al., 1990). The aerosol optical depth is estimated using

a dark-object-based approach where water vapor content is derived from Landsat 8 (obtained from

MODIS database) and Sentinel-2 (estimated on a pixel-specific basis) imagery. In addition, clouds

and shadows are detected using a modified version of Fmask (Zhu and Woodcock, 2012) for

Sentinel-2 images (Frantz et al., 2018)—see Planet Labs Team (2021b, 2019) for further details.

An assessment of the FORCE atmospheric correction was performed as part of the atmospheric

correction inter-comparison exercise (Doxani et al., 2018), and the FORCE implementation uses

the Landsat 8 and Sentinel-2 imagery mapped onto a common UTM grid to produce 30 m spatial-

resolution imagery. Seamline removal enhances the visual appearance of the Basemap image

edges. In this step, each PlanetScope scene used in the Basemap mosaic is set to match its

neighbor—pixel values near a scene boundary change more than values away from the boundary;

however, the pixel values are not modified. Specifically, we first calculated the Basemap mosaic

pixel values gradient, then set the gradient values between 1 and 0 (scene boundary) and fixed the

125
original pixel values along the Basemap mosaic edge. This process was applied independently for

each band; therefore, it may alter band ratios near scene edges—this is most apparent when scenes

do not match locally, for instance, for unmasked clouds. Lastly, the seamline removal may

introduce artifacts (e.g., straight lines, distortions) at the Basemap mosaic boundary, which is most

frequent over water when normalization cannot fully correct for differences between scenes due

to waves and sun glint.

2.2.3. Planet Fusion Surface Reflectance


We processed Planet Fusion images using an algorithm based on the CubeSat-enabled

spatiotemporal enhancement method (Houborg and McCabe, 2018b), which enhances, inter-

calibrates, and fuses satellite imagery from multiple sensors. Planet Fusion has unique features,

including (1) precise co-registration and sub-pixel fine alignment for different image sources, (2)

PlanetScope scenes with near-nadir field of view, resulting in minimal bidirectional reflectance

distribution function (BRDF) variation effects, and (3) pixel traceability to identify imagery

sources and to assess the confidence of daily gap-filled images.

To generate Planet Fusion surface-reflectance images, we used the same approach

described for Basemap (i.e., FORCE (Frantz, 2019)), with top-of-atmosphere (TOA) PlanetScope

scenes (3 m), Sentinel-2 TOA reflectance (10–20 m), Landsat 8 TOA reflectance (30 m), and daily

tile-based MODIS or VIIRS normalized to a nadir-view direction and local-solar-noon surface

reflectance. The Planet Fusion algorithm uses MODIS MCD43A4 surface-reflectance product in

seven spectral bands that are corrected for reflectance anisotropy using a semi-empirical BRDF

(Schaaf et al., 2002), which utilizes the best observations collected over a 16-day period centered

on the day of interest. In addition, VIIRS products (VNP43IA4 and VNP43MA4) are used as a

backup to ensure continuity if MODIS data is not available.

126
The Planet Fusion algorithm guarantees spatially complete and temporally continuous

images by gap-filling radiometric data (i.e., synthetic pixel values). The gap-filling process uses

both spatial (i.e., neighboring and class-specific pixel information) and temporal interpolation

techniques to estimate the pixel values. In general, uncertainty will vary based on Earth’s surface

characteristics (e.g., vegetation dynamic changes), and it will be higher for longer daily interval

gaps. Planet Fusion images are accompanied by a quality-assurance product, which is a thematic

raster layer using the same spatial grid (i.e., UTM grid system in 24 km by 24 km tiles) as the

corresponding Planet Fusion spectral data (Planet Labs Team, 2021b). We used the quality-

assurance product to assess the percentage of synthetic (i.e., gap-filled) versus observation data

(PlanetScope and Sentinel-2) used to generate the pixel value. The observation data can be a

combination of PlanetScope and Sentinel-2. A value of 1 indicates no gap-filling, whereas a value

of 100 indicates an entirely gap-filled pixel value. Specifically for our study case, when clipping

Planet Fusion images using the OFR boundaries, the clips can have real pixels, synthetic pixels,

or a combination of both. Additionally, there are known issues associated with the gap-filling

process used by the Planet Fusion algorithm, including false cloud or cloud-shadow detection and

image artifacts (e.g., strips, distortions). These issues are most common during prolonged

cloudiness and in study regions with significant terrain shadowing.

2.3. Data Analysis


To classify the OFR surface area from PlanetScope, Basemap, and Planet Fusion, we

clipped all available images using each OFR shapefile buffered to 100 m. Then, we calculated the

normalized difference water index (NDWI) using the green and NIR bands (McFeeters, 1996), and

we applied an adaptive Otsu threshold (Otsu, 1979) for each image in the time series to separate

water from non-water pixels. The Otsu threshold is a well-known algorithm used to classify surface

127
water of inland water bodies (Du et al., 2014; Li and Wang, 2015; Liu et al., 2016; Sheng et al.,

2016; Vanthof and Kelly, 2019; Wang et al., 2019). In addition, the Otsu threshold optimizes the

separability of pixel values is contingent on the bimodal distribution of the pixel values (i.e., water

and non-water pixels), which was ensured by clipping the satellite imagery using each OFR

shapefile. After calculating the Otsu threshold and separating water from non-water pixels, we

clipped the images one more time using the OFRs’ shapefiles buffered at 20 m. This last step was

done to minimize the impact (i.e., inflating surface area) of adjacent water bodies when estimating

OFR surface area. All surface area image classification was done in Google Earth Engine (Gorelick

et al., 2017).

PlanetScope has a near-daily revisit time; however, the number of usable satellite images

varies throughout the year due to the presence of clouds and sensor-related issues. To assess the

number of different PlanetScope observations, we first counted the total number of observations

for each OFR and for each month; then, we plotted the monthly distribution of this number,

including all OFRs (i.e., one boxplot for each month of the year that represents the variability in

the number of monthly observations according to different OFRs). In addition, we evaluated the

total number of different observations for each OFR along the year (i.e., a histogram that represents

the distribution of the total number of observations for each OFR). A similar approach was used

to assess the number of different Basemap observations and to count the number of Planet Fusion

observations that were real, mixed (i.e., including real and synthetic pixels), and synthetic.

Basemap images with the same PlanetScope scene composition were counted only once.

For the different OFR surface area size classes (Fig. 1), we assessed the uncertainties in

the Basemap and Planet Fusion images by pairwise comparing them with PlanetScope and

calculating the percent error (Equation (1)) monthly distribution and the monthly mean absolute

128
percent error (MAPE; Equation (2)). In addition, for Planet Fusion, we divided the pairwise

comparisons between real, mixed and synthetic surface area observations. We illustrated the

surface area time series derived from PlanetScope, Basemap, and Planet Fusion for six OFRs of

different sizes (Table 2). These OFRs were chosen to demonstrate the surface area time series

variability from the different images and for OFRs located under different environmental

conditions (e.g., presence of vegetation inside the OFR, close to adjacent water bodies, a multi-

part OFR). In addition, we overlaid the OFRs’ shapefile on high-resolution Google Maps satellite

imagery to show the environmental conditions where the OFRs are located.
𝑦 𝑖 − 𝑥𝑖
Percent error (%) = ( ) * 100 (2)
𝑥𝑖

1 𝑦 𝑖 − 𝑥𝑖
Mean absolute percent error (%) = 𝑛 𝛴| | ∗ 100 (3)
𝑥𝑖

Where 𝑥i is the SkySat or PlanetScope surface area, and yi is the Basemap or Planet Fusion surface

area.

Table 2–Selected OFRs (see Fig. 8) to illustrate PlanetScope, Basemap, and Planet Fusion surface

area time series, and their size according to the OFRs dataset.

OFR id OFR size (ha)


A 13.62
B 22.60
C 9.83
D 12.73
E 13.82
F 29.72

2.4. Validation Scheme


To validate the surface area classification using the Otsu thresholding approach, we

downloaded five orthorectified and multispectral SkySat images (Planet Labs Team, 2021b) (Blue:

0.450–0.515 µm, Green: 0.515–0.595 µm, Red: 0.605–0.695 µm, and NIR: 0.740–0.900 µm) at

129
sub-meter (0.66–0.73 m) spatial resolution (Table 3). For each image, we overlaid the OFR

geometry and manually delineated the OFR surface area, which resulted in 144 validation surface

areas from 71 different OFRs for multiple observations in time. Then, we conducted a pairwise

comparison of the validation surface area with the surface area obtained from PlanetScope,

Basemap, and Planet Fusion. When the PlanetScope surface area date did not correspond exactly

to the SkySat dates, we used the closest observation in time, which had a maximum difference of

three days before or after the SkySat date. In addition, we assessed the uncertainties of

PlanetScope, Basemap, and Planet Fusion for different surface area size classes: 0.1–5 ha (n = 50),

5–10 ha (n = 46), and 10–50 ha (n = 48).

Table 3–SkySat image identification, acquisition date, number of OFRs surface area observations

per image, percent clear (indicates the presence or absence of cloud cover; higher values indicate

fewer clouds), ground-control ratio (defines the image positional accuracy; values closer to 1 mean

higher accuracy), and ground-sampling distance in meters.

Ground
OFRs Percent clear Ground control sampling
SkySat image Date observations (%) ratio distance (m)
20200929_193409
_ssc10_u0001 2020-09-29 35 87 0.91 0.66
20201013_194518
_ssc6_u0001 2020-10-13 44 99 0.97 0.73
20201102_193752
_ssc9_u0001 2020-11-02 16 100 0.97 0.67
20201102_193752
_ssc9_u0002 2020-11-02 10 100 0.96 0.67
20201210_194154
_ssc11_u0001 2020-12-10 39 99 0.91 0.68

130
3. RESULTS

3.1. Surface Water Area Validation Using SkySat Imagery


The surface area obtained from PlanetScope, Basemap, and Planet Fusion showed great

agreement (r2 ≥ 0.98) with the validation dataset. In addition, PlanetScope had the smallest MAPE

(8.09%), followed by Basemap (8.21%) and Planet Fusion (9.17%) (Fig. 3). When splitting the

validation surface area observations into different size classes (Table 4), all three image sources

presented similar agreement (r2 ≥ 0.87), and the highest r2 values were found for surface area

observations between 10 and 30 ha (r2 ≥ 0.95). All three sources had a similar MAPE for

observations between 0.1 and 5 ha (~7.55%) and between 10 and 30 ha (~7.98%), while the highest

values were found for observations between 5 and 10 ha (~10.27%).

Figure 3–Pairwise comparisons between the SkySat validation surface area and the surface area

obtained from PlanetScope, Basemap, and Planet Fusion for multiple observations in time

131
Table 4–Pairwise comparisons between the SkySat validation surface area and the surface area

obtained from PlanetScope, Basemap, and Planet Fusion for multiple observations in time and

divided into three size classes.

PlanetScope Basemap Planet Fusion


0.1–5 ha 5–10 ha 10–30 ha 0.1–5 ha 5–10 ha 10–30 ha 0.1–5 ha 5–10 ha 10–30 ha
r2 0.91 0.91 0.94 0.91 0.87 0.90 0.96 0.96 0.95
MAPE
(%) 7.05 7.14 7.68 9.91 10.1 10.8 7.37 7.46 9.11

3.2. Number of Surface Area Observations per Dataset


The number of PlanetScope observations varied throughout the year and varied across

different OFRs (Fig. 4). The months with the highest number of PlanetScope images were

November–December 2020 (~17) and March–April 2021 (~14), while the months with the lowest

numbers were July–September 2020 (< 10) and February 2021 (< 3) (Fig. 4A). In addition, most

of the OFRs (~60%) had 80–100 PlanetScope observations per year (Fig. 4B). Basemap images

were processed at a daily cadence, and we considered a new Basemap observation every time a

new image composite was used. In this regard, the number of Basemap images followed a similar

pattern found for PlanetScope; however, the mean number of Basemap observations per month

was higher than of PlanetScope in 10 out of the 12 months analyzed, and most of the OFRs (~75%)

had 90–120 Basemap observations per year (Fig. 4A,B).

132
Figure 4–Number of surface area observations per month for PlanetScope and Basemap (A), and

for Planet Fusion real, mixed, and synthetic (C). The frequency distribution of the total number of

observations per OFR per year for PlanetScope and Basemap (B), and for Planet Fusion real,

mixed, and synthetic (D).

Planet Fusion images were derived from real and synthetic pixel values, and the number of

real and synthetic observations varied throughout the year (Fig. 4C). The number of images

derived from real pixels reached its peak (~13–15) between November and December 2020, and

the lowest numbers were found in February 2021 (~2) and between May and June 2021 (< 5). In

general, most of the OFRs had ~ 80–100 real observations per year. The number of mixed images

(i.e., composed by real and synthetic pixels) tended to be < 10 for all months, and most of the

OFRs had < 50 mixed observations per year (Fig. 4C,D). The number of synthetic images is higher

than real and mixed observations for all months of the year, and the highest values (~22–26)

occurred in July 2021 and May–June 2020, with the lowest values between November and

133
December 2020 (~14–16) (Fig. 4C). In addition, most of the OFRs had ~250–260 synthetic

observations per year (Fig. 4D).

3.3. Planet Fusion Comparison with PlanetScope


We found a high agreement (r2 ≥ 0.90) for the same-day surface area pairwise comparisons

between Planet Fusion and PlanetScope for all size classes (Fig. 5). MAPE decreased as

observations increased in size, and the highest MAPE values were found for synthetic, mixed, and

real for all size classes (Fig. 5). The number of pairwise comparisons for real was higher than

mixed and synthetic, to a large extent (~60%). This finding is somewhat expected, as the Planet

Fusion algorithm uses PlanetScope images as an input to generate daily Planet Fusion imagery.

134
Figure 5–Same day pairwise comparisons between PlanetScope and Planet Fusion real, mixed,

and synthetic, for multiple observations in time and for all OFRs divided into three size classes

(0.1–5 ha, 5–10 ha, and 10–30 ha). Brighter colors indicate higher point density.

3.4. Monthly Comparisons between Basemap and Planet Fusion with PlanetScope
When comparing each OFR surface area time series derived from Basemap and Planet

Fusion with PlanetScope, for both datasets, most of the OFRs (63% and 61% for Basemap and

135
Planet Fusion, respectively) showed good agreement with r2 ≥ 0.55, and 74% and 70% of the

OFRs presented small uncertainties with MAPE <5% (Fig. 6).

Figure 6–Frequency distribution of r2 and MAPE calculated by comparing the OFRs time series

from Basemap and Planet Fusion with PlanetScope.

The mean monthly percent error—calculated by comparing Basemap and Planet Fusion

with PlanetScope—for Basemap and Planet Fusion varied between −2.45–1.48% and between

−3.36%–1.66% for 0.1–5 ha, between −2.88–1.11% and between −3.56–0.51% for 5–10 ha, and

between −2.23–0.53% and −3.13–0.76% for 10–30 ha. These values were stable throughout the

year (Fig. 7). The percent error variability decreased as the surface area observations increased in

size, and the observations between 10 and 30 ha had the least variability. In addition, Planet Fusion

presented smaller percent error variability when compared to Basemap for all size classes (Fig. 7).

The highest MAPE values for Basemap (4.73%) and Planet Fusion (5.80%) were found for

observations between 0.1 and 5 ha, and the MAPE was <4.40% for all months for both Basemap

136
and Planet Fusion for observations between 5 and 10 ha and 10 and 30 ha, respectively. This

indicates that even when there are fewer PlanetScope images available to generate Basemap and

Planet Fusion due to clouds, shadow, and haze, both products tend to have surface area

uncertainties <5%.

Figure 7–Monthly percent error variability and MAPE calculated from the same day pairwise

comparisons between Basemap and Planet Fusion with PlanetScope for the three size classes (0.1–

5 ha, 5–10 ha, and 10–30 ha).

3.5. OFR Surface Area Time Series


We selected six OFRs (Table 2) to illustrate the surface area time series derived from

PlanetScope, Basemap, and Planet Fusion (Fig. 8). The surface area time series show that different

OFRs have different surface area change patterns. In general, the OFR surface area decreased

between 20 July and 20 November (e.g., Fig. 8, OFRs A–D), period of the year when farmers are

137
irrigating their crops, and it increased between 21 January and 21 May, which are the months when

the study region receives most of its annual precipitation (Perin et al., 2021).

When compared to PlanetScope and Basemap, Planet Fusion had a smoother surface area

time series with less variability (e.g. Fig. 8 OFRs A–D). In addition, the Planet Fusion time series

was less affected by the presence of clouds and haze that can increase or decrease the OFRs' surface

area. Even though we used a small cloud cover threshold (< 10%) for PlanetScope, there are several

PlanetScope and Basemap images contaminated with cloud shadows and haze (e.g. Fig. 8 OFR A

between 08/20–09/20) indicating surface area ~ 20% larger than Planet Fusion. Other examples

were observed between 07/20–08/20 and 05/21–06/21 in Fig. 8 OFR B, in which there were no

PlanetScope images available, and the Basemap shows abrupt changes in surface area—a drop of

20% and 15% for both dates—that are caused by the presence of cloud shadows and haze. In Fig.

9, we highlighted the impact of clouds and haze for OFR A (08/16/2020) and OFR B (08/30/2020).

For OFR A, PlanetScope and Basemap surface areas were ~ 20% larger than Planet Fusion which

is explained by the misclassification of water on the lower-right corner of the OFR. For OFR B,

while the PlanetScope image had surface area ~ 13% larger than Planet Fusion, the Basemap image

indicated surface area ~ 14% smaller, these discrepancies are caused by the presence of clouds on

the PlanetScope image and haze on the Basemap image.

The OFRs' surface water classification is impacted by the OFRs environmental conditions

and their shape geometry—OFRs with complex geometries (e.g., not circular or square and shapes

with a high number of edges) tend to have higher surface area classification uncertainties (Perin et

al., 2022). For example, Fig. 8 OFR D shows a multi-part OFR that may not have all parts

inundated at the same time, which can explain part of the variability in the surface area time series

for PlanetScope, Basemap, and Planet Fusion. The surface area time series from OFR E and OFR

138
F (Fig. 8) are influenced by the presence of vegetation inside the OFRs—the presence of vegetation

impacts the surface water classification (DeVries et al., 2017; Hondula et al., 2021) leading to

noisy surface area time series and abrupt changes (e.g. OFR E between 09/20–01/21). In addition,

the high-variability in surface area for OFRs E and F is related to the presence of adjacent water

bodies, which can inundate during flood events and contribute to change the OFRs' boundary

limits. We highlighted the impact of vegetation on the OFR E time series for two different

occasions 07/14/2020 and 10/16/2020 (Fig. 10). During the first occasion, PlanetScope and

Basemap indicated surface area (~ 9.5 ha) 95% higher than Planet Fusion (0.5 ha); on the second

occasion, a contrasting scenario, in which Planet Fusion (12.25 ha) was 86% higher than

PlanetScope and Basemap (~ 2 ha). These results shed light on the importance of assessing the

OFRs environmental conditions and how they are impacting the OFRs' surface area time series

before employing these datasets to monitor the OFRs' surface area changes.

139
Figure 8–OFRs (see Table 2) surface area time series derived from PlanetScope, Basemap and

Planet Fusion, and the OFRs shapefile overlaid on high-resolution Google Satellite imagery.

140
141
142
Figure 9–OFRs A and B (see Table 2) PlanetScope, Basemap, and Planet Fusion false color

composites (Blue: Red, Green: Green, and Red: NIR), and the surface water classification for

08/16/2020 (OFR A) and 08/30/2020 (OFR B).

Figure 10–OFR E (see Table 2) PlanetScope, Basemap, and Planet Fusion false color composites

(Blue: Red, Green: Green, and Red: NIR), and the surface water classification for 10/16/2020 and

07/14/2020.

4. DISCUSSION

The surface area validation carried out using multiple SkySat imagery showed that the

methodology used to classify OFR surface area performed well for PlanetScope, Basemap, and

Planet Fusion, with high agreement r2 ≥ 0.87 and MAPE between 7.05% and 10.08% for all image

143
sources and all size classes (Table 4). Comparisons between Basemap and Planet Fusion with

PlanetScope highlighted that most of the OFRs had good agreement with 61% of the OFRs with

r2 ≥ 0.55, and small uncertainties with 70% of the OFRs with MAPE < 5% (Fig. 6). Basemap and

Planet Fusion presented similar monthly mean percent error (~ −3–3%) and MAPE (~2.20–5.80%)

throughout the year (Fig. 7). In addition, percent error variability and MAPE decreased for the

larger surface area observations (Fig. 7). The highest monthly MAPE (5.80%) was found for Planet

Fusion for observations between 0.1 and 5 ha, and the MAPE was ≤4.40% for Basemap and Planet

Fusion for observations between 5 and 10 ha and between 10 and 30 ha. Furthermore, when

analyzing the three Planet Fusion data categories (i.e., real, mixed, and synthetic), the greatest

uncertainties were found for the synthetic images (MAPE ~ 5%), followed by mixed (MAPE ~

4%) and real (MAPE ~ 3%) (Fig. 5). These findings indicate that Basemap and Planet Fusion

images can be employed to monitor OFRs with uncertainties < 10% when the sources are

compared to the validation dataset and with uncertainties < 5% when compared to PlanetScope.

However, time series obtained from Basemap and Planet Fusion can be highly variable (Fig. 8E,F),

as surface water classification can be impacted by the size of water bodies (Table 4, Fig. 2, Fig.

4 and Fig. 6) and the environment in which OFRs are located (e.g., presence of vegetation within

the OFRs; Fig. 8, OFRs D–F).

The number of cloud-free observations offered by Basemap and Planet Fusion enlightens

the potential of these datasets to monitor OFR surface area changes (Fig. 4). Both datasets pose

advantages when compared to a single sensor approach—employing PlanetScope alone (Fig. 4),

or other sensors, for example, Landsat (Avisse et al., 2017; Jones et al., 2017; Perin et al., 2022),

Sentinel 1 (López-Caloca et al., 2020), and Sentinel-2 (Pena-Regueiro et al., 2020; Yang et al.,

2020, 2017)—or a multi-sensor approach (Ogilvie et al., 2018a; Perin et al., 2022; Vanthof and

144
Kelly, 2019). Briefly, the use of a single sensor is limited to a few observations a month, and in

some periods of the year in eastern Arkansas, there could be weeks without a cloud-free image

(Perin et al., 2022). Although the number of observations is improved when employing a multi-

sensor approach, daily to sub-weekly monitoring is not attainable unless an assimilation algorithm

(Perin et al., 2022) is implemented. In addition, when implementing a multi-sensor approach, it is

necessary to acquire the data from different platforms (e.g., Planet Explore, Sentinel Hub, and

Google Earth Engine), which can be time-consuming and a limiting factor if it is necessary to

process, download, and move the satellite imagery across multiple platforms. In this study, we

demonstrated that Basemap and Planet Fusion imagery processing can be done entirely in the cloud

environment by leveraging the integration of Planet’s Platform, Google Cloud Storage, and Google

Earth Engine. This integration allows for swift analysis, and it can be used for other study regions

without the need to acquire data from multiple platforms.

Daily OFR surface area time series derived from Basemap and Planet Fusion revealed

important differences between the two datasets. In general, Basemap had higher surface area

variability, and it was more susceptible to the presence of cloud shadows and haze when compared

to Planet Fusion, which had a smoother time series with less variability and fewer abrupt changes

throughout the year (Fig. 8). The Planet Fusion algorithm combines data from multiple satellites

to establish a baseline of OFR surface area time series by filling gaps with synthetic pixels.

Nonetheless, the smoothing effect should be interpreted cautiously, as some changes in the time

series due to large rainfall events or frequent irrigation activities may be smoothed out. This is

especially relevant for the periods of the year when there are more synthetic observations (Fig. 4)

and the uncertainties in surface area tend to be higher (Fig. 5). Additionally, because Planet Fusion

is based on a robust algorithm that uses data from various satellites, this dataset requires more

145
image processing steps and higher computing power when compared to Basemap, which is

generated at a faster speed with lower processing costs. Meanwhile, the Basemap time series may

contain a “stair-step” effect caused by repeated observations when the Basemap scene composition

was kept constant due to a lack of new cloud-free scenes (e.g., Fig. 8, OFR D, early March 2021).

By keeping the same image composition, the Basemap algorithm avoids generating synthetic pixel

values while still providing a cloud-free observation. Nonetheless, it is important to keep in mind

that there could be scenarios (e.g., when there is a lack of a new cloud-free scene for weeks or

more) in which the Basemap may have the same number of observations as PlanetScope, hence

decreasing its monitoring capabilities.

Our findings have important implications to future hydrological studies that aim to monitor

small water bodies at large scale and high temporal frequency. For the OFRs in eastern Arkansas,

the Basemap and Planet Fusion surface area time series helped unravel sub-weekly changes in

OFR surface area, as well as yearly seasonality (Fig. 8). OFRs surface area changes are pivotal

information for calculation of OFR water volume inflows and outflows. This can be achieved by

combining the surface area time series with the area-volume equations (e.g., hypsometry), which

are derived using the OFRs’ geometric shape and depth (Avisse et al., 2017; Yao et al., 2018;

Zhang et al., 2016). Estimating OFRs volume change helps bridge one of the key limitations when

modeling the cumulative impacts of OFRs on surface hydrology, as OFR water volume change is

commonly assumed to be equal to all OFRs located in a watershed (Fowler et al., 2015; Habets et

al., 2018; Hughes and Mantel, 2010). In addition, as the number of OFRs is projected to increase

globally (Downing, 2010; Habets et al., 2018), understanding the impact of OFRs on surface

hydrology is pivotal when seeking indicators to determine the optimal spatial distribution and

number of OFRs, as well as their storage capacities and water management plans aiming to

146
mitigate downstream impacts. Beyond implications to hydrological studies, we demonstrated that

Basemap and Planet Fusion can be used to monitor surface area changes for a network of OFRs

(Fig. 8). This information can be used by regulatory agencies to create water status reports to

improve regional water management and water use efficiency. These reports would be especially

relevant during the dry critical period of the year when farmers are frequently irrigating.

5. KNOWN ISSUES AND LIMITATIONS

We applied Basemap and Planet Fusion imagery for a one-year analysis. More research is

necessary to assess the performance of these datasets for a longer study period (e.g., including

periods of prolonged droughts ~3–5 years) and in other study regions—for example, in Southern

India, where OFRs are common (Vanthof and Kelly, 2019), and where there is a monsoon climate

in which there could be weeks without a clear-sky image (Ahmad et al., 2020). In addition, the

validation of this study was conducted using cloud-free SkySat imagery; therefore, there is still a

need to further evaluate the performance of both datasets under cloudy conditions. However, this

will require extensive field work, including visiting multiple OFRs on cloudy days, which imposes

several challenges, as most of the OFRs in eastern Arkansas are located on private properties.

Furthermore, we assumed that OFR surface area would vary within known and limited boundaries

(i.e., OFR shapefile buffered to 20 m). However, different results might be obtained if the Basemap

and Planet Fusion images are used to monitor water bodies that frequently change their

boundaries—water impoundments that are located close to water streams and rivers that flood

frequently, impacting the edges of water bodies. Lastly, although we calculated the uncertainties

introduced by Basemap and Planet Fusion, when using these datasets for monitoring purposes, it

would be helpful to have an estimated uncertainty accompanying every surface area observation.

For instance, whenever there are repeated observations by the Basemap or continuous synthetic

147
observations from Planet Fusion, the uncertainties from these images will be higher; however, as

of now, we cannot estimate an observation based uncertainty.

6. CONCLUSIONS

We presented a novel application of Basemap and Planet Fusion analysis ready datasets to

monitor sub-weekly OFRs surface area changes. We tested both datasets to monitor 340 OFRs of

different sizes, and we found that these datasets can be employed to monitor OFRs with

uncertainties < 10% when compared to an independent validation dataset and with uncertainties <

5% when compared to PlanetScope imagery. While Basemap had higher surface area variability

and it was more susceptible to the presence of cloud shadows and haze, Planet Fusion had a

smoother time series with less variability and fewer abrupt changes throughout the year. Given

that the surface area classification can be impacted by the OFR environmental conditions (e.g.,

presence of vegetation inside the OFR), therefore limiting the use of these datasets, we recommend

assessing the OFRs’ surface area time series before employing them for monitoring purposes. As

the number of OFRs is expected to increase globally, the use of these datasets is of great

importance to understanding OFR sub-weekly, seasonal and inter-annual surface area changes,

and to improving freshwater management by allowing better assessment and management of

OFRs.

148
REFERENCES

Ahmad, S.K., Hossain, F., Eldardiry, H., Pavelsky, T.M., 2020. A Fusion Approach for Water
Area Classification Using Visible, Near Infrared and Synthetic Aperture Radar for South
Asian Conditions. IEEE Trans. Geosci. Remote Sens. 58, 2471–2480.
https://doi.org/10.1109/TGRS.2019.2950705

Avisse, N., Tilmant, A., Müller, M.F., Zhang, H., 2017. Monitoring small reservoirs’ storage
with satellite remote sensing in inaccessible areas. Hydrol. Earth Syst. Sci. 21, 6445–
6459. https://doi.org/10.5194/hess-21-6445-2017

Cheng, Y., Vrieling, A., Fava, F., Meroni, M., Marshall, M., Gachoki, S., 2020. Phenology of
short vegetation cycles in a Kenyan rangeland from PlanetScope and Sentinel-2. Remote
Sens. Environ. 248, 112004. https://doi.org/10.1016/j.rse.2020.112004

Cooley, S., Smith, L., Stepan, L., Mascaro, J., 2017. Tracking Dynamic Northern Surface Water
Changes with High-Frequency Planet CubeSat Imagery. Remote Sens. 9, 1306.
https://doi.org/10.3390/rs9121306

Csillik, O., Asner, G.P., 2020. Near-real time aboveground carbon emissions in Peru. PLOS
ONE 15, e0241418. https://doi.org/10.1371/journal.pone.0241418

Csillik, O., Kumar, P., Asner, G.P., 2020. Challenges in Estimating Tropical Forest Canopy
Height from Planet Dove Imagery. Remote Sens. 12, 1160.
https://doi.org/10.3390/rs12071160

Csillik, O., Kumar, P., Mascaro, J., O’Shea, T., Asner, G.P., 2019. Monitoring tropical forest
carbon stocks and emissions using Planet satellite data. Sci. Rep. 9, 17831.
https://doi.org/10.1038/s41598-019-54386-6

Dadap, N.C., Hoyt, A.M., Cobb, A.R., Oner, D., Kozinski, M., Fua, P.V., Rao, K., Harvey, C.F.,
Konings, A.G., 2021. Drainage Canals in Southeast Asian Peatlands Increase Carbon
Emissions. AGU Adv. 2, e2020AV000321.

DeVries, B., Huang, C., Lang, M., Jones, J., Huang, W., Creed, I., Carroll, M., 2017. Automated
Quantification of Surface Water Inundation in Wetlands Using Optical Satellite Imagery.
Remote Sens. 9, 807. https://doi.org/10.3390/rs9080807

Downing, J.A., 2010. Emerging global role of small lakes and ponds: little things mean a lot.
Limnetica 29, 9–24. https://doi.org/10.23818/limn.29.02

Downing, J.A., Prairie, Y.T., Cole, J.J., Duarte, C.M., Tranvik, L.J., Striegl, R.G., McDowell,
W.H., Kortelainen, P., Caraco, N.F., Melack, J.M., Middelburg, J.J., 2006. The global
abundance and size distribution of lakes, ponds, and impoundments. Limnol. Oceanogr.
51, 2388–2397. https://doi.org/10.4319/lo.2006.51.5.2388

149
Doxani, G., Vermote, E., Roger, J.-C., Gascon, F., Adriaensen, S., Frantz, D., Hagolle, O.,
Hollstein, A., Kirches, G., Li, F., Louis, J., Mangin, A., Pahlevan, N., Pflug, B.,
Vanhellemont, Q., 2018. Atmospheric Correction Inter-Comparison Exercise. Remote
Sens. 10, 352. https://doi.org/10.3390/rs10020352

Du, Z., Li, W., Zhou, D., Tian, L., Ling, F., Wang, H., Gui, Y., Sun, B., 2014. Analysis of
Landsat-8 OLI imagery for land surface water mapping. Remote Sens. Lett. 5, 672–681.
https://doi.org/10.1080/2150704X.2014.960606

Fowler, K., Morden, R., Lowe, L., Nathan, R., 2015. Advances in assessing the impact of hillside
farm dams on streamflow. Australas. J. Water Resour. 19, 96–108.
https://doi.org/10.1080/13241583.2015.1116182

Frantz, D., 2019. FORCE—Landsat + Sentinel-2 Analysis Ready Data and Beyond. Remote
Sens. 11, 1124. https://doi.org/10.3390/rs11091124

Frantz, D., Haß, E., Uhl, A., Stoffels, J., Hill, J., 2018. Improvement of the Fmask algorithm for
Sentinel-2 images: Separating clouds from bright surfaces based on parallax effects.
Remote Sens. Environ. 215, 471–481. https://doi.org/10.1016/j.rse.2018.04.046

Gorelick, N., Hancher, M., Dixon, M., Ilyushchenko, S., Thau, D., Moore, R., 2017. Google
Earth Engine: Planetary-scale geospatial analysis for everyone. Remote Sens. Environ.
202, 18–27. https://doi.org/10.1016/j.rse.2017.06.031

Habets, F., Molénat, J., Carluer, N., Douez, O., Leenhardt, D., 2018. The cumulative impacts of
small reservoirs on hydrology: A review. Sci. Total Environ. 643, 850–867.
https://doi.org/10.1016/j.scitotenv.2018.06.188

Habets, F., Philippe, E., Martin, E., David, C.H., Leseur, F., 2014. Small farm dams: Impact on
river flows and sustainability in a context of climate change. Hydrol. Earth Syst. Sci. 18,
4207–4222. https://doi.org/10.5194/hess-18-4207-2014

Hondula, K.L., DeVries, B., Jones, C.N., Palmer, M.A., 2021. Effects of Using High Resolution
Satellite-Based Inundation Time Series to Estimate Methane Fluxes From Forested
Wetlands. Geophys. Res. Lett. 48, e2021GL092556.
https://doi.org/10.1029/2021GL092556

Houborg, R., McCabe, M., 2018a. Daily Retrieval of NDVI and LAI at 3 m Resolution via the
Fusion of CubeSat, Landsat, and MODIS Data. Remote Sens. 10, 890.
https://doi.org/10.3390/rs10060890

Houborg, R., McCabe, M.F., 2018b. A Cubesat enabled Spatio-Temporal Enhancement Method
(CESTEM) utilizing Planet, Landsat and MODIS data. Remote Sens. Environ. 209, 211–
226. https://doi.org/10.1016/j.rse.2018.02.067

Hughes, D.A., Mantel, S.K., 2010. Estimation des incertitudes lors de la simulation des impacts

150
de petites retenues agricoles sur les régimes d’écoulement en Afrique du Sud. Hydrol.
Sci. J. 55, 578–592. https://doi.org/10.1080/02626667.2010.484903

Jones, S.K., Fremier, A.K., DeClerck, F.A., Smedley, D., Pieck, A.O., Mulligan, M., 2017. Big
data and multiple methods for mapping small reservoirs: Comparing accuracies for
applications in agricultural landscapes. Remote Sens. 9.
https://doi.org/10.3390/rs9121307
Kääb, A., Altena, B., Mascaro, J., 2019. River-ice and water velocities using the Planet optical
cubesat constellation. Hydrol Earth Syst Sci 23, 4233–4247. https://doi.org/10.5194/hess-
23-4233-2019

Kimm, H., Guan, K., Jiang, C., Peng, B., Gentry, L.F., Wilkin, S.C., Wang, S., Cai, Y.,
Bernacchi, C.J., Peng, J., Luo, Y., 2020. Deriving high-spatiotemporal-resolution leaf
area index for agroecosystems in the U.S. Corn Belt using Planet Labs CubeSat and
STAIR fusion data. Remote Sens. Environ. 239, 111615.
https://doi.org/10.1016/j.rse.2019.111615

Kington, J.D., IV, Jordahl, K.A., Kanwar, A.N., Kapadia, A., Schönert, M., Wurster, K., 2019.
Spatially and Temporally Consistent Smallsat-Derived Basemaps for Analytic
Applications 2019, IN13B-0716.

Knapp, T., Kovacs, K., Huang, Q., Henry, C., Nayga, R., Popp, J., Dixon, B., 2018. Willingness
to pay for irrigation water when groundwater is scarce. Agric. Water Manag. 195, 133–
141. https://doi.org/10.1016/j.agwat.2017.10.013

Kong, J., Ryu, Y., Huang, Y., Dechant, B., Houborg, R., Guan, K., Zhu, X., 2021. Evaluation of
four image fusion NDVI products against in-situ spectral-measurements over a
heterogeneous rice paddy landscape. Agric. For. Meteorol. 297, 108255.
https://doi.org/10.1016/j.agrformet.2020.108255

Kotchenova, S.Y., Vermote, E.F., 2007. Validation of a vector version of the 6S radiative
transfer code for atmospheric correction of satellite data Part II Homogeneous
Lambertian and anisotropic surfaces. Appl. Opt. 46, 4455.
https://doi.org/10.1364/AO.46.004455

Kotchenova, S.Y., Vermote, E.F., Matarrese, R., Klemm, Jr., F.J., 2006. Validation of a vector
version of the 6S radiative transfer code for atmospheric correction of satellite data Part I:
Path radiance. Appl. Opt. 45, 6762. https://doi.org/10.1364/AO.45.006762

Li, J., Wang, S., 2015. An automatic method for mapping inland surface waterbodies with
Radarsat-2 imagery. Int. J. Remote Sens. 36, 1367–1384.
https://doi.org/10.1080/01431161.2015.1009653

Liu, Z., Yao, Z., Wang, R., 2016. Assessing methods of identifying open water bodies using
Landsat 8 OLI imagery. Environ. Earth Sci. 75, 873. https://doi.org/10.1007/s12665-016-
5686-2

151
López-Caloca, A.A., Escalante-Ramírez, B., Henao, P., 2020. Mapping small and medium-sized
water reservoirs using Sentinel-1A: a case study in Chiapas, Mexico. J. Appl. Remote
Sens. 14, 1–24. https://doi.org/10.1117/1.JRS.14.036503

McFeeters, S.K., 1996. The use of the Normalized Difference Water Index (NDWI) in the
delineation of open water features. Int. J. Remote Sens. 17, 1425–1432.
https://doi.org/10.1080/01431169608948714

Mime, M.M., Young, D.W., 1989. The Impact Of Stockwatering Ponds (Stockponds) On Runoff
From Large Arizona Watersheds. JAWRA J. Am. Water Resour. Assoc. 25, 165–173.
https://doi.org/10.1111/j.1752-1688.1989.tb05678.x

Mishra, V., Limaye, A.S., Muench, R.E., Cherrington, E.A., Markert, K.N., 2020. Evaluating the
performance of high-resolution satellite imagery in detecting ephemeral water bodies
over West Africa. Int. J. Appl. Earth Obs. Geoinformation 93, 102218.
https://doi.org/10.1016/j.jag.2020.102218

Ogilvie, A., Belaud, G., Massuel, S., Mulligan, M., Le Goulven, P., Calvez, R., 2018a. Surface
water monitoring in small water bodies: Potential and limits of multi-sensor Landsat time
series. Hydrol. Earth Syst. Sci. 22, 4349–4380. https://doi.org/10.5194/hess-22-4349-
2018

Ogilvie, A., Belaud, G., Massuel, S., Mulligan, M., Le Goulven, P., Malaterre, P.-O., Calvez, R.,
2018b. Combining Landsat observations with hydrological modelling for improved
surface water monitoring of small lakes. J. Hydrol. 566, 109–121.
https://doi.org/10.1016/j.jhydrol.2018.08.076

Otsu, N., 1979. A Threshold Selection Method from Gray-Level Histograms. IEEE Trans. Syst.
Man Cybern. 9, 62–66. https://doi.org/10.1109/TSMC.1979.4310076

Pena-Regueiro, J., Sebastiá-Frasquet, M.-T., Estornell, J., Aguilar-Maldonado, J.A., 2020.


Sentinel-2 Application to the Surface Characterization of Small Water Bodies in
Wetlands. Water 12, 1487. https://doi.org/10.3390/w12051487

Perin, V., Tulbure, M.G., Gaines, M.D., Reba, M.L., Yaeger, M.A., 2022. A multi-sensor
satellite imagery approach to monitor on-farm reservoirs. Remote Sens. Environ.
https://doi.org/10.1016/j.rse.2021.112796

Perin, V., Tulbure, M.G., Gaines, M.D., Reba, M.L., Yaeger, M.A., 2021. On-farm reservoir
monitoring using Landsat inundation datasets. Agric. Water Manag. 246, 106694.
https://doi.org/10.1016/j.agwat.2020.106694

Planet Labs Team, 2021a. Planet imagery product specifications. URL


https://assets.planet.com/docs/Planet_Combined_Imagery_Product_Specs_letter_screen.p
df (accessed 9.8.21).

152
Planet Labs Team, 2021b. Planet Fusion Monitoring Technical Specification. URL
https://assets.planet.com/docs/Planet_fusion_specification_March_2021.pdf (accessed
9.8.21).

Planet Labs Team, 2019. Planet Basemaps Product Specification. URL


https://assets.planet.com/products/basemap/planet-basemaps-product-specifications.pdf
(accessed 9.8.21).

Renwick, W.H., Smith, S.V., Bartley, J.D., Buddemeier, R.W., 2005. The role of impoundments
in the sediment budget of the conterminous United States. Geomorphology 71, 99–111.
https://doi.org/10.1016/j.geomorph.2004.01.010

Roy, D.P., Huang, H., Houborg, R., Martins, V.S., 2021. A global analysis of the temporal
availability of PlanetScope high spatial resolution multi-spectral imagery. Remote Sens.
Environ. 264, 112586. https://doi.org/10.1016/j.rse.2021.112586

Sadeh, Y., Zhu, X., Dunkerley, D., Walker, J.P., Zhang, Y., Rozenstein, O., Manivasagam, V.S.,
Chenu, K., 2021. Fusion of Sentinel-2 and PlanetScope time series data into daily 3 m
surface reflectance and wheat LAI monitoring. Int. J. Appl. Earth Obs. Geoinformation
96, 102260. https://doi.org/10.1016/j.jag.2020.102260

Schaaf, C.B., Gao, F., Strahler, A.H., Lucht, W., Li, X., Tsang, T., Strugnell, N.C., Zhang, X.,
Jin, Y., Muller, J.-P., Lewis, P., Barnsley, M., Hobson, P., Disney, M., Roberts, G.,
Dunderdale, M., Doll, C., d’Entremont, R.P., Hu, B., Liang, S., Privette, J.L., Roy, D.,
2002. First operational BRDF, albedo nadir reflectance products from MODIS. Moderate
Resolut. Imaging Spectroradiometer MODIS New Gener. Land Surf. Monit. 83, 135–
148. https://doi.org/10.1016/S0034-4257(02)00091-3

Sheng, Y., Song, C., Wang, J., Lyons, E.A., Knox, B.R., Cox, J.S., Gao, F., 2016. Representative
lake water extent mapping at continental scales using multi-temporal Landsat-8 imagery.
Remote Sens. Environ. 185, 129–141. https://doi.org/10.1016/j.rse.2015.12.041
Shults, D.D., Nowlin, W.J., Yaeger, M.A., Massey, J.H., Reba, M.L., 2020. A spatiotemporal
analysis quantifying the need for more on-farm reservoirs to reduce groundwater use in
the Cache and L’Anguille River Regions in Northeastern AR, in: ESRI User Conference.
Online.

Tanre, D., Deroo, C., Duhaut, P., Herman, M., Morcrette, J.J., Perbos, J., Deschamps, P.Y.,
1990. Technical note Description of a computer code to simulate the satellite signal in the
solar spectrum: the 5S code. Int. J. Remote Sens. 11, 659–668.
https://doi.org/10.1080/01431169008955048

Vanthof, V., Kelly, R., 2019. Water storage estimation in ungauged small reservoirs with the
TanDEM-X DEM and multi-source satellite observations. Remote Sens. Environ. 235,
111437. https://doi.org/10.1016/j.rse.2019.111437

153
Wang, J., Yang, D., Chen, S., Zhu, X., Wu, S., Bogonovich, M., Guo, Z., Zhu, Z., Wu, J., 2021.
Automatic cloud and cloud shadow detection in tropical areas for PlanetScope satellite
images. Remote Sens. Environ. 264, 112604. https://doi.org/10.1016/j.rse.2021.112604

Wang, Z., Zhang, R., Zhang, Q., Zhu, Y., Huang, B., Lu, Z., 2019. An Automatic Thresholding
Method For Water Body Detection From SAR Image, in: 2019 IEEE International
Conference on Signal, Information and Data Processing (ICSIDP). pp. 1–4.
https://doi.org/10.1109/ICSIDP47821.2019.9172964

Yaeger, M.A., Massey, J.H., Reba, M.L., Adviento-Borbe, M.A.A., 2018. Trends in the
construction of on-farm irrigation reservoirs in response to aquifer decline in eastern
Arkansas: Implications for conjunctive water resource management. Agric. Water
Manag. 208, 373–383. https://doi.org/10.1016/j.agwat.2018.06.040

Yaeger, M.A., Reba, M.L., Massey, J.H., Adviento-Borbe, M.A.A., 2017. On-farm irrigation
reservoirs in two Arkansas critical groundwater regions: A comparative inventory. Appl.
Eng. Agric. 33, 869–878. https://doi.org/10.13031/aea.12352

Yang, X., Qin, Q., Yésou, H., Ledauphin, T., Koehl, M., Grussenmeyer, P., Zhu, Z., 2020.
Monthly estimation of the surface water extent in France at a 10-m resolution using
Sentinel-2 data. Remote Sens. Environ. 244, 111803.
https://doi.org/10.1016/j.rse.2020.111803

Yang, X., Zhao, S., Qin, X., Zhao, N., Liang, L., 2017. Mapping of Urban Surface Water Bodies
from Sentinel-2 MSI Imagery at 10 m Resolution via NDWI-Based Image Sharpening.
Remote Sens. 9, 596. https://doi.org/10.3390/rs9060596

Yao, F., Wang, J., Yang, K., Wang, C., Walter, B.A., Crétaux, J.F., 2018. Lake storage variation
on the endorheic Tibetan Plateau and its attribution to climate change since the new
millennium. Environ. Res. Lett. 13. https://doi.org/10.1088/1748-9326/aab5d3

Zhang, S., Foerster, S., Medeiros, P., de Araújo, J.C., Motagh, M., Waske, B., 2016. Bathymetric
survey of water reservoirs in north-eastern Brazil based on TanDEM-X satellite data. Sci.
Total Environ. 571, 575–593. https://doi.org/10.1016/j.scitotenv.2016.07.024

Zhu, Z., Woodcock, C.E., 2012. Object-based cloud and cloud shadow detection in Landsat
imagery. Remote Sens. Environ. 118, 83–94. https://doi.org/10.1016/j.rse.2011.10.028

154
CHAPTER 5: ASSESSING THE IMPACTS OF ON-FARM RESERVOIRS ON

MODELED SURFACE HYDROLOGY

Prepared for submission to Water Research Resources

Authors
Vinicius Perin1, Mirela G. Tulbure1, Shiqi Fang2, A. Sankarasubramanian Arumugam2 , Michele
L. Reba3, Mary A. Yaeger4

Affiliations
1
Center for Geospatial Analytics, North Carolina State University, 2800 Faucette Drive, Raleigh,
NC, 27606, USA
2
Department of Civil, Construction and Environmental Engineering, North Carolina State
University, 961 Partners Way, Raleigh 27695, USA
3
USDA-Agricultural Research Service, Delta Water Management Research Unit, Jonesboro, AR,
72467, USA
4
Center for Applied Earth Science and Engineering Research, The University of Memphis, 3675
Alumni Drive, Memphis, TN, 38152, USA

Acknowledgements
The first author is supported by NASA through the Future Investigators in NASA Earth

and Space Science and Technology fellowship.

Abstract
On-farm reservoirs (OFRs) are essential water bodies to meet global irrigation needs. Farmers use

OFRs to store water from precipitation and runoff during the rainy season to irrigate their crops

during the dry season. Despite their importance to crop irrigation, OFRs can have a cumulative

impact on surface hydrology by decreasing flow and peak flow. Nonetheless, there is limited

knowledge on the spatial and temporal variability of the OFRs' impacts on surface hydrology.

155
Therefore, we leveraged the latest developments in the Soil Water Assessment Tool+ (SWAT+)

in combination with a novel remote sensing-based algorithm—used to monitor the OFRs

dynamics—to assess the impacts of OFRs in a small sub-watershed located in eastern Arkansas,

which is the third most irrigated state in the USA. Our results show that the presence of the OFRs

on the watershed may decrease annual flow on average between 14 and 24%, and the mean

reduction in peak flow could vary between 43 and 60%. In addition, the impact of the OFRs is not

equally distributed across the watershed, and it varies according to the OFRs spatial distribution,

and their capacity (i.e., size of the OFR). We provide a new framework that can support water

agencies with information to assess the cumulative impacts of OFRs on the watersheds where they

are constructed, aiming to support surface water resources management. This is relevant as the

number of OFRs is expected to increase globally, partially to adapt to climate change under severe

drought conditions.

Keywords: On-farm reservoirs, Surface hydrology, SWAT+, Water management

156
1. INTRODUCTION

1.1 The impacts of OFRs on surface hydrology and different modeling techniques
Inland water bodies (e.g., lakes and reservoirs) comprise a small fraction of Earth’s surface;

however, they are responsible for storing the vast majority of the accessible fresh water resources

available on Earth. In addition, water bodies are pivotal components of surface hydrology, having

key roles in ecosystems and human and animal habitats (Khazaei et al., 2022; Verpoorter et al.,

2014). On-farm reservoirs are essential to meet global irrigation needs (Döll et al., 2009; Downing,

2010; Van Den Hoek et al., 2019). Farmers use OFRs to store water from precipitation and runoff

during the rainy season to irrigate their crops during the dry season (Habets et al., 2018; Perin et

al., 2021; Vanthof & Kelly, 2019; Yaeger et al., 2017; Yaeger et al., 2018). The number of OFRs

is expected to rise worldwide in the coming decades, and estimates show that there are more than

2.6 million OFRs in the US alone (Downing, 2010; Renwick et al., 2005). The OFRs are often

built to manage surface water resources more efficiently, and to help mitigate the impact of

extreme droughts driven by climate change (Habets et al., 2018; Van Der Zaag & Gupta, 2008).

Although the OFRs are small water bodies (< 50 ha), they can have cumulative impacts on the

local and remote hydrology in the watersheds where they occur (e.g., decreasing flow and peak

flow) (Habets et al., 2018), and their impact may contribute to worsening the surface water stress

already intensified by climate change and population growth (Vörösmarty et al., 2010). Most

studies have focused on the cumulative impact of cascade of multi-purpose reservoirs on surface

water availability and flow alteration (Chalise et al., 2021; Mukhopadhyay et al., 2021), but limited

analysis has been performed on the impact of OFRs on surface water availability.

To quantify the impact of OFRs on surface hydrology, it is necessary to understand the

spatial and temporal variability of OFRs, as well as how the impacts are related to the OFRs

157
network, as the impacts of OFRs are not the sum of the individual OFR impacts, but rather the sum

and their interaction effects (Canter & Kamath, 1995; Habets et al., 2018). By gathering

information from more than 30 studies conducted in multiple countries (e.g., USA, France, Brazil),

Habets et al., (2018) did a thorough assessment of the OFRs’ impact on surface hydrology, and

the different types of models and ways to represent the OFRs on the watershed. The authors

concluded that the modeled OFRs impacts have a wide range, and that most of the studies reported

the mean annual reduction on flow, which ranged between 0.2 and 36%. In addition, the variability

of the impact was higher when assessing the low flows, with reductions between 0.3 and 60%. In

general, the estimated mean annual reduction in flow was 13.4% ± 8.0%, and the mean decrease

in peak flow could be up to 45% (Habets et al., 2018).

According to Habets et al., (2018), the different procedures used to quantify the impacts of

the OFRs can be divided into two classes: 1) data-driven methods, and 2) hydrological modeling.

Within the data-driven class, there are three main methods. The first method relies on assessing

measured inflows and outflows of selected OFRs aiming to quantify their hydrological functioning

with the assumption that the cumulative impacts are the sum of individual impacts (Culler et al.,

1961; Dubreuil & Girard, 1973; Kennon, 1966). A variation of the cumulative impact assessment

approach has been recently suggested by Hwang et al., (2021) by comparing the naturalized flows

and the controlled flows for assessing the impact of large reservoir systems. The second method

is based on doing a statistical analysis of the observed discharge time series of a watershed as the

OFRs increased in numbers over a period of time (Galéa et al., 2005; Schreider et al., 2002). This

approach is limited when discriminating the specific impact of OFRs from those of land use and

land cover change, and when explicitly representing the OFRs in the models, given that OFRs tend

to be aggregated within the entire basin. And the third method relies on conducting a paired-

158
catchment experiment by comparing the flows from two adjacent and similar catchments, one with

the OFRs and the other without the OFRs (Thompson, 2012). This technique requires that the

catchment properties (e.g., soils, topology, lithology, land cover) to be spatially homogeneous,

which is practically unattainable at a large scale, hence limiting this method’s applications.

The second class of methods relate to hydrological modeling, and it is the most widely used

approach for assessing the OFRs’ impacts. A variety of models have been proposed by coupling

the OFRs’ water balance with a quantitative approach to estimate the OFRs’ water volume change

(Fowler et al., 2015; Habets et al., 2014; Jalowska & Yuan, 2019; Yongbo et al., 2014; Ni &

Parajuli, 2018; Perrin, 2012; Zhang et al., 2012). In general, the models have three main

components: the OFR water balance, the quantitative approach to quantify the OFR inflows, and

the spatial representation of the OFRs network. These different model components result in

different limitations and assumptions—a complete assessment of these three components and how

they impact the hydrological simulations is provided in a recent review (Habets et al., 2018).

Therefore, when selecting a specific model to assess the impacts of the OFRs, it is important to

account for the model’s suitability for the target issue to be addressed, as well as the model

limitations and assumptions. In addition, from a modeler’s perspective, the easy access to

documentation and guidelines is crucial when deciding which model to use. In this regard, the Soil

and Water Assessment Tool (SWAT) (Arnold et al., 2012) has been widely used to model the

impacts of the OFRs (Jalowska & Yuan, 2019; Kim & Parajuli, 2014; Yongbo et al., 2014; Ni et

al., 2020; Ni & Parajuli, 2018; Perrin, 2012; Rabelo et al., 2021; Zhang et al., 2012), in part given

by a comprehensive collection of model documentation and guidelines available online

(https://swat.tamu.edu/).

159
1.2 The Soil Water Assessment Tool to model the impacts of OFRs on surface hydrology
The SWAT model is a time-continuous semi-distributed hydrological model widely used

across the globe—more than 5,000 peer reviewed publications since its launch in the early 1980s

(SWAT Team, 2022). The large number of SWAT applications globally revealed the model

development needs and its limitations. To address the present and future challenges when

modeling with SWAT, the model source code has undergone major modifications, and a

completely revised version of the model was proposed in SWAT+ (Bieger et al., 2017). SWAT+

uses the same equations as SWAT to simulate the hydrological processes; however, it offers more

flexibility to users when configuring the model (e.g., when defining management schedules,

routing constituents, and connecting managed flow systems to the natural stream network) (Bieger

et al., 2017).

The SWAT+ is under constant improvements (Chawanda et al., 2020; Molina-Navarro et

al., 2018), and a new module (Molina-Navarro et al., 2018) was recently developed to allow the

optimal integration of a water body and its drainage area within the simulated hydrological

processes. In previous versions of the model, when delineating the watershed area draining into a

water body, the users were required to place an outlet in a certain point of the water stream's

network, and the areas in-between the rivers’ subbasins flowing into the water body were therefore

excluded—if these areas are disregarded, important hydrological processes (e.g., evaporation,

overland and/or groundwater flow) flowing into the water body are not accounted for (Molina-

Navarro et al., 2018). This former approach can lead to inaccuracies when delineating the

watershed areas, especially when the results are used as input to an OFR model component. The

newest versions of SWAT+ consider the OFRs’ outline (i.e., shape and surface area) when

delineating the watersheds; hence, accounting for the entire drainage area flowing into the

waterbody (Mollina-Navarro et al., 2018). In addition, the latest versions allow adding more than

160
one OFR per subbasin by associating the OFR with channels—components of the watersheds, and

finer divisions and extensions of water stream reaches—enabling the modeling analyses at the

channel scale. When simulating the impact of the OFRs at the channel scale, there is a higher level

of detail of where and when the OFRs are contributing to changes in surface hydrology, unlike the

previous versions of the model, which allowed adding only a single OFR per subbasin placed at

the subbasin outlet as a point (Arnold et al., 2012), and therefore, the analyses were conducted at

the subbasin scale.

Given that OFRs tend to occur in high numbers (e.g., > 100), multiple studies leveraged

the latest developments and availability of satellite imagery to monitor the occurrence and

dynamics of OFRs (Jones et al., 2017; Ogilvie et al., 2018, 2020; Perin et al., 2022; Perin et al.,

2021a, 2021b; Van Den Hoek et al., 2019; Vanthof & Kelly, 2019). These studies allowed

quantifying the number of OFRs and their variability in surface water area and storage in the

watershed where they occur, providing relevant information for modeling the impact of the OFRs

on surface hydrology. Nonetheless, only a few studies have coupled remote sensing derived

information with hydrological modeling for quantifying OFR impacts (Yongbo et al., 2014; Ni

& Parajuli, 2018; Zhang et al., 2012).

In this study, we propose a new framework to quantify the impact of the OFRs on surface

hydrology using the latest SWAT+ model developments in combination with remote sensing

derived information. We focused on eastern Arkansas, a region that has experienced a steady

development of OFRs during the past four decades (Shults et al., 2020; Yaeger et al., 2017);

however, there is no study assessing the impact of the OFRs on surface water availability. The

proposed framework aims to assess the new module added to SWAT+ (i.e., to improve the

simulations of OFRs) that has not been thoroughly tested since it is relatively new when compared

161
to SWAT’s previous versions. Furthermore, the framework integrates remote sensing based

information—which is lacking in most of studies assessing the OFRs impacts—by leveraging a

digitally-mapped OFRs dataset (Yaeger et al., 2017) available for the study region to account for

the OFRs spatial variability in the watershed, and a novel remote sensing assimilation algorithm

(Perin et al., 2022) to account for the OFRs variability in surface area. Therefore, the objective of

this manuscript is to employ the new framework to account for the OFRs spatial and temporal

variability in surface water area, and to quantify the intra- and-inter annual impacts of the OFRs

on flow and peak flow at the channel scale.

2. METHODS

2.1 Study region


The study region is located in eastern Arkansas, USA, the third most-irrigated state in the

country (ERS-USDA, 2017). The area has a humid subtropical climate with a 30-year annual

average precipitation of ~1300 mm/year (PRISM Climate Group, 2022). The precipitation is

distributed mostly between March and May, receiving an average of ~400 mm during these months

(Perin et al., 2021b). The region has experienced a steady increase in irrigated agriculture, with

commonly irrigated crops including corn, rice, and soybeans (NASS-USDA, 2017). Furthermore,

a recent study (Yaeger et al., 2017) digitally mapped 330 OFRs located in our study region (Fig.

1) using the high-resolution (1-m) National Agricultural Imagery Program archive in combination

with 2015 sub-meter spatial resolution Google Earth satellite imagery. Most of the OFRs (95%) in

our study area have a surface area < 50 ha, and they are concentrated in the eastern portion of the

study region (Fig. 1).

162
Figure 1–Study region located in eastern Arkansas, USA, the subbasins and surface water streams

and channels delineated with SWAT+, the model outlet, the United States Geological Survey

(USGS) stations (USGS, 2022) used for flow calibration and validation, the digitized OFRs

(Yaeger et al., 2017), and the Digital Elevation Model (DEM) used in the modeling (Farr et al.,

2007).

163
2.2 SWAT+ model setup

We modeled the impact of OFRs on surface hydrology using the QSWAT+ (v.2.1.9)

SWAT+ model interface together with SWAT+ Editor (v.2.1.0) to set up the model, to input the

required datasets (e.g., DEM, land use and land cover layer, interpolated meteorological climate

information), and to run the different modeling scenarios.

The modeled watershed (710,700 ha, Fig. 1) included 68 subbasins and a total of 642

Hydrological Response Units (HRUs)—HRUs are unique portions of the subbasins that have

unique land use and management, and soil attributes. We set up daily simulations for 30 years

(1990–2020), including five years of model warm up to establish the initial soil water conditions

and hydrological processes. The watershed was delineated using the Shuttle Radar Topography

Mission DEM (30 m) (Farr et al., 2007). In addition, we set the channel length threshold to 6 km2,

and the stream length threshold to 60 km2. We placed an outlet in the southern part of the study

region—where the lowest part of the watershed is located (Fig. 1). We created the HRUs using the

dominant option—this option selects the largest HRU within the subbasin as the general HRU—

within QSWAT+ interface, and used the National Land Cover Database (30 m) (Homer et al.,

2020), and Gridded Soil Survey Geographic Database (gSSURGO) (Soil Survey Staff, USDA-

NRCS, 2021) (100 m) as inputs to the model. The gSSURGO layers were processed according to

their guidelines when using them on QSWAT+ (George, 2020). For climate data, we extracted the

centroid coordinates of each subbasin (Muche et al., 2020), and used these centroids to download

30 years of daily precipitation, minimum and maximum temperature, surface downward shortwave

radiation, wind velocity, and relative humidity from the Gridded Surface Meteorological Datasets

(Abatzoglou, 2013), available in Google Earth Engine (Gorelick et al., 2017). The time series of

each subbasin centroid was added into the SWAT+ Editor as independent weather stations.

164
2.3 Model calibration and validation procedures
We used monthly measured flow from three USGS stations (Fig. 1 and Table 1) to calibrate

and validate the model flow simulations. The USGS flow time series length varied between 14 and

25 years, and we used 60% of the timeseries for calibration and 40% for validation for each USGS

station (Table 1). We assessed the performance of the model by calculating the Coefficient of

determination (r2), Percent bias (PBIAS, %, Eq. 1) (Yapo et al., 1996), and the Nash–Sutcliffe

model efficiency coefficient (NSE, Eq. 2) (Nash & Sutcliffe, 1970). PBIAS is the relative mean

difference between the simulated and the measured flow values, and it reflects the ability of the

model to simulate monthly flows. The optimal PBIAS is zero, and low-magnitude values indicate

better model performance. Positive PBIAS indicates overestimation bias, whereas negative values

denote underestimation bias. The NSE expresses how well the model simulates flows, and it ranges

from a negative value to one, with one indicating a perfect fit between the simulated and measured

flow values. In general, the model simulations of monthly flow are considered satisfactory when

r2 ranges from 0.60 to 0.75, PBIAS ranges from ±10% to ±15%, and NSE ranges from 0.50 to 0.70

(Moriasi et al., 2015).

Table 1–USGS stations, drainage areas, and the periods used for flow calibration and validation.
USGS station Station id Drainage Area (ha) Period (years)

Calibration Validation

07264000 (A) 53,600 1995–2010 2010–2020

07263555 (B) 25,400 2007–2014 2014–2020

07263580 (C) 5,300 1997–2011 2011–2020

∑𝑛𝑖= 1 (𝑌𝑖 − 𝑋𝑖)


PBIAS = (1)
∑𝑛𝑖= 1 𝑋𝑖

∑𝑛𝑖= 1 (𝑋𝑖 − 𝑌𝑖)2


NSE = 1 − (2)
∑𝑛𝑖= 1 (𝑋𝑖 − x̅i )2

165
Where Xi is the measured flow and Yi is the simulated flow.

We conducted a sensitivity analysis using the SWAT+ ToolBox (v.0.7.6) (Chawanda,

2022) to reveal the most sensitive parameters when simulating flow—a total of 10 parameters

(Table B1) were tested based on previous studies that used SWAT/SWAT+ to model the impact

of water impoundments on surface hydrology (Jalowska & Yuan, 2019; Yongbo et al., 2014; Ni

et al., 2020; Ni & Parajuli, 2018; Perrin, 2012; Rabelo et al., 2021; Zhang et al., 2012). Following

the sensitivity analysis, we selected the five most sensitive parameters (Table 2), and proceeded

with a manual calibration using the SWAT+ Toolbox. We aimed to improve the model's monthly

flow predictions by testing the parameters one at time and changing their values between -20% to

20% with 5% increments based on their range values. The final calibrated parameters and their

fitted values are shown in Table 2.

Table 2–Monthly flow calibration parameters.


Parameter Description Range Value

CN2 SCS runoff curve number 35–95 0.20*

SOL_AWC Available water capacity (mm/mm) 0.01–1 -0.20*

ESCO Soil evaporation compensation coefficient 0.01–1 0.50

PERCO Percolation coefficient (fraction) 0–1 0.60

CANMX Maximum canopy storage (mm) 0–100 75


*Denotes relative percentage change.
2.4 OFRs representation in SWAT+
Multiple OFRs can be added to the same subbasin by associating them with channels (Dile

et al., 2022). The OFRs need to have at least one outlet channel, and they may have none or

multiple inlets. Therefore, most OFR-related processes within the model involve determining what

channels form inflowing and outflowing channels for each OFR. Ideally, each OFR would interact

with a channel, and therefore, have a channel entering, leaving, or within the OFR. Nonetheless,

166
it is common to have OFRs that do not intersect with any channel (Dile et al., 2022)—this is the

case for 93% of the OFRs in our study region. The OFRs from our study region are not dammed

along the streams, but rather they are engineered water impoundments that are indirectly connected

to the main streams via pipes and pumps (Yeager et al., 2017). A possible solution would be

modifying the OFRs’ shapes by dragging them to the closest channel (Dile et al., 2022). However,

this would require extensive modifications of the OFRs’ shapes. In addition, when an OFR is added

to a channel, this channel is split into two channels, and the model needs to account for the two

newly created channels during the water routing calculations. For this reason, adding multiple

OFRs to the same channel, or adding multiple OFRs closely located to the same channel, can be a

cumbersome process that leads to numerous routing errors.

To overcome these challenges, we aggregated the OFRs’ surface area, and added

aggregated OFRs to the model. This adaptation involved two steps. First, for each of the 330 OFRs,

we searched for the closest channel by calculating the distance between the OFRs’ centroid and

the multiple channels within each subbasin. Then, we aggregated all the OFRs that were associated

with each channel by summing up their surface area, and adding a polygon of the aggregated area

to represent the aggregated OFR. This approach resulted in 69 aggregated OFRs that were added

to 67 different channels located in 17 subbasins. The surface area of the aggregated OFRs varied

between 3.05 ha and 165.67 ha, and the number of OFRs in each aggregated OFR varied between

2 and 12. To avoid confusion, for the rest of the manuscript, we refer to OFRs as the aggregated

OFRs, and not the individual OFRs shown in Fig. 1.

2.5 Scenario Analysis


Given our representation of the OFRs in SWAT+, we assessed the impact of the OFRs on

surface hydrology at the channel scale. To do so, we established the model baseline scenario

167
without the presence of the OFRs on the watershed. In addition, we divided the channels into four

classes (i.e., low and high flow classes) according to their mean baseline flow. The different class

intervals were calculated using the mean flow quartiles accounting for all channels, which resulted

in the following baseline flow classes: (1) 0.001–0.25 m3/s, (2) 0.25–0.50 m3/s, (3) 0.50–2.11 m3/s,

and (4) 2.11–17.50 m3/s.

To account for the OFRs variation in surface area, we created three modeling scenarios

using daily OFRs surface area time series—these scenarios were based on the methodology

proposed by Perin et al., (2022). The authors used a multi-sensor satellite imagery approach with

the Kalman filter (Kalman, 1960) to derive daily OFRs’ surface area change between 2017 and

2020. The proposed algorithm accounts for the uncertainties in both the sensor's observations and

the resulting surface areas. By improving the OFRs surface area observations cadence, the

algorithm allows further understanding of the OFRs surface area intra- and inter-annual changes,

which are key pieces information that can be used to better assess and manage the water stored by

the OFRs (Perin et al., 2022). The daily surface area time series—derived by combining

PlanetScope, RapidEye, and Sentinel-2 satellite imagery (Perin et al., 2022)—of each OFR was

used to simulate three scenarios (i.e., lower, mean, and upper) representing the OFRs’ capacity in

terms of surface area. The mean scenario represents the regular condition of the OFRs, and it is

the mean of the daily surface area time series derived from the Kalman filter. The lower and upper

scenarios represent the lowest and highest capacities of the OFRs, and they are based on the surface

area 95% confidence interval limits, calculated using the daily time series. For each scenario, the

OFRs were simulated at full capacity (i.e., maximum storage at the lower, mean and upper

scenarios), and this capacity was kept constant during the simulation period (Ni et al., 2020; Ni &

Parajuli, 2018; Perrin, 2012). To assess the impact of the OFRs on surface hydrology, we compared

168
the baseline flow with the flow simulated by each surface area scenario—i.e., comparing the flow

changes with and without OFRs, a common approach used by previous studies (Habets et al.,

2018).

We estimated the impact of the OFRs on surface hydrology by calculating the percent

change (Eq. 3) of monthly flow between the baseline and the three surface area scenarios including

all OFRs. The annual impact on flow was calculated by averaging the mean percent change along

the months. We also calculated the distribution of the percent change for each baseline flow class.

The distribution was assessed using 2-D Kernel Density estimation (KDE) plots. Different from

discrete bins (e.g., histograms), the KDE plots show a continuous density estimate of the

observations using a Gaussian kernel. In addition, we assessed the percent changes in peak flow.

For the purposes of this analysis, peak flow is defined as equal or higher than the 99 th flow

percentile calculated using the entire flow time series.

𝑌𝑖 − 𝑋𝑖
Percent change (%) = ( ) ∗ 100 (3)
𝑋𝑖

Where Xi is the baseline flow and Yi is the simulated flow of each surface area scenario.

3. RESULTS

3.1 Model calibration and validation


The model calibration and validation were done using the three USGS stations presented

in Fig. 1 and Table 1. When comparing the monthly simulated flow with the measured flow for

the calibration period, there was a good agreement (0.71 ≦ r2 ≦ 0.93), and a satisfactory model

efficiency (0.68 ≦ NSE ≦ 0.90) for all three stations (Fig. 2). In addition, the PBIAS magnitude

was < 3% for station A, and < 12% for stations B and C. Meanwhile, the validation period had r 2

ranging between 0.69 and 0.86, and the NSE between 0.68 and 0.83, with PBIAS magnitude <

10% for stations A and B, and 18.12% for station C. In general, for stations A and C, the model

169
overestimated flow values (i.e., positive PBIAS) mostly during flow events < 3 m3/s, and the model

underestimated flow (i.e., negative PBIAS) for station B during flows > 20 m3/s (Fig. 2). These

findings are consistent with a previous study conducted in western Mississippi near our study

region (Ni & Parajuli, 2018). Even though during the validation period the station B had PBIAS

magnitude higher than 15%, the r2 and NSE values from the calibration and validation periods

indicate satisfactory modeling performance when simulating monthly flow (Moriasi et al., 2015).

170
Figure 2–Flow calibration and validation time series for the three USGS stations A (07264000),

B (07263555) and C (07263580). See Fig. 1 and Table 1 for more information about the USGS

stations. The precipitation time series represents the monthly accumulated precipitation at the

watershed scale (i.e., for the entire study region).

171
3.2 Percent change in flow
We assessed the impact of the OFRs on flow by comparing the baseline flow (i.e., without the

OFRs) with the three surface area scenarios generated from the Kalman filter approach—lower,

mean, and upper. The total surface area (i.e., summing all OFRs surface area) was 2.176 ha for the

lower, 2.766 ha for the mean, and 3.370 ha for the upper, and the three scenarios had a similar

OFRs surface area distribution (Fig. 3). In addition, most of the OFRs had surface areas < 50 ha—

78%, 71%, and 62% of the OFRs for the lower, mean, and upper scenarios.

Figure 3–OFRs surface area distribution for the three surface area scenarios, lower, mean, and

upper.

The impact of the OFRs varied throughout the year, and the largest impacts occurred

between January and May for all flow classes (Fig. 4). During these months, including all surface

area scenarios, the mean decrease in flow (i.e., negative mean percent change) was -34.4 ± 6% for

172
class 1, -37.6 ± 5% for class 2, -30.0 ± 6% for class 3, and -34.1 ± 6% for class 4. For all classes,

the greatest reduction in flow occurred during the month of March (~ -40%). Meanwhile, the

impact of the OFRs was smaller during the second half of the year, in which the mean percent

change in flow was -12.0 ± 3.% for class 1, -12.5 ± 5% for class 2, -1.4 ± 4% for class 3, and -2.6

± 10% for class 4 (Fig. 4).

When assessing the mean percent change per month, for all surface area scenarios, the

lower flow classes (i.e., (1) 0.001–0.25 m3/s and (2) 0.25–0.50 m3/s) had a negative mean percent

change for all months. Nonetheless, we observed a mean positive percent change (i.e., increase in

flow) for the months of August (5.0 ± 1%) and October (5.2 ± 0.2%) for class 3, and during June

(8.2 ± 0.3%), August (7.3 ± 0.4%), and October (8.7 ± 0.4%) for class 4 (Fig. 4). Furthermore, the

different surface area scenarios had similar impacts on flow for all months of the year with

differences smaller than 5% for all scenarios. Between January and May, for all flow classes, the

mean percent change was -32.0 ± 6% for the lower, –34.6 ± 7% for the mean, and -35.8 ± 5% for

the upper. Between June and December, the impact on flow was -5.4 ± 6% for the lower, -7.3 ±

8% for the mean, and -8.9 ± 5% for the upper.

173
Figure 4–Monthly percent change in flow between the baseline scenario (vertical dotted blue line)

and the three surface area scenarios (lower, mean, and upper), and for the four flow classes (1)

0.001–0.25 m3/s, (2) 0.25–0.50 m3/s, (3) 0.50–2.11 m3/s, and (4) 2.11–17.50 m3/s.

174
In general, the OFRs contributed to decreased monthly flow. However, the OFRs' impact

on flow had a significant intra- and inter-annual variability, and it varied according to different

OFRs and channels—this is highlighted by the boxplots size variability in Fig. 4, in which the

variability was lower during the first part of the year, and greater between July and August. In

addition, the monthly percent change in flow KDE plots (Fig. 5) shows that for the three scenarios,

and all flow classes, most of the changes in flow ranged between -40% and 0%. In addition, all

KDE plots have a triangular shape with its base on the smaller flows, denoting where most of the

changes occur. Even though the majority of the percent change in flow is negative, there are

circumstances in which the OFRs could positively impact flow—the increase in flow is represented

by faded colors in each surface area scenario (Fig. 5). The positive mean percent change could be

as high as 80%—see Fig. 5 for the larger flow classes, (3) 0.50–2.11 m3/s and (4) 2.11–17.50 m3/s.

The positive impact on flow for these classes occurred during the months of June, August and

October when a mean positive change is observed (Fig. 4 classes 3 and 4).

The annual mean percent change, for all surface area scenarios, was -22.5 ± 3% for class

1, -24.2 ± 4% for class 2, -14.6 ± 3% for class 3, and -16.6 ± 3% for class 4. In addition, the surface

area scenarios annual changes were -18.0 ± 5% for the lower, -19.6 ± 5% for the mean, and -20.8

± 6% for the upper, including all flow classes. The differences between the surface area scenarios

shown in Fig. 4 and Fig. 5 are related to the variability of the OFRs surface area.

175
Figure 5–Kernel density estimation plots smoothed using a Gaussian kernel for the monthly

percent change in flow between the baseline scenario (vertical dotted blue line) and the three

surface area scenarios (lower, mean, and upper) for the four flow classes (1) 0.001–0.25 m3/s, (2)

0.25–0.50 m3/s, (3) 0.50–2.11 m3/s, and (4) 2.11–17.50 m3/s. Note the different range of values of

the y-axis for all four flow classes.

176
177
3.3 Impact on peak flow
For each channel, we calculated the impact of the OFRs on peak flow (Fig. 6). The impact

on peak flow was -60.7 ± 13% for class 1, -56.2 ± 11% for class 2, -46.7 ± 19% for class 3, and -

43.9 ± 12% class 4. When assessing the impact on peak flow based on different surface area

scenarios, the mean percent change was -49.4 ± 18% for the lower, -50.4 ± 17% for the mean, and

-52.7 ± 18% for the upper. All peak flows occurred between January and May, which is the period

of the year when the study region receives most of its precipitation (Perin et al., 2021). With the

exception of a few outliers, there was no increase in peak flow, even though the OFRs contributed

to a positive mean percent change in flow in certain months of the year (Fig. 4 classes 3 and 4).

Figure 6–Percent change in peak flow between the baseline scenario (vertical dotted blue line)

and the three surface area scenarios (lower, mean, and upper) for the four flow classes (1) 0.001–

0.25 m3/s, (2) 0.25–0.50 m3/s, (3) 0.50–2.11 m3/s, and (4) 2.11–17.50 m3/s.

3.4 Simulated flow time series


We randomly selected a channel within the flow class 3 to demonstrate the baseline and

the three surface area scenarios’ flow time series between 1995 and 2005 (Fig. 7). For this channel,

the annual mean percent changes in flow when comparing the baseline scenario with the lower,

178
mean, and upper surface area scenarios were 0.99 ± 11.8%, -1.9 ± 13%, and -2.0 ± 19%—the high

standard deviation for the three scenarios is explained by the interannual variability. The upper

surface area scenario resulted in lower flows (i.e., higher impact) when compared to the lower and

mean scenarios for the majority of the flow events—67.8% and 57.6% for the lower and mean

scenarios. Nonetheless, there are circumstances when the upper scenario yielded higher flows—

32.2% and 42.4% of the events for the lower and mean scenarios (e.g., see the two insets 03/1997–

08/1998 and 05/2002–02/2004). These findings indicate that the impacts that the OFRs have on

flow are not entirely governed by the presence and surface area of the OFRs (i.e., the different

surface area scenarios), instead by a combination of the OFRs with different modeling components

(e.g., terrain, land use, climate information), and different hydrological processes (e.g., run-off,

precipitation, evaporation). In addition, the impact on peak flow for this channel was -45.7 ± 19.7%

for all surface area scenarios—this is highlighted on two occasions (08/2002 and 08/2003) during

the second inset.

179
Figure 7–A subset of the time series of simulated flow for baseline and the three surface area

scenarios (lower, mean, and upper) between 1995 and 2005 for a selected channel within the flow

class 3.

4. DISCUSSION

When simulating water impoundments in SWAT/SWAT+, it is common practice to

validate and calibrate the model using flow measurements (Evenson et al., 2018; Habets et al.,

2018; Jalowska & Yuan, 2019; Ni & Parajuli, 2018). In addition, other studies have validated and

calibrated the model using alternative variables. For example, Perrin et al., (2012) employed

monthly measurements of piezometric variations to assess aquifer recharge processes, and

Jalowska & Yuan (2019) used sediment loadings (concentration and budget), from field

monitoring reports to evaluate sediment simulations. Ideally, we would calibrate and validate the

180
model by accounting for the parameters governing the OFRs’ water budget (e.g., inflows and

outflows) (e.g., Kim and Parajuli, 2014). Nonetheless, these measurements were not available for

the OFRs in our study region. Furthermore, a thorough calibration and validation of the model

would require extra streamflow data, covering other parts of the study region, as the three USGS

stations—the only data available—used in this study are located in the upper part of the modeled

watershed. Similar to Evenson et al., (2018)—who proposed a module to better represent spatially

distributed wetlands, and validated their model using a direct (i.e., flow measurement) and an

indirect (i.e., the wetlands surface area) approach—our validation and calibration was done using

the flow measurements, and the OFRs surface area scenarios were based on an algorithm that was

validated with an independent higher spatial resolution dataset (Perin et al., 2022).

There is a consensus within the scientific community that the OFRs will have a cumulative

impact on surface hydrology by decreasing flow and peak flow, and the impact will vary from

basin to basin due to the number of OFRs, and the OFRs’ different purposes (e.g., different

irrigation schedule) (Ayalew et al., 2017; Fowler et al., 2015; Habets et al., 2018; Nathan & Lowe,

2012; Pinhati et al., 2020; Rabelo et al., 2021). As pointed out by Habets et al., (2018) the mean

annual decrease in flow from all studies was -13.4% ± 8%. This value is aligned with our results,

which varied between -24.2 ± 4% and -14.6 ± 3% for all flow classes. In addition, OFRs can reduce

peak flow on average by 45% (Habets et al., 2018; Nathan and Lowe, 2012; Thompson, 2012),

and it can reach up to 70% (Ayalew et al., 2017) for certain flow events. Likewise, our results are

consistent with these findings, in which the mean impact on peak flow varied between -60.7 ± 12%

and -43.9 ± 12%. Furthermore, differently from previous research, our results (Fig. 4, classes 3

and 4) showed that the OFRs may have a positive (< 9%) impact on flow. This could be explained

by the level of details in our analyses. While we calculated the monthly impact on flow at the

181
channel scale by aggregating the OFRs to the closest channel, previous studies have mostly

reported the annual impact on flows (Habets et al., 2018), and they performed their analysis at the

subbasin scale by aggregating the OFRs to a single point at the outlet of each subbasin in SWAT

(Evenson et al., 2018; Kim & Parajuli, 2014; Perrin, 2012; Zhang et al., 2012), or they used

different modeling approaches (see Habet et al., (2018)).

Be leveraging the latest improvements in SWAT to simulate water impoundments (Molina-

Navarro et al., 2018) in combination with a novel algorithm to monitor OFRs (Perin et al., 2022),

we modeled the impact of the OFRs on flow at the channel scale. In addition, the surface area

scenarios enabled us to account for events when the OFRs were at the lowest, regular, and fullest

capacities according to their surface area. This is an improvement over previous studies (e.g., Ni

et al., 2020; Ni and Parajuli, 2018; Perrin, 2012) that used a single surface area (i.e., one snapshot

in time) to represent the OFRs in SWAT. The small differences (< 5%) between the surface area

scenarios in terms of mean percent change on monthly flow indicates that the OFRs’ surface area

variation had a low impact on the model. For instance, during January and May the mean monthly

percent change ranged between -35.8 ± 6% and -32.0 ± 7%, and during June and December it

varied between -8.8 ± 5% and -5.4 ± 6% for the three surface area scenarios. The same was

observed for peak flow, with a mean monthly impact ranging between -52.7 ± 17% and -49.4 ±

18%. This small variability on flow impact was observed even though the total OFR surface area

increased by 590 ha and 1194 ha when comparing the lower scenario with the mean and upper

scenarios (Fig. 5). However, the OFRs remained a small portion (< 1%) of the total area of the

modeled watershed (Fig. 1). These findings could be related to the fact that flow simulations are

governed by several hydrological processes (e.g., run-off, precipitation, evapotranspiration)

besides the presence of the OFRs on the channel (Bieger et al., 2017; Dile et al., 2022; Arnold et

182
al., 2012). In addition, when assessing the percent change in flow at the channel scale, the

differences in surface area between the scenarios occur at a lower magnitude when compared to

the total OFRs surface area. For instance, an OFR with surface area smaller than 10 ha, and with

surface area variations between 10 and 20% for the three scenarios, may not lead to differences

(e.g., > 10%) between the three scenarios.

Our findings highlight that the impacts of the OFRs on flow and peak flow have a

significant intra- and inter-annual variability (Figs. 4, 5, 7), and the impacts vary according to

different OFRs and channels (Fig. 4). The largest impacts on flow occurred during the first part of

the year between January and May, the period of the year when the peak flows occur. In addition,

this time of the year also coincides with the period when the region receives most of its

precipitation (Perin et al., 2021b), and the OFRs are at their fullest capacity (i.e., OFRs storing

their maximum amount of water) (Perin et al., 2022). During the second part of the year, we

observed a milder mean percent change in flow for all flow classes and all scenarios, and a greater

variability in percent change, notably for the months of July and August (Fig. 4). Moreover, most

of the irrigation activities happen between June and September (Perin et al., 2021b, Yaeger et al.,

2017), and it is when the OFRs are at their lowest capacities (i.e., smaller surface areas) (Perin et

al., 2022), which could explain their moderate impact and higher variability during these months—

even though we are not accounting for the OFRs inflows and outflows, and not simulating

irrigation events. Additionally, the variability of the OFRs impacts is related to the OFRs’ physical

properties (e.g., surface area and location in the watershed). For example, the OFRs’ surface area

will have an impact on flow and peak flow, as shown by the different surface area scenarios, and

depending on where the OFRs are located in the watershed, given that they may be connected to

lower or higher flow channels, which contributes to their impact variability during the year (Figs.

183
3 and 4). Besides the OFRs’ physical properties, the built-in complexity of the model—when

simulating the presence of the OFRs and the various hydrological processes (e.g., run-off,

precipitation, evapotranspiration) governing the water cycle—contributes to the differences in the

OFRs impacts. This complexity is illustrated in Fig. 7 showing that the upper scenario can have a

higher or lower impact on flow when compared to the lower and mean scenarios.

Overall, we presented a new framework to quantitatively analyze the impact of a network

of OFRs on flow and peak flow, and we described the various potential reasons behind the

variability of the impacts. Our results indicate that OFRs do not have an equally distributed impact

on flow and peak flow across the watershed. Hence, assessing the OFRs location as well as their

numbers across the watershed is important when aiming to manage the construction of new OFRs.

In particular, the geospatial variability of the OFRs impacts could be considered by water agencies

when planning and developing a network of OFRs, given it is possible to identify the areas that

are under high pressure (e.g., regions with multiple OFRs that are having a significant impact on

flow and peak), and to identify areas that could benefit from the construction of new OFRs,

targeting improvements on water resources management and irrigation activities. Furthermore,

even though the impact of the OFRs may vary significantly in different watersheds (Habets et al.,

2018), our framework could be transferable to other places across the world as it integrates satellite

based inputs derived using novel algorithms and Earth Observations data sources in combination

with the latest SWAT+ hydrological modeling developments to quantify the impact of OFRs on

surface hydrology. This is relevant as the number of OFRs is expected to increase globally (Althoff

et al., 2020; Habets et al., 2014; Habets et al., 2018; Krol et al., 2011; Rodrigues et al., 2012), with

a limited knowledge of how the OFRs may impact surface hydrology in different watersheds, and

under diverse environmental conditions. Finally, in tandem with the OFRs’ key role on irrigated

184
food production, in part to adapt to climate change (Habets et al., 2018) and to alleviate the pressure

on groundwater resources (Vanthof & Kelly, 2019; Yaeger et al., 2017; Yaeger et al., 2018), their

impacts on surface hydrology need to be considered to avoid exacerbating the surface water stress

already intensified by climate change and population growth (Vörösmarty et al., 2010).

5. FUTURE IMPROVEMENTS

Future improvements should focus on how to better represent the OFRs water management

(i.e., OFRs inflows and outflows) in SWAT+. Given that each OFR has an independent water

balance, accounting for the OFRs water volume change would be a more realistic representation

of the OFRs when compared to the three surface area scenarios tested in this study. Estimating the

OFRs volume change can be done by combining the OFRs surface area time series with area-

elevation equations—these equations describe the OFRs’ bathymetry, and allow volume

estimation by inputting the OFRs’ surface area (Liebe et al., 2005; Meigh, 1995; Sawunyama et

al., 2006). After carefully assessing different methods to derive these equations (Arvor et al., 2018;

Avisse et al., 2017; Li et al., 2021; Meigh, 1995; Sawunyama et al., 2006; Vanthof & Kelly, 2019;

Yao et al., 2018; Zhang et al., 2016), we decided that measured ground-data of the OFRs’ depth—

which is not available—is required to estimate the equations with an acceptable uncertainty.

Estimating the area-elevation equations entails several challenges, including: 1) despite the fact

that there are several DEMs available for the study region (Arkansas GIS Office, 2022)—DEMs

can be used to estimate the OFRs bottom elevation—the DEMs were collected when most of the

OFRs were full (i.e., bathymetry was not exposed), which limits their use in this case; and 2)

although the OFRs are located within the same geomorphological region, they have different

depth, shape and physical characteristics (Perin et al., 2022; Yaeger et al., 2017). Therefore, even

if a generalized area-elevation equation was calculated for our study region—this is a common

185
approach done by other studies (Mady et al., 2020; Vanthof & Kelly, 2019)—that would still lead

to high uncertainties of water volume changes. Ideally, each OFR would have its own equation,

which was not possible when this study was done.

Efforts should also be made to improve SWAT+ capabilities to receive measured OFRs’

inflows and outflows. The latest version of the model has improved the hydrological representation

of small water impoundments in SWAT+ (Mollina-Navarro et al., 2018). Nonetheless, at the time

of our study, the newest version of the model does not allow users to input measured or calculated

OFRs’ inflows and outflows. Instead, the model developers recommend simulating the OFRs

water balance using decision tables (Arnold et al., 2018; Dile et al., 2022). However, there are very

limited guidelines on how to create these decision tables. In addition, the tables would simulate

the OFRs water balance instead of using the measured or calculated volume change, which could

introduce more uncertainties to the modeling scenarios.

6. CONCLUSION

We assessed the latest developments in SWAT+ to simulate the impacts of a network of

OFRs in combination with satellite based inputs derived using novel algorithms and Earth

Observations data sources. Our study showed that the OFRs may have an impact on flow and peak

flow, which can have a significant inter- and intra-annual variability. The impact of the OFRs is

not equally distributed across the watershed, and it varies according to the OFRs spatial

distribution, and their surface area. As the number of OFRs is expected to increase globally—

partially to adapt to climate change and to alleviate pressure on groundwater resources—and

therefore, also increase their relevance to irrigated food production, it is imperative to develop new

frameworks to further understand the OFRs impacts on surface hydrology. In this regard, we

186
provided a combination of different methods that can be used in other watersheds, which can

support water agencies with information to improve surface water resources management.

187
REFERENCES

Abatzoglou, J. T. (2013). Development of gridded surface meteorological data for ecological


applications and modelling. International Journal of Climatology, 33(1), 121–131.
https://doi.org/10.1002/joc.3413

Althoff, D., Rodrigues, L. N., & da Silva, D. D. (2020). Impacts of climate change on the
evaporation and availability of water in small reservoirs in the Brazilian savannah.
Climatic Change, 159(2), 215–232. https://doi.org/10.1007/s10584-020-02656-y

Arkansas GIS Office. (2022, June 20). Arkansas GIS Office | Elevation datasets. Retrieved June
20, 2022, from https://gis.arkansas.gov/

Arnold, J. G., Moriasi, D. N., Gassman, P. W., Abbaspour, K. C., M. J. White, M. J., Srinivasan,
R., et al. (2012). SWAT: Model Use, Calibration, and Validation. Transactions of the
ASABE, 55(4), 1491–1508. https://doi.org/10.13031/2013.42256

Arnold, J. G., Bieger, K., White, M. J., Srinivasan, R., Dunbar, J. A., & Allen, P. M. (2018). Use
of Decision Tables to Simulate Management in SWAT+. Water, 10(6), 713.
https://doi.org/10.3390/w10060713

Arvor, D., Daher, F. R. G., Briand, D., Dufour, S., Rollet, A.-J., Simões, M., & Ferraz, R. P. D.
(2018). Monitoring thirty years of small water reservoirs proliferation in the southern
Brazilian Amazon with Landsat time series. ISPRS Journal of Photogrammetry and
Remote Sensing, 145, 225–237. https://doi.org/10.1016/j.isprsjprs.2018.03.015

Avisse, N., Tilmant, A., Müller, M. F., & Zhang, H. (2017). Monitoring small reservoirs’ storage
with satellite remote sensing in inaccessible areas. Hydrology and Earth System Sciences,
21(12), 6445–6459. https://doi.org/10.5194/hess-21-6445-2017

Ayalew, T. B., Krajewski, W. F., Mantilla, R., Wright, D. B., & Small, S. J. (2017). Effect of
Spatially Distributed Small Dams on Flood Frequency: Insights from the Soap Creek
Watershed. Journal of Hydrologic Engineering, 22(7), 04017011.
https://doi.org/10.1061/(ASCE)HE.1943-5584.0001513

Bieger, K., Arnold, J. G., Rathjens, H., White, M. J., Bosch, D. D., Allen, P. M., et al. (2017).
Introduction to SWAT+, A Completely Restructured Version of the Soil and Water
Assessment Tool. JAWRA Journal of the American Water Resources Association, 53(1),
115–130. https://doi.org/10.1111/1752-1688.12482

Canter, L. W., & Kamath, J. (1995). Questionnaire checklist for cumulative impacts.
Environmental Impact Assessment Review, 15(4), 311–339. https://doi.org/10.1016/0195-
9255(95)00010-C

Chalise, D. R., Sankarasubramanian, A., & Ruhi, A. (2021). Dams and Climate Interact to Alter
River Flow Regimes Across the United States. Earth’s Future, 9(4).
https://doi.org/10.1029/2020EF001816

188
Chawanda, C. J. (2022, June 17). SWAT+ Toolbox. Retrieved June 17, 2022, from
https://celray.github.io/docs/swatplus-toolbox/v1.0/index.html

Chawanda, C. J., Arnold, J., Thiery, W., & van Griensven, A. (2020). Mass balance calibration
and reservoir representations for large-scale hydrological impact studies using SWAT+.
Climatic Change, 163(3), 1307–1327. https://doi.org/10.1007/s10584-020-02924-x

Culler, R. C., Hadley, R. F., & Schumm, S. A. (1961). Hydrology of the upper Cheyenne River
basin: Part A. Hydrology of stock-water reservoirs in upper Cheyenne River basin; Part
B. Sediment sources and drainage-basin characteristics in upper Cheyenne River basin
(USGS Numbered Series No. 1531) (Vol. 1531). U.S. Geological Survey.
https://doi.org/10.3133/wsp1531

Dile, Y., Srinivasan, R., & George, C. (2022). QGIS Interface for SWAT+: QSWAT+ (No. 2.2)
(p. 137). Retrieved from https://swatplus.gitbook.io/docs/user/qswat+

Döll, P., Fiedler, K., & Zhang, J. (2009). Global-scale analysis of river flow alterations due to
water withdrawals and reservoirs. Hydrology and Earth System Sciences, 13(12), 2413–
2432. https://doi.org/10.5194/hess-13-2413-2009

Downing, J. A. (2010). Emerging global role of small lakes and ponds: little things mean a lot.
Limnetica, 29(1), 9–24. https://doi.org/10.23818/limn.29.02

Dubreuil, P., & Girard, G. (1973). Influence of a very large number of small reservoirs on the
annual flow regime of a tropical stream. Washington DC American Geophysical Union
Geophysical Monograph Series, 17, 295–299.

Economic Research Service-United States Department of Agriculture (ERS-USDA). (2017).


Irrigation & Water Use. Retrieved October 30, 2021, from
https://www.ers.usda.gov/topics/farm-practices-management/irrigation-water-use/

Evenson, G. R., Jones, C. N., McLaughlin, D. L., Golden, H. E., Lane, C. R., DeVries, B., et al.
(2018). A watershed-scale model for depressional wetland-rich landscapes. Journal of
Hydrology X, 1, 100002. https://doi.org/10.1016/j.hydroa.2018.10.002

Farr, T. G., Rosen, P. A., Caro, E., Crippen, R., Duren, R., Hensley, S., et al. (2007). The Shuttle
Radar Topography Mission. Reviews of Geophysics, 45(2).
https://doi.org/10.1029/2005RG000183

Fowler, K., Morden, R., Lowe, L., & Nathan, R. (2015). Advances in assessing the impact of
hillside farm dams on streamflow. Australasian Journal of Water Resources, 19(2), 96–
108. https://doi.org/10.1080/13241583.2015.1116182

Galéa, G., Vasquez-Paulus, B., Renard, B., & Breil, P. (2005). L’impact des prélèvements d’eau
pour l’irrigation sur les régimes hydrologiques des sous-bassins du Tescou et de la
Séoune (bassin Adour-Garonne, France). Revue des sciences de l’eau / Journal of Water
Science, 18(3), 273–305. https://doi.org/10.7202/705560ar

189
George, C. (2020). Using SSURGO soil data with QSWAT and QSWAT+ (No. 3) (p. 10).
Retrieved from https://swat.tamu.edu/media/116594/us_soilsv3.pdf

Gorelick, N., Hancher, M., Dixon, M., Ilyushchenko, S., Thau, D., & Moore, R. (2017). Google
Earth Engine: Planetary-scale geospatial analysis for everyone. Remote Sensing of
Environment, 202, 18–27. https://doi.org/10.1016/j.rse.2017.06.031

Habets, F., Philippe, E., Martin, E., David, C. H., & Leseur, F. (2014). Small farm dams: Impact
on river flows and sustainability in a context of climate change. Hydrology and Earth
System Sciences, 18(10), 4207–4222. https://doi.org/10.5194/hess-18-4207-2014

Habets, Florence, Molénat, J., Carluer, N., Douez, O., & Leenhardt, D. (2018). The cumulative
impacts of small reservoirs on hydrology: A review. Science of The Total Environment,
643, 850–867. https://doi.org/10.1016/j.scitotenv.2018.06.188

Homer, C., Dewitz, J., Jin, S., Xian, G., Costello, C., Danielson, P., et al. (2020). Conterminous
United States land cover change patterns 2001–2016 from the 2016 National Land Cover
Database. ISPRS Journal of Photogrammetry and Remote Sensing, 162, 184–199.
https://doi.org/10.1016/j.isprsjprs.2020.02.019

Hwang, J., Kumar, H., Ruhi, A., Sankarasubramanian, A., & Devineni, N. (2021). Quantifying
Dam‐Induced Fluctuations in Streamflow Frequencies Across the Colorado River Basin.
Water Resources Research, 57(10). https://doi.org/10.1029/2021WR029753

Jalowska, A. M., & Yuan, Y. (2019). Evaluation of SWAT Impoundment Modeling Methods in
Water and Sediment Simulations. JAWRA Journal of the American Water Resources
Association, 55(1), 209–227. https://doi.org/10.1111/1752-1688.12715

Jones, S. K., Fremier, A. K., DeClerck, F. A., Smedley, D., Pieck, A. O., & Mulligan, M. (2017).
Big data and multiple methods for mapping small reservoirs: Comparing accuracies for
applications in agricultural landscapes. Remote Sensing, 9(12).
https://doi.org/10.3390/rs9121307

Kalman, R. E. (1960). A new approach to linear filtering and prediction problems.

Kennon, F. W. (1966). Hydrologic effects of small reservoirs in Sandstone Creek Watershed,


Beckham and Roger Mills Counties, western Oklahoma. U.S. G.P.O. Retrieved from
https://doi.org/10.3133/wsp1839C

Khazaei, B., Read, L. K., Casali, M., Sampson, K. M., & Yates, D. N. (2022). GLOBathy, the
global lakes bathymetry dataset. Scientific Data, 9(1), 36. https://doi.org/10.1038/s41597-
022-01132-9

Kim, H. K., & Parajuli, P. B. (2014). Impacts of Reservoir Outflow Estimation Methods in
SWAT Model Calibration. Transactions of the ASABE, 1029–1042.
https://doi.org/10.13031/trans.57.10156

Krol, M. S., de Vries, M. J., van Oel, P. R., & de Araújo, J. C. (2011). Sustainability of Small

190
Reservoirs and Large Scale Water Availability Under Current Conditions and Climate
Change. Water Resources Management, 25(12), 3017–3026.
https://doi.org/10.1007/s11269-011-9787-0

Li, Y., Gao, H., Allen, G. H., & Zhang, Z. (2021). Constructing Reservoir Area–Volume–
Elevation Curve from TanDEM-X DEM Data. IEEE Journal of Selected Topics in
Applied Earth Observations and Remote Sensing, 14, 2249–2257.
https://doi.org/10.1109/JSTARS.2021.3051103

Liebe, J., van de Giesen, N., & Andreini, M. (2005). Estimation of small reservoir storage
capacities in a semi-arid environment. Physics and Chemistry of the Earth, 30(6-7 SPEC.
ISS.), 448–454. https://doi.org/10.1016/j.pce.2005.06.011

Mady, B., Lehmann, P., Gorelick, S. M., & Or, D. (2020). Distribution of small seasonal
reservoirs in semi-arid regions and associated evaporative losses. Environmental
Research Communications, 2(6), 061002. https://doi.org/10.1088/2515-7620/ab92af

Meigh, J. (1995). The impact of small farm reservoirs on urban water supplies in Botswana.
Natural Resources Forum, 19(1), 71–83. https://doi.org/10.1111/j.1477-
8947.1995.tb00594.x
Molina-Navarro, E., Nielsen, A., & Trolle, D. (2018). A QGIS plugin to tailor SWAT watershed
delineations to lake and reservoir waterbodies. Environmental Modelling & Software,
108, 67–71. https://doi.org/10.1016/j.envsoft.2018.07.003

Moriasi, D. N., Gitau, M. W., Pai, N., & Daggupati, P. (2015). Hydrologic and water quality
models: Performance measures and evaluation criteria. Transactions of the ASABE,
58(6), 1763–1785. https://doi.org/10.13031/trans.58.10715

Muche, M. E., Sinnathamby, S., Parmar, R., Knightes, C. D., Johnston, J. M., Wolfe, K., et al.
(2020). Comparison and Evaluation of Gridded Precipitation Datasets in a Kansas
Agricultural Watershed Using SWAT. JAWRA Journal of the American Water Resources
Association, 56(3), 486–506. https://doi.org/10.1111/1752-1688.12819

Mukhopadhyay, S., Sankarasubramanian, A., & de Queiroz, A. R. (2021). Performance


Comparison of Equivalent Reservoir and Multireservoir Models in Forecasting
Hydropower Potential for Linking Water and Power Systems. Journal of Water
Resources Planning and Management, 147(4), 04021005.
https://doi.org/10.1061/(ASCE)WR.1943-5452.0001343

Nash, J. E., & Sutcliffe, J. V. (1970). River flow forecasting through conceptual models part I —
A discussion of principles. Journal of Hydrology, 10(3), 282–290.
https://doi.org/10.1016/0022-1694(70)90255-6

Nathan, R., & Lowe, L. (2012). The Hydrologic Impacts of Farm Dams. Australasian Journal of
Water Resources, 16(1), 75–83. https://doi.org/10.7158/13241583.2012.11465405

National Agricultural Statistics Service-United States Department of Agriculture (NASS-


USDA). (2017). Census of agriculture. Retrieved February 10, 2021, from

191
https://www.nass.usda.gov/AgCensus/index.php

Ni, X., & Parajuli, P. B. (2018). Evaluation of the impacts of BMPs and tailwater recovery
system on surface and groundwater using satellite imagery and SWAT reservoir function.
Agricultural Water Management, 210, 78–87.
https://doi.org/10.1016/j.agwat.2018.07.027

Ni, X., Parajuli, P. B., & Ouyang, Y. (2020). Assessing Agriculture Conservation Practice
Impacts on Groundwater Levels at Watershed Scale. Water Resources Management,
34(4), 1553–1566. https://doi.org/10.1007/s11269-020-02526-3

Ogilvie, A., Belaud, G., Massuel, S., Mulligan, M., Le Goulven, P., Malaterre, P.-O., & Calvez,
R. (2018). Combining Landsat observations with hydrological modelling for improved
surface water monitoring of small lakes. Journal of Hydrology, 566, 109–121.
https://doi.org/10.1016/j.jhydrol.2018.08.076

Ogilvie, A., Poussin, J.-C., Bader, J.-C., Bayo, F., Bodian, A., Dacosta, H., et al. (2020).
Combining Multi-Sensor Satellite Imagery to Improve Long-Term Monitoring of
Temporary Surface Water Bodies in the Senegal River Floodplain. Remote Sensing,
12(19), 3157. https://doi.org/10.3390/rs12193157

Perin, V., Roy, S., Kington, J., Harris, T., Tulbure, M. G., Stone, N., et al. (2021). Monitoring
Small Water Bodies Using High Spatial and Temporal Resolution Analysis Ready
Datasets. Remote Sensing, 13(24), 5176. https://doi.org/10.3390/rs13245176

Perin, V., Tulbure, M. G., Gaines, M. D., Reba, M. L., & Yaeger, M. A. (2021). On-farm
reservoir monitoring using Landsat inundation datasets. Agricultural Water Management,
246, 106694. https://doi.org/10.1016/j.agwat.2020.106694

Perin, V., Tulbure, M. G., Gaines, M. D., Reba, M. L., & Yaeger, M. A. (2022). A multi-sensor
satellite imagery approach to monitor on-farm reservoirs. Remote Sensing of
Environment. https://doi.org/10.1016/j.rse.2021.112796

Perrin, J. (2012). Assessing water availability in a semi-arid watershed of southern India using a
semi-distributed model. Journal of Hydrology, 13.

Pinhati, F. S. C., Rodrigues, L. N., & Aires de Souza, S. (2020). Modelling the impact of on-
farm reservoirs on dry season water availability in an agricultural catchment area of the
Brazilian savannah. Agricultural Water Management, 241, 106296.
https://doi.org/10.1016/j.agwat.2020.106296

PRISM Climate Group, Oregon State University. (2020). PRISM Gridded Climate Data.
Retrieved January 2, 2022, from https://prism.oregonstate.edu/

Rabelo, U. P., Dietrich, J., Costa, A. C., Simshäuser, M. N., Scholz, F. E., Nguyen, V. T., &
Lima Neto, I. E. (2021). Representing a dense network of ponds and reservoirs in a semi-
distributed dryland catchment model. Journal of Hydrology, 603, 127103.
https://doi.org/10.1016/j.jhydrol.2021.127103

192
Renwick, W. H., Smith, S. V., Bartley, J. D., & Buddemeier, R. W. (2005). The role of
impoundments in the sediment budget of the conterminous United States.
Geomorphology, 71(1), 99–111. https://doi.org/10.1016/j.geomorph.2004.01.010

Rodrigues, L. N., Sano, E. E., Steenhuis, T. S., & Passo, D. P. (2012). Estimation of Small
Reservoir Storage Capacities with Remote Sensing in the Brazilian Savannah Region.
Water Resources Management, 26(4), 873–882. https://doi.org/10.1007/s11269-011-
9941-8

Sawunyama, T., Senzanje, A., & Mhizha, A. (2006). Estimation of small reservoir storage
capacities in Limpopo River Basin using geographical information systems (GIS) and
remotely sensed surface areas: Case of Mzingwane catchment. Physics and Chemistry of
the Earth, Parts A/B/C, 31(15), 935–943. https://doi.org/10.1016/j.pce.2006.08.008

Schreider, S. Yu., Jakeman, A. J., Letcher, R. A., Nathan, R. J., Neal, B. P., & Beavis, S. G.
(2002). Detecting changes in streamflow response to changes in non-climatic catchment
conditions: farm dam development in the Murray–Darling basin, Australia. Journal of
Hydrology, 262(1), 84–98. https://doi.org/10.1016/S0022-1694(02)00023-9

Shults, D. D., Nowlin, W. J., Yaeger, M. A., Massey, J. H., & Reba, M. L. (2020). A
spatiotemporal analysis quantifying the need for more on-farm reservoirs to reduce
groundwater use in the Cache and L’Anguille River Regions in Northeastern AR. In ESRI
User Conference. Online. Retrieved from https://www.esri.com/en-
us/about/events/uc/esri-uc-map-gallery#/map-detail/5ef64560f3b616cb344765e2

Soil Survey Staff, USDA-NRCS. (2021). Gridded Soil Survey Geographic (gSSURGO)
Database for the Conterminous United States. United States Department of Agriculture,
Natural Resources Conservation Service. Retrieved December 20, 2021, from
https://datagateway.nrcs.usda.gov/

SWAT Team. (2022, January 6). Publications | Soil & Water Assessment Tool (SWAT).
Retrieved June 17, 2022, from https://swat.tamu.edu/publications/

Thompson, J. C. (2012, January 1). Impact and Management of Small Farm Dams in Hawke’s
Bay, New Zealand (thesis). Open Access Te Herenga Waka-Victoria University of
Wellington. https://doi.org/10.26686/wgtn.16997929.v1

United States Geological Survey (USGS). (2022). United States Geological Survey Water Data
for the Nation. Retrieved June 22, 2022, from https://waterdata.usgs.gov/nwis

Van Den Hoek, J., Getirana, A., Jung, H., Okeowo, M., & Lee, H. (2019). Monitoring Reservoir
Drought Dynamics with Landsat and Radar/Lidar Altimetry Time Series in Persistently
Cloudy Eastern Brazil. Remote Sensing, 11(7), 827. https://doi.org/10.3390/rs11070827

Van Der Zaag, P., & Gupta, J. (2008). Scale issues in the governance of water storage projects.
Water Resources Research, 44(10). https://doi.org/10.1029/2007WR006364

193
Vanthof, V., & Kelly, R. (2019). Water storage estimation in ungauged small reservoirs with the
TanDEM-X DEM and multi-source satellite observations. Remote Sensing of
Environment, 235, 111437. https://doi.org/10.1016/j.rse.2019.111437

Verpoorter, C., Kutser, T., Seekell, D. A., & Tranvik, L. J. (2014). A global inventory of lakes
based on high-resolution satellite imagery. Geophysical Research Letters, 41(18), 6396–
6402. https://doi.org/10.1002/2014GL060641

Vörösmarty, C. J., McIntyre, P. B., Gessner, M. O., Dudgeon, D., Prusevich, A., Green, P., et al.
(2010). Global threats to human water security and river biodiversity. Nature, 467(7315),
555–561. https://doi.org/10.1038/nature09440

Yaeger, M. A., Reba, M. L., Massey, J. H., & Adviento-Borbe, M. A. A. (2017). On-farm
irrigation reservoirs in two Arkansas critical groundwater regions: A comparative
inventory. Applied Engineering in Agriculture, 33(6), 869–878.
https://doi.org/10.13031/aea.12352

Yaeger, Mary A., Massey, J. H., Reba, M. L., & Adviento-Borbe, M. A. A. (2018). Trends in the
construction of on-farm irrigation reservoirs in response to aquifer decline in eastern
Arkansas: Implications for conjunctive water resource management. Agricultural Water
Management, 208, 373–383. https://doi.org/10.1016/j.agwat.2018.06.040

Yao, F., Wang, J., Yang, K., Wang, C., Walter, B. A., & Crétaux, J. F. (2018). Lake storage
variation on the endorheic Tibetan Plateau and its attribution to climate change since the
new millennium. Environmental Research Letters, 13(6). https://doi.org/10.1088/1748-
9326/aab5d3

Yapo, P. O., Gupta, H. V., & Sorooshian, S. (1996). Automatic calibration of conceptual rainfall-
runoff models: sensitivity to calibration data. Journal of Hydrology, 181(1–4), 23–48.
https://doi.org/10.1016/0022-1694(95)02918-4

Yongbo, L., Wanhong, Y., Zhiqiang, Y., Ivana, L., Jim, Y., Jane, E., & Kevin, T. (2014).
Assessing Effects of Small Dams on Stream Flow and Water Quality in an Agricultural
Watershed. Journal of Hydrologic Engineering, 19(10), 05014015.
https://doi.org/10.1061/(ASCE)HE.1943-5584.0001005

Zhang, C., Peng, Y., Chu, J., Shoemaker, C. A., & Zhang, A. (2012). Integrated hydrological
modelling of small- and medium-sized water storages with application to the upper
Fengman Reservoir Basin of China. Hydrol. Earth Syst. Sci., 15.
https://doi.org/10.5194/hess-16-4033-2012

Zhang, S., Foerster, S., Medeiros, P., de Araújo, J. C., Motagh, M., & Waske, B. (2016).
Bathymetric survey of water reservoirs in north-eastern Brazil based on TanDEM-X
satellite data. Science of the Total Environment, 571, 575–593.
https://doi.org/10.1016/j.scitotenv.2016.07.024

194
CHAPTER 6: CONCLUSION

This dissertation is a cohesive body of work of different approaches to monitor a network

of OFRs, and it offers a new framework using satellite imagery based information combined with

hydrological modeling to quantitatively assess the impacts of OFRs on surface water availability.

The results presented throughout the four manuscripts are relevant to regions where OFRs lack

monitoring and understanding of their impacts on surface hydrology. In particular, by exploring

the latest Earth Observation datasets—including analysis ready datasets, and freely available and

commercial satellite imagery—to monitor OFRs, this dissertation sets the foundation on which

datasets to employ when aiming to monitor OFRs, and what are the eventual challenges (e.g., data

acquisition and processing) and limitations (e.g., spatial and temporal resolutions) of each dataset.

Furthermore, this dissertation offers an alternative to an issue that is often ignored when modeling

the OFRs impacts on surface hydrology, which is to account for the OFRs variability (e.g., changes

in surface area) on the modeling scenarios.

Leveraging the long term Landsat inundation datasets to monitor OFRs has several

advantages, including assessing the OFRs historical surface area seasonality and frequency of

inundation, and their maximum extent. In Chapter 2, we proposed the first exploration of the

DWSE and JRC to widely monitor more than 750 OFRs for a study region in the USA. In addition,

we described the main limitations of both datasets—i.e., inconsistent time series and spatial

resolution to monitor OFRs smaller than 5 ha—when aiming to apply these datasets to other study

regions. Despite these limitations, both datasets are freely available and can be easily accessed

through Google Earth Engine (GEE)—although the DWSE was originally processed for the USA

only, the dataset can be reprocessed to other countries on GEE using open source algorithms. The

easy access to the datasets (i.e., in terms of data availability and imagery processing steps) on GEE

195
offers a robust yet inexpensive alternative to monitoring OFRs where there are limited resources,

for example, the remote regions on the Jaguaribe basin—located in Northeastern Brazil—that it is

known for having a conglomerate of OFRs.

To overcome the limitations of Landsat-based inundation datasets, in Chapter 3, we

introduced a multi-sensor satellite imagery algorithm based on the Kalman filter to decrease the

latency of satellite observations, and to increase the accuracy of OFRs’ surface area estimates. The

algorithm combines data from multiple sensors that have higher spatial and temporal resolutions

when compared to Landsat. In addition, this new approach allowed assimilating optical and radar

satellite data to obtain sub-weekly OFRs surface area observations while accounting for the

uncertainties in both the satellite observations and the algorithm estimates. Assessing the OFRs

surface area dynamics at a sub-weekly scale allows further understanding the spatial and temporal

variability of a network of OFRs, which could enable water agencies to monitor water user trends

across the watershed where the OFRs occur, and to act on management recommendations aiming

to increase water use efficiency. To make the algorithm more accessible to field managers and

other scientists, the next step for the Kalman filter algorithm would be setting up an operational

online platform in which users could easily access the OFR surface area information already

processed by the algorithm. This would be possible with online geospatial data platforms (e.g.,

GEE), and it would not only provide insights to a broader scientific community, but would also

enable professionals who do not have the background in remote sensing and GIS to access the

data.

Although the algorithm proposed in Chapter 3 enabled monitoring of the OFRs at sub-

weekly scale, the algorithm involved downloading and processing satellite imagery of different

spatial and temporal resolutions from multiple platforms, which can be a demanding task, and a

196
limiting factor when implementing the proposed methodology. An alternative to the multi-sensor

algorithm is to employ the next generation of analysis ready datasets, Planet Fusion and Planet

Basemap, to monitor OFRs. In Chapter 4, we demonstrated that these datasets allow monitoring

OFRs with high accuracy by providing daily imagery at 3 m spatial resolution. In addition, the

datasets require minimal user processing steps, as the users can swiftly derive insights (e.g., OFRs

water storage status based on surface area), and insert these insights into their data pipeline and

workflows. Future analysis should focus on expanding the use of these datasets by monitoring

OFRs in other study regions, and accounting for more than one year of analysis—in our study we

only processed Planet Fusion and Planet Basemap for one year. Ideally, these datasets should be

tested in study regions where there is continuous cloud cover during several days or weeks (e.g.,

Southern India). This would allow assessing the performance of these datasets and their

uncertainties when there is a significant amount of missing data due to cloud cover. Even though

these datasets are not free of costs; Planet Labs is focusing on making these datasets more

accessible to researchers across the globe through partnerships with government and research

institutions.

Chapters 2, 3, and 4 were dedicated to the first part of the overarching goal of this

dissertation, which was to improve OFRs monitoring while using reproducible workflows that

leveraged different satellite imagery datasets and sensors. Chapter 5 relates to the second part of

the overarching goal, and it was dedicated to combine satellite imagery based information with

hydrological modeling to estimate the OFRs’ impacts on surface water availability.

Chapter 5 presented a new framework that combines the OFRs surface area information

from the Kalman filter (Chapter 3) with the latest improvements in SWAT+ hydrological model.

This represents an advancement from previous studies that did not incorporate changes in OFRs

197
variability (e.g., surface area), and it is the first assessment of how the OFRs are impacting the

surface hydrology in eastern Arkansas. Nonetheless, there are several limitations of this study that

should be improved in future iterations of the model. For example, when the study was conducted

there was no ground-data to calibrate and validate any of the parameters controlling the OFRs

water balance inflows and outflows (e.g., total evaporation, overland flow). Another improvement

would be modifying the model so that it would allow adding measured OFRs inflows and outflows.

The current version of the model only allows the users to simulate inflows and outflows—this can

be done using decision tables; however, there is very limited documentation on how to create these

tables. Also, given that these tables simulate the OFRs volume changes instead of measuring them,

this could introduce extra uncertainties to the modeling scenarios.

The work presented in this dissertation should be considered complementary to ground

assessments and monitoring—even though OFRs occur in high numbers (i.e., thousands), which

can make field evaluations time consuming and labor intensive. For example, in order to take full

advantage of the OFRs surface area inter- and intra-annual variability information, field managers

and water agencies can combine the OFRs surface area changes with area-elevation equations

(hypsometry). These equations represent the OFRs’ bathymetry (i.e., measurement of the OFRs

depth), and allow to estimate the amount of water stored in each OFR. In general, these equations

are estimated using digital elevation models that were collected when the OFRs were empty with

the bathymetry exposed. Nonetheless, accurate measurements of the OFRs depth are required for

calibration and validation of these equations—we explored multiple methods to derive these

equations solely based on remote sensing data; however, without the ground-data to check the

validity of the equations, they were only rough estimations of the OFRs storage, which is why we

have not used the equations. Future work should focus on collecting ground-data on the OFRs

198
depth to derive these equations, and to provide the spatial and temporal variability of the water

stored by the OFRs network. The water storage information enables the monitoring of water use

trends, therefore, has the potential to improve water irrigation management recommendations

aiming to increase water use efficiency.

The latest availability and improvements on satellite imagery enabled widespread

monitoring of the dynamics of a network of OFRs across space and time. In addition, assessing

the OFRs dynamics and combining satellite based information with hydrological modeling

enhances our understanding of the OFRs impacts on surface water availability, especially where

the OFRs are poorly monitored. With the increasing importance of the OFRs to irrigated food

production, the cohesive body of work presented in this dissertation can be employed to other

regions to identify the areas where the OFRs are exerting high pressure on surface water

availability, and to identify areas that could benefit from the construction of new OFRs, while

targeting improvements on water use efficiency.

199
APPENDICES

200
APPENDIX A: SUPPLEMENTAL MATERIAL ACCOMPANYING CHAPTER 2: ON-

FARM RESERVOIR MONITORING USING LANDSAT INUNDATION DATASET

201
Table A1–All generalized additive mixed models fitted employing SAratio as the response variable

for the DSWE and JRC. Precipitation, air temperature, and NDVI were modelled as fixed effects,

and month and year as random effects. The best model was selected using the lowest Akaike

Information Criterion (AIC), the highest adjusted coefficient of determination (r2), and the least

number of parameters. The relative contribution (contri.) of each explanatory variable depicts the

variable contribution to r2. The selected models are highlighted with a bounding box.

202
DSWE HUC
Models A B C D
AIC r2 contri. AIC r2 contri. AIC r2 contri. AIC r2 contri.
1 SAratio ~ Precip. -479 0.080 - -499 0.051 - -450 0.05 - -382 0.044 -
2 SAratio ~ Precip. + Tmax -518 0.241 >100% -537 0.211 >100% -443 0.06 28% -419 0.201 >100%
3 SAratio ~ Precip. + Tmax + NDVI -489 0.277 15% -508 0.251 19% -408 0.08 32% -399 0.277 38%
4 SAratio ~ Precip. + Tmax + NDVI + (1|month) -494 0.279 1% -509 0.252 0% -406 0.08 0% -397 0.276 0%
5 SAratio ~ Precip. + Tmax + NDVI + (1|month) + (1|year) -577 0.253 -9% -572 0.238 -5% -527 0.07 -16% -424 0.274 -1%
JRC
1 SAratio ~ Precip. -310 0.082 - -356 0.076 - -345 0.048 - -267 0.042 -
2 SAratio ~ Precip. + Tmax -362 0.275 >100% -407 0.266 >100% -363 0.144 >100% -320 0.248 >100%
3 SAratio ~ Precip. + Tmax + NDVI -333 0.275 0% -372 0.266 0% -327 0.143 0% -297 0.295 19%
4 SAratio ~ Precip. + Tmax + NDVI + (1|month) -353 0.237 -14% -390 0.242 -9% -325 0.141 -1% -296 0.286 -3%
5 SAratio ~ Precip. + Tmax + NDVI + (1|month) + (1|year) -480 0.170 -28% -502 0.181 -25% -467 0.109 -23% -357 0.238 -17%

DSWE HUC
Models E F G H
AIC r2 contri. AIC r2 contri. AIC r2 contri. AIC r2 contri.
1 SAratio ~ Precip. -488 0.033 - -326 0.004 - -438 0.015 - 142 0.032 -
2 SAratio ~ Precip. + Tmax -478 0.038 16% -319 0.020 >100% -438 0.051 >100% 149 0.039 19%
3 SAratio ~ Precip. + Tmax + NDVI -445 0.056 45% -311 0.095 368% -394 0.102 100% 144 0.039 1%
4 SAratio ~ Precip. + Tmax + NDVI + (1|month) -443 0.055 0% -310 0.091 -4% -394 0.099 -3% 146 0.039 0%
5 SAratio ~ Precip. + Tmax + NDVI + (1|month) + (1|year) -574 0.039 -30% -319 0.092 1% -455 0.092 -7% -155 -0.001 -102%
JRC
1 SAratio ~ Precip. -336 0.049 - -267 0.020 - -300 0.015 - 168 0.016 -
2 SAratio ~ Precip. + Tmax -349 0.133 >100% -298 0.169 >100% -323 0.126 >100% 176 0.019 19%
3 SAratio ~ Precip. + Tmax + NDVI -314 0.129 -3% -268 0.174 3% -327 0.230 83% 168 0.031 64%
4 SAratio ~ Precip. + Tmax + NDVI + (1|month) -312 0.129 0% -266 0.174 0% -332 0.226 -2% 170 0.031 0%
5 SAratio ~ Precip. + Tmax + NDVI + (1|month) + (1|year) -454 0.106 -17% -320 0.171 -1% -397 0.205 -10% -134 -0.023 -175%

203
Table A2–Global generalized additive mixed models fitted employing SAratio as the response

variable for the DSWE and JRC. Precipitation, air temperature, and NDVI were modelled as fixed

effects, and month and year as random effects. The best model was selected using the lowest

Akaike Information Criterion (AIC), the highest adjusted coefficient of determination (r2), and the

least number of parameters. The relative contribution (contri.) of each explanatory variable depicts

the variable contribution to r2. The selected model is highlighted with a bounding box.

DSWE JRC
Models AIC r2 contri. AIC r2 contri.
1 SAratio ~ Precip. + (1|HUC) -2214.89 0.03 - -1595.53 0.03 -
2 SAratio ~ Precip. + Tmax + (1|HUC) -2290.72 0.06 >100 % -1787.68 0.11 >100 %
3 SAratio ~ Precip. + Tmax + NDVI + (1|HUC) -2222.43 0.09 45% -1717.98 0.13 17%
4 SAratio ~ Precip. + Tmax + NDVI + (1|HUC) + (1|month) -2231.35 0.09 -1% -1741.02 0.13 -3%
5 SAratio ~ Precip. + Tmax + NDVI + (1|HUC) + (1|month) + (1|year) -2704.97 0.08 -10% -2527.86 0.10 -19%

204
Figure A1–The relationship between SAratio derived from the DSWE, between climate factors and

NDVI, and the distribution of all variables, and their correlation coefficient.

205
Figure A2–The relationship between SAratio derived from the JRC, between climate factors and

NDVI, and the distribution of all variables, and their correlation coefficient.

206
Figure A3–The distribution of temperature, NDVI, and precipitation for HUCs: C, D, E, F, G,

and H, and considering the entire time series (1995–2018), and all OFRs located on these HUCs.

The black and red dashed lines represent the average, maximum and minimum temperature. The

boxplots represent the NDVI monthly variation. The violin plots represent the monthly variation

of precipitation. Please see Table 1 for HUC ID information.

207
HUC: A DSWE HUC: G JRC

HUC: G DSWE HUC: G JRC

Figure A4–Example of the moving average filter with smoother of magnitude 2 for SAratio time

series-series for HUCs A and G, and both DSWE and JRC. Please see Table 1 for HUC ID

information.

208
APPENDIX B: SUPPLEMENTAL MATERIAL ACCOMPANYING CHAPTER 5:

ASSESSING THE IMPACTS OF ON-FARM RESERVOIRS ON MODELED SURFACE

HYDROLOGY

Table B1–SWAT+ parameters tested in sensitivity analysis. Parameters are ordered from most to
least important.
Parameter Description Range

1 CN2 SCS runoff curve number 35–95

2 SOL_AWC Available water capacity (mm/mm) 0.01–1

3 ESCO Soil evaporation compensation 0.01–1


coefficient

4 PERCO Percolation coefficient (fraction) 0–1

5 CANMX Maximum canopy storage (mm) 0–100

6 EPCO Plant uptake compensation factor 0.01–1

7 REVAP_C Groundwater “revap” coefficient 0.02–0.2


O

8 ALPHA_B Base flow recession constant (days) 0.01–1


F

9 BF_MAX Baseflow rate when the entire area is 0.1–2


contributing to baseflow (mm)

10 SURLAG Surface runoff coefficient 0.05-24

209

You might also like