ArticlePDF Available

Cotton Yield Estimation from UAV-Based Plant Height

Authors:

Abstract and Figures

Accurate estimation of crop yield before harvest, especially in early growth stages, is important for farmers and researchers to optimize field management and evaluate crop performance. However, existing in-field methods for estimating crop yield are not efficient. The goal of this research was to evaluate the performance of a UAV-based remote sensing system with a low-cost RGB camera to estimate cotton yield based on plant height. The UAV system acquired images at 50 m above ground level over a cotton field at the first flower growth stage. Waypoints and flight speed were selected to allow >70% image overlap in both forward and side directions. Images were processed to develop a geo-referenced orthomosaic image and a digital elevation model (DEM) of the field that was used to extract plant height by calculating the difference in elevation between the crop canopy and bare soil surface. Twelve ground reference points with known height were deployed in the field to validate the UAV-based height measurement. Geo-referenced yield data were aligned to the plant height map based on GPS and image features. Correlation analysis between yield and plant height was conducted row-by-row with and without row registration. Pearson correlation coefficients between yield and plant height with row registration for all individual rows were in the range of 0.66 to 0.96 and were higher than those without row registration (0.54 to 0.95). A linear regression model using plant height was able to estimate yield with root mean square error of 550 kg ha ⁻¹ and mean absolute error of 420 kg ha ⁻¹ . Locations with low yield were analyzed to identify the potential reasons, and it was found that water stress and coarse soil texture, as indicated by low soil apparent electricity conductivity (EC a ), might contribute to the low yield. The findings indicate that the UAV-based remote sensing system equipped with a low-cost digital camera was potentially able to monitor plant growth status and estimate cotton yield with acceptable errors. Keywords: Cotton, Geo-registration, Plant height, UAV-based remote sensing, Yield estimation.
Content may be subject to copyright.
Transactions of the ASABE
Vol. 62(2): 393-403 2019 American Society of Agricultural and Biological Engineers ISSN 2151-0032 https://doi.org/10.13031/trans.13067 393
C
OTTON
Y
IELD
E
STIMATION FROM
UAV-B
ASED
P
LANT
H
EIGHT
A. Feng, M. Zhang, K. A. Sudduth, E. D. Vories, J. Zhou
A
BSTRACT
. Accurate estimation of crop yield before harvest, especially in early growth stages, is important for farmers
and researchers to optimize field management and evaluate crop performance. However, existing in-field methods for esti-
mating crop yield are not efficient. The goal of this research was to evaluate the performance of a UAV-based remote
sensing system with a low-cost RGB camera to estimate cotton yield based on plant height. The UAV system acquired images
at 50 m above ground level over a cotton field at the first flower growth stage. Waypoints and flight speed were selected to
allow >70% image overlap in both forward and side directions. Images were processed to develop a geo-referenced ortho-
mosaic image and a digital elevation model (DEM) of the field that was used to extract plant height by calculating the
difference in elevation between the crop canopy and bare soil surface. Twelve ground reference points with known height
were deployed in the field to validate the UAV-based height measurement. Geo-referenced yield data were aligned to the
plant height map based on GPS and image features. Correlation analysis between yield and plant height was conducted
row-by-row with and without row registration. Pearson correlation coefficients between yield and plant height with row
registration for all individual rows were in the range of 0.66 to 0.96 and were higher than those without row registration
(0.54 to 0.95). A linear regression model using plant height was able to estimate yield with root mean square error of 550 kg
ha
-1
and mean absolute error of 420 kg ha
-1
. Locations with low yield were analyzed to identify the potential reasons, and
it was found that water stress and coarse soil texture, as indicated by low soil apparent electricity conductivity (EC
a
), might
contribute to the low yield. The findings indicate that the UAV-based remote sensing system equipped with a low-cost digital
camera was potentially able to monitor plant growth status and estimate cotton yield with acceptable errors.
Keywords. Cotton, Geo-registration, Plant height, UAV-based remote sensing, Yield estimation.
ield estimation at early growth stages is an im-
portant objective for precision agriculture as it
can assist in optimizing field management, espe-
cially for management-intensive crops such as
cotton (Gossypium hirsutum L.), and in evaluating crop per-
formance (Komm and Moyer, 2015). For many crops, plant
height is an important indicator of yield because larger and
healthier plants usually have higher yield than smaller plants
in the same field. For example, Chang et al. (2017) found
that plant height was a good indicator of sorghum growth
status and yield. Many other studies also showed that plant
height has potential to estimate crop yield or biomass in dif-
ferent crops, including onion (Ballesteros et al., 2018), corn
(Sharma et al., 2016), barley (Bendig et al., 2014, 2015;
Brocks and Bareth, 2018), and wheat (Walter et al., 2018).
However, excessive height in cotton (i.e., rank growth) can
reduce yield, making the relationship weaker. For example,
Huang et al. (2016) reported that cotton in the middle of the
height range had higher yield than taller or shorter cotton.
Cotton plant height is usually adequately controlled with
plant growth regulators; in those cases, cotton yield has been
positively correlated with plant height (Sui et al., 2012; Chu
et al., 2016).
Plant height has been measured using different ground-
based sensing technologies, such as a mobile platform with
ultrasonic sensors (Scotford and Miller, 2004; Sui et al.,
2012; Sharma et al., 2016) and tractor-mounted light detec-
tion and ranging (LiDAR) sensors (Zhang and Grift, 2012).
However, ground-based sensing methods have limitations in
travel speed and may not be efficient for data collection in
large fields. In recent years, unmanned aerial vehicle (UAV)
technology has been used as a high-throughput data collec-
tion tool in many studies due to its flexibility and ability to
cover large areas quickly (Yang et al., 2017; Zhou et al.,
2018). A number of different sensors, including imaging
sensors (e.g., visible cameras, multispectral and hyperspec-
tral cameras, and thermal cameras) and LiDAR sensors, can
be integrated with a UAV platform to measure target crop
traits, such as canopy color, vegetation indices, and canopy
Submitted for review in August 2018 as manuscript number ITSC
13067; approved for publication by the Information Technology, Sensors,
& Control Systems Community of ASABE in January 2019.
Mention of company or trade names is for description only and does no
t
imply endorsement by the USDA. The USDA is an equal opportunity
p
rovider and employer.
The authors are Aijing Feng, Graduate Research Assistant, Division o
f
Food Systems and Bioengineering, University of Missouri, Columbia,
Missouri; Meina Zhang, Research Assistant, Institute of Agricultural
Facilities and Equipment, Jiangsu Academy of Agricultural Sciences,
N
anjing, China; Kenneth A. Sudduth, Research Agricultural Engineer,
and Earl D. Vories, Research Agricultural Engineer, USDA-ARS
Cropping Systems and Water Quality Research Unit, Columbia, Missouri;
Jianfeng Zhou, Assistant Professor, Division of Food Systems an
d
Bioengineering, University of Missouri, Columbia, Missouri.
Corresponding author: Jianfeng Zhou, 211 Ag Engineering Building,
University of Missouri, Columbia, MO 65211; phone: 509-882-2495;
e-mail: zhoujianf@missouri.edu.
Y
394 TRANSACTIONS OF THE ASABE
structural features, including plant height (Sankaran et al.,
2015). UAV-based plant height measurement using a Li-
DAR sensor may be very expensive (Sankey et al., 2018;
Wang et al., 2017). In recent years, low-cost cameras
mounted on a UAV to measure plant height have been stud-
ied in field conditions (de Castro et al., 2018; Jay et al.,
2015). Sequential images collected with UAV systems have
been used to reconstruct the 3D structure of individual plants
or plant rows based on the structure-from-motion (SfM)
method (Westoby et al., 2012), which is able to develop dig-
ital elevation models (DEMs) for the scouted fields using
overlapped images (Smith et al., 2016; Glendell et al., 2017).
These DEMs can be used to extract the height of objects,
including plants (Chang et al., 2017; Malambo et al., 2018;
Walter et al., 2018).
A UAV-based plant height map includes high-resolution
and site-specific information that can be used for estimation
of crop yield based on calibrated estimation models. How-
ever, development of calibration models requires accurate
alignment or registration of UAV-based height data with
ground-based yield data (Yang et al., 2017). Although geo-
referenced height and yield can be registered based on coor-
dinates from their respective GPS systems, this process may
lead to large errors in registration due to low accuracy of the
GPS receivers (Jiao et al., 2018). Low-accuracy geo-refer-
encing information plays a less important role in plot-based
research, where the plots are easily segmented from images
and registered to labeled ground data (Bendig et al., 2014;
Liu et al., 2016; Du and Noguchi, 2017; Duan et al., 2017;
Brocks and Bareth, 2018). These data are typically analyzed
at plot level, i.e., mean values of imagery features and
ground data in each plot are calculated to represent all plants
in the plot, and then the reduced number of observations
(equal to the number of plots) are used to develop estimation
models (Huang et al., 2013, 2016; Van Der et al., 2017).
Similarly, some studies have manually split a large field into
multiple plots by adding ground reference points that can be
identified in images, and then used single data points
(means) from those segmented plots for further analysis
(Duan et al., 2017; Maimaitijiang et al., 2017; Vega et al.,
2015). However, the spatial resolution of the acquired da-
tasets is relatively low, and the averaging process may re-
duce the variability to where it is not possible to develop re-
liable models.
The goal of this study was to estimate cotton yield using
imagery data from a UAV-based low-cost visible camera us-
ing a large data set. Geo-registration between UAV-esti-
mated plant height and ground-based yield data was con-
ducted. The specific objectives of this study were to: (1) ex-
tract and calibrate cotton plant height from UAV-based im-
agery data, (2) register geo-referenced yield data with UAV-
based plant height data, and (3) evaluate the feasibility of
estimating cotton yield using the low-cost imagery-based
plant height.
MATERIALS AND METHODS
EXPERIMENTAL FIELD AND GROUND
DATA COLLECTION
This study was conducted in a cotton research field at the
Fisher Delta Research Center of the University of Missouri,
located in the upper Mississippi Delta region near Portage-
ville, Missouri (36.411° N, 89.696° W). The spatial variabil-
ity of soil texture in the field is quite large due to both allu-
vial and seismic activities. Because soil texture is strongly
related to apparent soil electrical conductivity (ECa; Sudduth
et al., 2003), soil ECa data were collected in May 2016 using
a Veris 3100 instrument (Veris Technologies, Salina, Kan.)
following procedures described by Sudduth et al. (2003).
Cotton cultivar PHY 333WRF (Dow Agrosciences, Indian-
apolis, Ind.) was planted on May 23, 2017, using a commer-
cial planter (Almaco, Nevada, Iowa) at a planting rate of
12 seed m-1 and a row width of 0.97 m, resulting in 152
north-south rows over the field area of 350 m 170 m
(~6 ha).
The cotton was irrigated using a 150 m center-pivot irri-
gation system (8000 Series, Valley Irrigation, Valley, Neb.)
following a predefined protocol. The portion of the field un-
der the center pivot was divided into 15 plots with a range of
irrigation treatments, including areas with no irrigation.
There were also areas in the northeast and southeast portions
of the field that were not irrigated by the center-pivot system.
Because the objective was to estimate yield with height, in-
dependent of other factors, the irrigation treatments were not
considered in developing the yield model.
The field was harvested on October 6-7, 2017, using a
cotton harvester (1996 Case IH 2155, Racine, Wisc.)
equipped with an Ag Leader Insight yield monitor system
that measures the cotton yield using optical emitter and de-
tector arrays on opposite sides of each picker chute, for a
total of eight pairs of sensor arrays (Ag Leader Technology,
Ames, Iowa) and an Ag Leader GPS 1500 receiver. The sys-
tem recorded geo-referenced yield data points separately for
each row at 1 s intervals at approximately 6.5 km h-1, or ap-
proximately 1.8 m of row per data point. The spatially refer-
enced yield data were managed using Ag Leader SMS Basic
version 18.50 and ArcGIS 10.4 for Desktop (ESRI, Red-
lands, Cal.). The yield monitor was calibrated by weighing
loads of cotton in a boll buggy equipped with scales (Master
Scales, Greenwood, Miss.), and the resulting calibration er-
ror based on nine of the ten loads averaged 2.7% with a max-
imum of 3.6%. Some of the seed cotton from the first load
tended to fill voids in the picker basket, so the subsequent
loads emptied more consistently; therefore, the weight of the
first load was not included in the calibration. After harvest-
ing, yield data artefacts were removed with Yield Editor
software (Sudduth et al., 2012). The average seed cotton
yield on the study field was 2530 1130 kg ha-1 (mean with
standard deviation of all data points), and all yield data re-
ported in this study were seed cotton yield.
UAV DATA ACQUISITION
The UAV-based remote sensing system consisted of a
UAV platform (Matrice 600 Pro, DJI, Shenzhen, China) and
an RGB camera (HERO 5, GoPro, San Mateo, Cal.) with a
62(2): 393-403 395
resolution of 3000 4000 pixels, resulting in a spatial resolu-
tion of 26 mm pixel-1 at a flight altitude of 50 m above ground
level (AGL) from the take-off location. The RGB camera was
set to capture still images at a rate of 2 frames per second (fps),
and images were saved to an on-board SD card. Flight of the
remote sensing system was controlled using a UAV ground
control app (Autopilot, Hangar Technology, Austin, Tex.) that
allowed planning the flight path and setting waypoints, flight
speed, and altitude. Prior to flight, the flight path (waypoints)
and speed (10 km h-1) were planned according to the flight
altitude (50 m AGL) and minimum overlap requirement
(>70% in the forward and sideward directions). After the
flight, images and GPS coordinates of the flight path recorded
by the on-board system were downloaded for further pro-
cessing. Image data were collected near noon central daylight
time (CDT) on August 12, 2017, during the cotton flowering
period. In total, 1352 images were acquired by the UAV sys-
tem over the field. Environmental data on the test day were
obtained from a weather station (http://agebb.mis-
souri.edu/weather/stations/pemiscot/framepvl.htm) located
approximately 400 m from the study site. Table 1 shows the
environmental conditions during UAV data acquisition.
Twelve ground reference points (GRPs) consisting of fence
posts (~1.1 m above ground) topped with white polytechnic
boards (about 30 30 cm) were distributed at the edges and
in the middle of the field to validate the height measurement
from the UAV-based sensing system.
IMAGE PROCESSING AND PLANT HEIGHT EXTRACTION
Sequential images were processed using Agisoft Pho-
toScan Pro (ver. 1.4.2, Agisoft LLC, St. Petersburg, Russia)
to generate dense point cloud data based on the SfM method.
GPS coordinates of the flight path from the UAV’s on-board
GPS receiver were extracted and matched with each image
based on their time stamps. The geo-referenced images were
uploaded to Agisoft PhotoScan to generate an orthomosaic
image and DEM map of the study field, as shown in figure 1.
The DEM of the field included elevation information (rel-
ative to mean sea level) for the crop canopy, the bare soil
surface, and the GRPs. Plant height was calculated as the
difference in elevation between the crop canopy and the
nearby bare soil surface. The soil surface elevation could be
measured prior to crop planting; however, that was not done
in this study. Therefore, soil elevation data were extracted
from the same DEM map using procedures described below.
Although there were errors in the measured elevation due to
the UAV having a low-accuracy GPS, it was assumed that
the measurement error over a small region would be similar
for the soil surface and the plant canopy or other objects.
Therefore, the study field was split into 2,550 small plots by
evenly dividing the field with 150 lines in the north-south
direction and 17 lines in the east-west direction (red lines in
fig. 2), resulting in an area of 2 m 8.5 m for each plot. It
Table 1. Environmental data during data collection (0.5 h period).
Air
Temp.
(°C)
Relative
Humidity
(%)
Dew
Point
(°C)
Wind
Speed
(m s-1)
Wind
Direction
(degrees)
Solar
Irradiance
(W m-2)
26.8 63 19.1 1.3 1 989
Figure 1. Results of processed sequential images from UAV system: (a) orthomosaic image of the entire field, (b) dense point cloud of a section o
f
the field, and (c) plant height map of the same area as in (b).
396 T
RANSACTIONS OF THE
ASABE
was assumed that the soil elevation was uniform within a
single plot. The dimension of each plot in the north-south
direction was determined based on the number of yield data
points in each row (~150 yield data points in each row), and
the dimension in the east-west direction was selected to in-
clude a sufficient number of rows to increase the potential of
bare soil being visible within the plot. The elevation infor-
mation for each pixel in the plots was extracted from the
DEM map, and the mean value of the lowest 1% of the ele-
vation information for each plot was used as the average el-
evation of the soil surface in that plot.
Although each plot included about nine cotton rows, as
shown in figure 2, bare soil was not visible in the plots with
full canopy closure. To identify the closed-canopy plots, the
RGB orthomosaic image was converted to the CIE 1976
L*a*b* color space, where channel a* represents the green-
red color component, with negative values corresponding to
green pixels and positive values to the background (Liu et
al., 2017). Therefore, channel a* was used to discriminate
crop and soil in the orthomosaic image. The canopy cover
ratio was calculated using the ratio of the number of crop
canopy pixels to the total number of pixels in each plot. In
this study, plots with a canopy cover ratio of 0.95 or higher
were considered to have a fully closed canopy, where bare
soil could not be seen. To calculate the plant height in the
plots without a visible soil surface, a linear interpolation of
the soil surface elevation between adjacent plots was used to
calculate the approximate elevation of the soil surface. As
shown in figure 2, some portion of the soil in plots 1, 2, and
6 (canopy cover ratio < 0.95) was visible, while the soil in
plots 3, 4, and 5 (canopy cover ratio > 0.95) was not visible.
Linear interpolation between the soil surface elevations of
plots 2 and 6 was used to estimate the soil surface elevation
of plots 3, 4, and 5. A plant height map was then generated
for each plot by subtracting the plot-mean soil elevation
from the plant canopy elevation of every pixel. A section of
a plant height map is shown in figure 1c.
G
EO
-R
EGISTRATION AND
R
OW
S
EPARATION
Because neither the UAV nor the yield monitor used in
this study employed a RTK-GPS system, their horizontal po-
sitioning accuracies were 1 to 2 m (about one to two cotton
rows) and sub-meter (about half to one cotton row), respec-
tively. Therefore, the geo-referenced data for UAV-based
plant height and for yield were not properly aligned (regis-
tered) in some rows when only GPS coordinate information
was used. In this study, an image feature-based geo-registra-
tion method was developed to improve the registration accu-
racy. The major steps included: (1) obtaining the GPS coor-
dinates of the GRPs using Google Earth (Mohammed et al.,
2013; Benker et al., 2011) and identifying the coordinates of
the plant height map, (2) converting the GPS coordinates of
yield data to the image coordinates of the plant height map,
(3) segmenting individual crop rows, and (4) adjusting the
image coordinates of yield data to the corresponding row po-
sition in the plant height map.
The positions of five GRPs located in the four corners and
middle of the field were identified in the plant height map,
and their GPS coordinates were matched to the image coor-
dinates of the plant height map. To convert the GPS coordi-
nates to image coordinates, a conversion scale was defined
and calculated using equation 1:
1 2 5
gk gm
k
ik im
gk gm
k
ik im
xx
xxx,m,k , , , ,m k
yy
yyy

(1)
where (x
gk
, y
gk
) and (x
ik
, y
ik
) are the GPS coordinates and im-
age coordinates of the kth GRP, respectively; and (x
gm
, y
gm
)
and (x
im
, y
im
) are the GPS coordinates and image coordinates
of the mth GRP (m k). Ten sets of conversion scales were
obtained, and their mean value was used as the conversion
scale to convert all yield GPS coordinate points to image co-
ordinate points with equation 2:


00
00
ii gi g k i
ii gi g k i
xxx xx
yyy yy


(2)
where (x
gi
, y
gi
) and (x
ii
, y
ii
) are the GPS coordinates and the
converted image coordinates of the ith yield datum, respec-
tively; ( ,
kk
x y
) is the average conversion scale between the
two coordinate systems; and (x
i0
, y
i0
) and (x
g0
, y
g0
) are the
image coordinates and GPS coordinates of a selected refer-
ence point.
To improve the registration accuracy, individual crop
rows were segmented from the plant height map and used to
align the corresponding rows of the yield data. The purpose
of row segmentation was to segment individual crop rows in
the images. The segmented rows were then directly regis-
tered to the yield monitor data, instead of relying only on the
GPS coordinates. Crop rows in the plant height map were
segmented based on the variation in elevation of the crop
rows and inter-rows. The elevation of crop rows was consist-
ently higher than that at inter-row locations, even with a
closed canopy (fig. 3). To search for inter-row locations,
strips with 50 pixels in the north-south (vertical) direction
and the same width as the whole plant height map (east-west
direction) were defined (shown as red boxes in fig. 3a). The
Figure 2. Section of the field showing plots used for DEM analysis. The red lines divided the field into 2550 small plots, each 2 m 8.5 m. Soil is
visible in plots 1, 2, and 6 but not in plots 3, 4, and 5. The canopy cover ratio for plots 1 to 6 was 0.67, 0.63, 1.00, 0.99, 0.97, and 0.91, respectively.
62(2): 393-403 397
sum of all plant height pixel values in the north-south direc-
tion in each strip was then calculated and plotted along the
east-west direction, as shown in figure 3b. The local valleys
in each curve in figure 3b, i.e., the lowest plant heights, rep-
resented the inter-row locations, which were used to segment
the crop rows, as shown in figure 3c. The coordinates of each
crop row in the plant height map were recorded for further
processing. Meanwhile, the yield data were also divided into
individual rows based on their GPS coordinates. The crop
rows and yield rows were aligned based on their order from
one edge of the field to the other. This created a one-to-one
correspondence between the image rows and the yield rows,
and the position of every yield GPS point (x
gi
, y
gi
) calculated
from equation 2 was adjusted to match the corresponding
image row position.
The geo-referenced yield data were recorded at 1 s inter-
vals during harvesting, resulting in about 25,000 data points
with each representing the cotton weight in an area of about
2 m
2
. However, data from the plant height map had a spatial
resolution of 2.36 10
-3
m
2
pixel
-1
, which resulted in 7559
3880 data points (pixels). To match the two data sets, a small
region of interest (ROI) with 30 30 pixels (equivalent to
about 2.3 m
2
) centered on the coordinate of each yield data
point in the plant height map was selected. The average crop
height in each ROI was extracted from the plant height map
and matched with the yield data for further analysis. During
harvesting, the yield monitor measured the yield separately
for each of the four rows harvested in one pass (harvest
pass), and the software applied offsets from each row to the
GPS receiver to determine the locations of the individual
rows. However, the yield monitor was calibrated using the
total yield of all four rows, and separate calibration for each
row was not possible. Therefore, the data from the 38 four-
row harvest passes were believed to be more accurate than
individual-row data and were used in the subsequent analy-
sis.
D
ATA
A
NALYSIS
The accuracy of the image-based height measurement
was evaluated by comparing the height extracted from the
DEM with that manually measured at the 12 GRPs using a
paired t-test. The Pearson correlation coefficient was calcu-
lated to evaluate the correlation between image-based plant
height and yield. Both processes were conducted in R soft-
ware (RStudio Desktop 1.1.453, RStudio, Boston, Mass.). A
yield prediction model using plant height was developed us-
ing linear regression. The yield estimation model was writ-
ten as equation 3 (Härdle, 1990):

ii
ˆ
ymX
(3)
where
i
ˆ
y is the estimated yield (kg ha
-1
), X
i
is a vector of the
image-based plant height, is the error (or noise) of the data,
and m(X
i
) is the model to be developed. A bin smoothing
method was used to smooth the raw data

1
,
n
ii
i
X y
, where
n is the number of raw data points (Irizarry, 2001), to reduce
the fluctuations and measurement error from the UAV sen-
sors and yield monitor so that the main relationship was eas-
ier to obtain. Prior to modeling, the plant height data corre-
sponding to yield data

1
,
n
ii
i
X y
were sorted in ascending
order based on the values of X
i
, and a bin with a length of
15 data points was used to smooth the raw data

1
,
n
ii
i
X y
.
The mean

, Xy
of each bin was obtained, and all the

1
,
n_bin
ii
i
Xy
made up a new set of data, i.e., smoothed data,
where n_bin is the number of bins. The relationship m(X
i
)
between yield and plant height was fit based on the smoothed
data using a least squares regression model (Hastie et al.,
2001). Residuals between the measured yield obtained from
the yield monitor and the predicted yield obtained from the
regression model were calculated using equation 4:
Residuals
ii
ˆ
yy (4)
where y
i
is the yield (kg ha
-1
) from the yield monitor, and
i
ˆ
y
is the predicted yield from equation 3.
Yield data less than 150 kg ha
-1
were removed from the
data set because those data only occurred at the edges of the
field or in non-cropped areas, where the measured yield
should be zero. Model performance was assessed using root
mean square error (RMSE) and mean absolute error (MAE),
as shown in equations 5 and 6 (Willmott and Matsuura,
2005):

2
1
RMSE
n
ii
i
ˆ
yy
n
(5)

1
MAE
n
ii
i
ˆ
yy
n
(6)
where y
i
and
i
ˆ
y
are the ith observed yield data and predicted
yield, respectively, and n is the total number of observations
Figure 3. Segmentation of crop rows: (a) section of the plant height map showing the field divided into east-west strips perpendicular to the crop
rows, (b) accumulated value of each north-south pixel column in each strip, and (c) plant height map with segmented crop rows.
398 T
RANSACTIONS OF THE
ASABE
in the data set. Although both error statistics evaluate the dif-
ference between predicted yield and measured yield, RMSE
imposes a higher penalty than MAE for predictions that are
far from the measured value (Willmott and Matsuura, 2005).
A 10-fold cross-validation was used to evaluate the pre-
diction performance of the model. The raw data

1
,
n
ii
i
Xy
were randomly partitioned into ten equal-size subsamples. In
each test, a single subsample was retained for testing the pre-
diction ability of the model, and the remaining nine subsam-
ples were used as training data. This process was repeated
ten times, with each of the ten subsamples used exactly once
as the test data. Bin smoothing was conducted on the training
data to obtain a smoothed data set

1
,
n_bin
ii
i
Xy
, and then
a linear model was built based on the smoothed data. The
test data were used to determine the model’s RMSE and
MAE.
R
ESULTS
A
CCURACY OF
DEM-D
ERIVED
H
EIGHT
The image-based height measurements of the twelve
GRPs and the manual measurements are shown in table 2. A
paired t-test indicated that there was no significant difference
in height due to measurement methods. The measurement
errors were in the range of -0.1 to 0.16 m with an average
value of 0.07 m, equivalent to a 6% error relative to the man-
ual measurement. This indicated that the UAV-based imag-
ing measurement method had the potential to quantify plant
height with less than 10% error.
C
ORRELATION
A
NALYSIS
Figure 4 shows the Pearson correlation coefficients (r) be-
tween cotton yield and plant height with and without row
segmentation for the 38 four-row harvest passes. The red
curve shows the correlation coefficient of each pass calcu-
lated based on GPS alignment, while the values on the black
curve were calculated based on row separation. The overall
correlation coefficient was improved from 0.81 to 0.83 after
row separation was added. Although the change in the over-
all correlation was not large, the correlation coefficients of
the rows with low correlation increased substantially from
about 0.7 to 0.8 (blue circles in fig. 4). The GPS coordinates
in the rows with the lowest correlation coefficients had
larger offsets than other rows, and aligning the yield data
more accurately could improve the correlation between plant
height and yield. The results indicate that the geo-registra-
tion based on row separation made an improvement in cor-
relation in these cases.
Figure 4 also shows that the Pearson correlation coeffi-
cients fluctuated over a large range from pass to pass. To
identify possible reasons for the low correlation in some
passes, a detailed analysis was conducted on two passes
(passes 31 and 38), which had the lowest correlation coeffi-
cients using the row separation method (fig. 4). Figure 5a
shows the yield and plant height in pass 31 (r = 0.66), where
points 135 to 145 and points 157 to 162 were in regions of
bare soil, and the yield and plant height were both zero.
Points 1 to 20 in pass 31 and points 10 to 40 in pass 38 had
low yield but taller plants. These points are in the lowest por-
tion of the field near the drain. It is likely that good moisture
conditions early in the season favored taller plants in this
area, but waterlogging later in the season caused boll shed-
ding, resulting in lower yield. The locations of pass 31 and
points 60 to 80 in the pass were identified on the orthomosaic
Table 2. Height measured from DEM and manually.
Ground
Reference
Point
Height
from DEM
(m)
Manual
Measurement
(m)
Absolute Error
(m) (%)
1 0.94 1.04 0.10 11
2 1.10 1.00 0.10 10
3 1.02 1.02 0.00 0
4 1.20 1.04 0.16 13
5 1.16 1.09 0.07 6
6 0.98 1.04 0.06 6
7 1.10 1.07 0.03 3
8 1.09 1.04 0.05 5
9 0.99 1.02 0.03 3
10 1.14 1.02 0.12 11
11 1.13 1.07 0.06 5
12 1.07 1.07 0.00 0
Average 1.08 1.04 0.07 6
Figure 4. Pearson coefficients of linear correlation between yield and plant height in all four-row harvest passes. The numbers on the dashed lines
show the overall correlation coefficients. The red line with circle markers represents the correlation with only GPS alignment, while the blac
k
line with triangle markers represents the correlation using GPS alignment plus position adjustment based on row separation. The blue circles
mark the passes where the correlation improved substantially after applying row separation.
62(2): 393-403 399
image, as shown in figure 6. The points coincide with a rain-
fed portion of the field, and water stress likely decreased the
cotton yield, reducing the correlation between plant height
and harvested yield. However, other rainfed areas had higher
correlations, so the soil texture probably exacerbated the
drought stress response in this area of pass 31.
Rainfall was adequate for early-season growth and devel-
opment of the cotton crop; however, only 2 mm of rain was
recorded between July 9 and August 5. The first irrigation
was applied to portions of the field on July 19, and up to five
irrigations were applied to parts of the field by the time
25 mm of rain was recorded on August 6. No additional irri-
gations were applied.
Y
IELD
E
STIMATION
There were over 25,000 data points in 152 crop rows of
harvested yield data, which was reduced to about 6500 data
points after combining the rows into 38 four-row harvest
passes and removing outliers. The data in each pass were
then smoothed with a 15-point bin smoothing method, re-
sulting in about 450 data points. A 10-fold cross-validation
was conducted, resulting in average errors of RMSE =
Figure 6. Explanation of low correlation between yield and plant height in pass 31 (rows 120 to 124). The red lines in (a) and (c) mark the four
rows of pass 31, and the blue numbers relate to data points in figure 5a. The graph (b) shows the cumulative precipitation from the date o
f
planting. Image data were collected on August 12, 2017, and the crop was harvested in early October (marked with dashed lines in the graph).
Figure 5. Details of cotton yield and plant height in two harvest passes with low Pearson coefficients (after row separation). Blue circles mark the
data points that have inconsistent trends between yield and plant height: (a) pass 31 (r = 0.66) and (b) pass 38 (r = 0.67).
400 T
RANSACTIONS OF THE
ASABE
550 kg ha
-1
(RMSE/mean = 21%) and MAE = 420 kg ha
-1
(MAE/mean = 16%). Figure 7 shows a scatter plot of cotton
yield versus plant height for all the smoothed and un-
smoothed data, in which the blue line shows a linear model
based on the smoothed data, and p-values < 0.01 were ob-
tained for both the plant height coefficient and intercept of
the linear model. Errors obtained for the full data set were
RMSE = 630 kg ha
-1
and MAE = 450 kg ha
-1
.
While the regression model established the general rela-
tionship of crop height and yield, it was instructive to con-
sider the points below the regression line, where the har-
vested yield was much lower than that estimated from plant
height (the unsmoothed data below the green dashed line in
fig. 7). To explore the causes of low yield and high negative
residuals, figure 8 shows the regions with low yield and the
regions with high negative residuals in the field. Figure 8a is
a residual plot of the unsmoothed data in which the red line
separates the points with negative residuals greater than
1000 kg ha
-1
. The high-residual points are plotted as red dots
in figure 8b, which shows that some of these points are in
rainfed regions and some are located adjacent to areas of yel-
low points that had yield less than 1500 kg ha
-1
. The yellow
points were in areas of lowest EC
a
, corresponding to areas
with a high-profile sand content (Sudduth et al., 2003) and
therefore a low soil water holding capacity. This resulted in
short plants and low yield, although the negative residuals
Figure 7. Scatter plot of seed cotton yield versus plant height. The linear function and R
2
value were obtained based on the smoothed data. The
points below the green dashed line are those with negative residuals greater than 1000 kg ha
-1
.
Figure 8. Unsmoothed data points with high negative residuals are below the red line in (a) and are shown as red dots in (b). The yellow points in
(b) have yield less than 1500 kg ha
-1
. “R” marks rainfed regions with no irrigation. The white circles mark regions of lowest apparent soil electrical
conductivity (EC
a
). The region within the purple circle has relatively low EC
a
compared to other parts of the field.
62(2): 393-403 401
were not greater than 1000 kg ha-1 because the plant height
and yield were both small. The red points located on the
boundaries of the sandy soil regions are where the water
shortage was not as severe or persistent, resulting in reduc-
tions in yield with less reduction in plant height and creating
highly negative residuals. Meanwhile, the red points in the
purple circle were in an area with low soil ECa, where the
coarse soil texture (i.e., high sand content) led to water
stress, even with irrigation, and contributed to low yield and
high negative residuals.
DISCUSSION
The primary goal of this study was to evaluate the perfor-
mance of using plant height acquired with UAV-based im-
aging data for cotton yield estimation. Plant height is an im-
portant index for plant growth status, biomass (Brocks and
Bareth, 2018), and yield (Sharma et al., 2016). The plant
height measurement accuracy obtained in this study
(RMSE = 0.08 m) is comparable to that of other studies us-
ing similar methods. For example, RMSE of 0.08 to 0.09 m
was obtained by Brocks and Bareth (2018), and RMSE of
0.05 to 0.1 m was obtained by Chang et al. (2017). With
RMSE values less than 0.1 m, UAV-based height measure-
ment has potential to replace manual measurement. Bendig
et al. (2015) showed that plant height was more accurate than
spectral indices for biomass estimation, although the combi-
nation of plant height and spectral indices was better than
either alone. Meanwhile, Ballesteros et al. (2018) found that
plant height had high correlation with spectral indices and
canopy cover. Therefore, improved methods of plant height
acquisition could provide another approach for cotton yield
estimation.
A unique aspect of this study was that UAV-based sens-
ing data (height, potentially other sensing data, such as veg-
etation indices, and temperature) were registered with
ground data (yield, irrigation, soil, and fertilizer) row-by-
row, which provided high-resolution site-specific infor-
mation. Other studies have correlated yield and remote sens-
ing image data at a lower resolution (e.g., plot level) than
this study. For example, Huang et al. (2016) analyzed the
correlation between yield and UAV-based vegetation indi-
ces from 20 plots where one yield point was acquired from
each plot. Van Der Meij et al. (2017) divided a field into
42 equal plots, and one yield point was acquired from each
plot. Similarly, Vega et al. (2015) used one data point from
each of 32 plots. In contrast, this study connected image data
and geo-referenced ground data using row separation at a
scale of 30 30 pixels, resulting in about 7000 data points.
Although the geo-registration might be improved using a
more accurate RTK-GPS unit, the row separation method
provides a low-cost alternative. The row-based geo-registra-
tion method provided flexibility in estimating yield for one
row or for the whole field. Meanwhile, this method makes it
easy to interpret and explore the data in each row, as shown
in figures 4 and 5.
A bin smoothing method was used to smooth the plant
height and corresponding yield data, which increased the
correlation between plant height and cotton yield. Figure 7
shows that the plant height and yield of cotton had a linear
relationship with R2 = 0.96 for the smoothed data. This result
was different from the results reported by Huang et al.
(2016), where cotton in the middle of the height range had
higher yield compared to cotton with low and high plant
heights (R2 = 0.43). The main reason for the different results
from the two studies may be that the height of the cotton in
this study was within the preferred range, without excessive
growth. Excessive height in cotton (i.e., rank growth) can be
detrimental but is usually adequately controlled with plant
growth regulators. Possible reasons for low yield were ana-
lyzed, and water stress and coarse soil texture, as indicated
by low soil ECa, might be the major reasons for the areas of
low yield. This result was consistent with the findings re-
ported by Vories et al. (2015).
CONCLUSION
A method was established for plant height extraction and
seed cotton yield estimation based on plant height data from
UAV-based remote sensing. A UAV equipped with a low-
cost RGB camera was used to collect images of cotton dur-
ing the flowering period. A DEM was generated from the
RGB images based on the SfM method. Plant height was ex-
tracted from the DEM by calculating the difference between
the canopy height and soil surface. Linear interpolation was
used to estimate the soil elevation in regions where the soil
was invisible due to canopy closure. A geo-registration
method based on row separation was developed to improve
the accuracy of geo-referencing the ground and UAV-based
sensing data. When the row separation method was included
in the geo-registration process, the alignment result was bet-
ter and the correlation between yield and plant height in-
creased from 81% to 83%, showing the effectiveness of row
separation. A bin smoothing method was used to smooth the
plant height corresponding with yield data. A linear model
showed that plant height and cotton yield had a positive re-
lationship. Yield estimation errors of RMSE = 550 kg ha-1
and MAE = 420 kg ha-1 were obtained using 10-fold cross-
validation. From the above results, we conclude that a UAV
equipped with a low-cost RGB camera has potential to map
plant height for cotton yield estimation. Data points with low
yield and high negative residuals (i.e., tall plants with low
yield) from the regression estimate were examined, and it
was found that water stress and low soil water holding ca-
pacity, as indicated by low soil ECa, might be the main con-
tributing factors.
ACKNOWLEDGEMENTS
The authors would like to thank the University of Mis-
souri Agricultural Experiment Station for partly funding this
project. We also thank colleagues Jing Zhou, Chin Nee
Vong, and Dr. Xiuqing Fu for their help in data collection,
data analysis, and improvement of the manuscript.
REFERENCES
Ballesteros, R., Ortega, J. F., Hernandez, D., & Moreno, M. A.
(2018). Onion biomass monitoring using UAV-based RGB
402 TRANSACTIONS OF THE ASABE
imaging. Prec. Agric., 19(5), 840-857.
https://doi.org/10.1007/s11119-018-9560-y
Bendig, J., Bolten, A., Bennertz, S., Broscheit, J., Eichfuss, S., &
Bareth, G. (2014). Estimating biomass of barley using crop
surface models (CSMs) derived from UAV-based RGB
imaging. Remote Sensing, 6(11), 10395-10412.
https://doi.org/10.3390/rs61110395
Bendig, J., Yu, K., Aasen, H., Bolten, A., Bennertz, S., Broscheit,
J., ... Bareth, G. (2015). Combining UAV-based plant height
from crop surface models, visible, and near-infrared vegetation
indices for biomass monitoring in barley. Intl. J. Appl. Earth
Obs. Geoinfo., 39, 79-87.
https://doi.org/10.1016/j.jag.2015.02.012
Benker, S. C., Langford, R. P., & Pavlis, T. L. (2011). Positional
accuracy of the Google Earth terrain model derived from
stratigraphic unconformities in the Big Bend region, Texas,
USA. Geocarto Intl., 26(4), 291-303.
https://doi.org/10.1080/10106049.2011.568125
Brocks, S., & Bareth, G. (2018). Estimating barley biomass with
crop surface models from oblique RGB imagery. Remote
Sensing, 10(2), 268. https://doi.org/10.3390/rs10020268
Chang, A., Jung, J., Maeda, M. M., & Landivar, J. (2017). Crop
height monitoring with digital imagery from unmanned aerial
system (UAS). Comput. Electron. Agric., 141, 232-237.
https://doi.org/10.1016/j.compag.2017.07.008
Chu, T., Chen, R., Landivar, J., Maeda, M. M., Yang, C., & Starek,
M. J. (2016). Cotton growth modeling and assessment using
unmanned aircraft system visual-band imagery. J. Appl. Remote
Sensing, 10, 036018. https://doi.org/10.1117/1.JRS.10.036018
De Castro, A. I., Torres-Sanchez, J., Pena, J. M., Jimenez-Brenes, F.
M., Csillik, O., & Lopez-Granados, F. (2018). An automatic
random forest-OBIA algorithm for early weed mapping between
and within crop rows using UAV imagery. Remote Sensing,
10(2), 285. https://doi.org/10.3390/rs10020285
Du, M., & Noguchi, N. (2017). Monitoring of wheat growth status
and mapping of wheat yield’s within-field spatial variations
using color images acquired from UAV-camera system. Remote
Sensing, 9(3), 289. https://doi.org/10.3390/rs9030289
Duan, T., Zheng, B., Guo, W., Ninomiya, S., Guo, Y., & Chapman,
S. C. (2017). Comparison of ground cover estimates from
experiment plots in cotton, sorghum, and sugarcane based on
images and ortho-mosaics captured by UAV. Funct. Plant Biol.,
44(1), 169-183. https://doi.org/10.1071/FP16123
Glendell, M., McShane, G., Farrow, L., James, M. R., Quinton, J.,
Anderson, K., ... Brazier, R. E. (2017). Testing the utility of
structure-from-motion photogrammetry reconstructions using
small unmanned aerial vehicles and ground photography to
estimate the extent of upland soil erosion. Earth Surf. Proc.
Landforms, 42(12), 1860-1871. https://doi.org/10.1002/esp.4142
Härdle, W. (1990). Applied nonparametric regression. Econometric
Society Monograph No. 19. Cambridge, UK: Cambridge
University Press. https://doi.org/10.1017/CCOL0521382483
Hastie, T., Tibshirani, R., & Friedman, J. (2001). The elements of
statistical learning: Data mining, inference, and prediction.
New York, NY: Springer Science+Business Media.
Huang, Y., Brand, H. J., Sui, R., Thomson, S. J., Furukawa, T., &
Wayne Ebelhar, M. (2016). Cotton yield estimation using very
high-resolution digital images acquired with a low-cost small
unmanned aerial vehicle. Trans. ASABE, 59(6), 1563-1574.
https://doi.org/10.13031/trans.59.11831
Huang, Y., Sui, R., Thomson, S. J., & Fisher, D. K. (2013).
Estimation of cotton yield with varied irrigation and nitrogen
treatments using aerial multispectral imagery. Intl. J. Agric. Biol.
Eng., 6(2), 37-41.
Irizarry, R. A. (2001). Applied nonparametric and modern statistics.
Baltimore, MD: Johns Hopkins University.
Jay, S., Rabatel, G., Hadoux, X., Moura, D., & Gorretta, N. (2015).
In-field crop row phenotyping from 3D modeling performed
using structure from motion. Comput. Electron. Agric., 110, 70-
77. https://doi.org/10.1016/j.compag.2014.09.021
Jiao, C., Chen, C., McGarvey, R. G., Bohlman, S., Jiao, L., & Zare,
A. (2018). Multiple instance hybrid estimator for hyperspectral
target characterization and sub-pixel target detection. ISPRS J.
Photogram. Remote Sensing, 146, 235-250.
https://doi.org/10.1016/j.isprsjprs.2018.08.012
Komm, B., & Moyer, M. (2015). Vineyard yield estimation.
EM086E. Pullman, WA: Washington State University
Extension.
Liu, Q.-S., Li, X.-Y., Liu, G.-H., Huang, C., & Guo, Y.-S. (2016).
Cotton area and yield estimation at Zhanhua County of China
using HJ-1 EVI time series. Proc. 3rd Annual Conf. Info. Tech.
and Applications (ITA 2016), article 09001.
https://doi.org/10.1051/itmconf/20160709001
Liu, S., Baret, F., Andrieu, B., Burger, P., & Hemmerle, M. (2017).
Estimation of wheat plant density at early stages using high-
resolution imagery. Front. Plant Sci., 8, 739.
https://doi.org/10.3389/fpls.2017.00739
Maimaitijiang, M., Ghulam, A., Sidike, P., Hartling, S.,
Maimaitiyiming, M., Peterson, K., ... Fritschi, F. (2017).
Unmanned aerial system (UAS)-based phenotyping of soybean
using multi-sensor data fusion and extreme learning machine.
ISPRS J. Photogram. Remote Sensing, 134, 43-58.
https://doi.org/10.1016/j.isprsjprs.2017.10.011
Malambo, L., Popescu, S. C., Murray, S. C., Putman, E., Pugh, N.
A., Horne, D. W., ... Bishop, M. (2018). Multitemporal field-
based plant height estimation using 3D point clouds generated
from small unmanned aerial systems high-resolution imagery.
Intl. J. Appl. Earth Obs. Geoinfo., 64, 31-42.
https://doi.org/10.1016/j.jag.2017.08.014
Mohammed, N. Z., Ghazi, A., & Mustafa, H. E. (2013). Positional
accuracy testing of Google Earth. Intl. J. Multidisc. Sci. Eng.,
4(6), 6-9.
Sankaran, S., Khot, L. R., Espinoza, C. Z., Jarolmasjed, S.,
Sathuvalli, V. R., Vandemark, G. J., ... Pavek, M. J. (2015).
Low-altitude, high-resolution aerial imaging systems for row
and field crop phenotyping: A review. European J. Agron., 70,
112-123. https://doi.org/10.1016/j.eja.2015.07.004
Sankey, T. T., McVay, J., Swetnam, T. L., McClaran, M. P.,
Heilman, P., & Nichols, M. (2018). UAV hyperspectral and lidar
data and their fusion for arid and semi-arid land vegetation
monitoring. Remote Sensing Ecol. Cons., 4(1), 20-33.
https://doi.org/10.1002/rse2.44
Scotford, I. M., & Miller, P. C. (2004). Combination of spectral
reflectance and ultrasonic sensing to monitor the growth of
winter wheat. Biosyst. Eng., 87(1), 27-38.
https://doi.org/10.1016/j.biosystemseng.2003.09.009
Sharma, L. K., Bu, H., Franzen, D. W., & Denton, A. (2016). Use
of corn height measured with an acoustic sensor improves yield
estimation with ground-based active optical sensors. Comput.
Electron. Agric., 124, 254-262.
https://doi.org/10.1016/j.compag.2016.04.016
Smith, M. W., Carrivick, J. L., & Quincey, D. J. (2016). Structure
from motion photogrammetry in physical geography. Prog.
Phys. Geogr., 40(2), 247-275.
https://doi.org/10.1177/0309133315615805
Sudduth, K. A., Drummond, S. T., & Myers, D. B. (2012). Yield
Editor 2.0: Software for automated removal of yield map errors.
ASABE Paper No. 121338243. St. Joseph, MI: ASABE.
Sudduth, K. A., Kitchen, N. R., Bollero, G. A., Bullock, D. G., &
Wiebold, W. J. (2003). Comparison of electromagnetic
induction and direct sensing of soil electrical conductivity.
Agron. J., 95(3), 472-482.
62(2): 393-403 403
https://doi.org/10.2134/agronj2003.4720
Sui, R., Thomasson, J. A., & Ge, Y. (2012). Development of sensor
systems for precision agriculture in cotton. Intl. J. Agric. Biol.
Eng., 5(4), 1-14.
Van Der Meij, B., Kooistra, L., Suomalainen, J., Barel, J. M., & De
Deyn, G. B. (2017). Remote sensing of plant trait responses to
field-based plant-soil feedback using UAV-based optical
sensors. Biogeosciences, 14(3), 733-749.
https://doi.org/10.5194/bg-14-733-2017
Vega, F. A., Ramirez, F. C., Saiz, M. P., & Rosua, F. O. (2015).
Multi-temporal imaging using an unmanned aerial vehicle for
monitoring a sunflower crop. Biosyst. Eng., 132, 19-27.
https://doi.org/10.1016/j.biosystemseng.2015.01.008
Vories, E. D., Stevens, W. E., Sudduth, K. A., Drummond, S. T., &
Benson, N. R. (2015). Impact of soil variability on irrigated and
rainfed cotton. J. Cotton Sci., 19(1), 1-14.
Walter, J., Edwards, J., McDonald, G., & Kuchel, H. (2018).
Photogrammetry for the estimation of wheat biomass and
harvest index. Field Crops Res., 216, 165-174.
https://doi.org/10.1016/j.fcr.2017.11.024
Wang, D., Xin, X., Shao, Q., Brolly, M., Zhu, Z., & Chen, J. (2017).
Modeling aboveground biomass in Hulunber grassland
ecosystem by using unmanned aerial vehicle discrete lidar.
Sensors, 17(1), 180. https://doi.org/10.3390/s17010180
Westoby, M. J., Brasington, J., Glasser, N. F., Hambrey, M. J., &
Reynolds, J. M. (2012). ‘Structure-from-motion’
photogrammetry: A low-cost, effective tool for geoscience
applications. Geomorphology, 179, 300-314.
https://doi.org/10.1016/j.geomorph.2012.08.021
Willmott, C. J., & Matsuura, K. (2005). Advantages of the mean
absolute error (MAE) over the root mean square error (RMSE)
in assessing average model performance. Climate Res., 30(1),
79-82. https://doi.org/10.3354/cr030079
Yang, G., Liu, J., Zhao, C., Li, Z., Huang, Y., Yu, H., ... Yang, H.
(2017). Unmanned aerial vehicle remote sensing for field-based
crop phenotyping: Current status and perspectives. Front. Plant
Sci., 8(1111). https://doi.org/10.3389/fpls.2017.01111
Zhang, L., & Grift, T. E. (2012). A LIDAR-based crop height
measurement system for Miscanthus giganteus. Comput.
Electron. Agric., 85, 70-76.
https://doi.org/10.1016/j.compag.2012.04.001
Zhou, J., Khot, L. R., Boydston, R. A., Miklas, P. N., & Porter, L.
(2018). Low-altitude remote sensing technologies for crop stress
monitoring: A case study on spatial and temporal monitoring of
irrigated pinto bean. Prec. Agric., 19(3), 555-569.
https://doi.org/10.1007/s11119-017-9539-0
... Simple linear regression models have also been developed to predict cotton yield using solely satellite-derived vegetation indices as independent variables [13][14][15][16], unmanned aerial vehicle (UAV) [17][18][19][20], airplane [21], and ground sensor [22]. While vegetation indices alone have limitations in capturing the complex relationships among production, plant physiology, climate, soil nutrition, soil water parameters, pest and disease infestation, and crop management characteristics [10], several studies have demonstrated their ability to predict crop yield with reasonable accuracy [7,23,24]. ...
... Table 4 provides a summary of the optimal periods, expressed in DAS, as inferred from the literature cited in Table 1. The most favorable correlations between cotton yield and the VIs have been reported across various stages, ranging from the flowering and boll development period [12,18,22,[26][27][28][29][30][31] to boll opening, maturation, and defoliant application [10,[15][16][17]19,21,27,30]. Notably, these findings align with the results obtained in the present study. ...
Article
Full-text available
Satellite remote sensing data expedite crop yield estimation, offering valuable insights for farmers’ decision making. Recent forecasting methods, particularly those utilizing machine learning algorithms like Random Forest and Artificial Neural Networks, show promise. However, challenges such as validation performances, large volume of data, and the inherent complexity and inexplicability of these models hinder their widespread adoption. This paper presents a simpler approach, employing linear regression models fitted from vegetation indices (VIs) extracted from MODIS sensor data on the Terra and Aqua satellites. The aim is to forecast cotton yields in key areas of the Brazilian Cerrado. Using data from 281 commercial production plots, models were trained (167 plots) and tested (114 plots), relating seed cotton yield to nine commonly used VIs averaged over 15-day intervals. Among the evaluated VIs, Enhanced Vegetation Index (EVI) and Triangular Vegetation Index (TVI) exhibited the lowest root mean square errors (RMSE) and the highest determination coefficients (R2). Optimal periods for in-season yield prediction fell between 90 and 105 to 135 and 150 days after sowing (DAS), corresponding to key phenological phases such as boll development, open boll, and fiber maturation, with the lowest RMSE of about 750 kg ha−1 and R2 of 0.70. The best forecasts for early crop stages were provided by models at the peaks (maximum value of the VI time series) for EVI and TVI, which occurred around 80–90 DAS. The proposed approach makes the yield predictability more inferable along the crop time series just by providing sowing dates, contour maps, and their respective VIs.
... Simple linear regression models have been developed to predict cotton yield using only satellitederived vegetation indices as independent variables [13][14][15][16], unmanned aerial vehicle (UAV) [17][18][19][20], airplane [21] and ground sensor [22]. If on the one hand VIs by themselves have shown some limitations for describing the complex relationships among production, plant physiology, climate, soil nutrition, soil water parameters, pest and disease infestation, crop design and management characteristics [10], on the other several studies have indicated that crop yield can be predicted with relatively good accuracies with this approach [7,23,24]. ...
... Table 4 summarizes the best periods, in DAS, inferred from the papers listed in Table 1. The best correlations between cotton yield and the VIs varied from the flowering and boll development period [12,18,22,37,38,39] to boll opening, maturation and defoliant application [10,15,16,17,19,21,40,41], with the latter in agreement with the results obtained in the present study. Particularly, the study carried out by Lang et al. [10] in 355 plots of cotton cultivated from 2012 to 2019 in Xinjiang Province, China, showed very similar results to ours, with lowest RMSE and highest R 2 for the fourth and fifth months after sowing (90-120 to 120-150 DAS). ...
Preprint
Full-text available
Crop yield estimation supported by satellite remote sensing data can provide expeditious and strategic information for farmers’ decision-making. Most recent forecasting methods have indicated a promising pathway based on machine learning algorithms. However, validation performances, demand for big data and their inherent inexplicability have not yet consolidated a substantial differential to replace methods based on simpler and more understandable models. This paper proposes an approach based on simple linear models fitted from vegetation indices (VIs) assessed at regular intervals of time series derived from MODIS satellite images, aiming to forecast cotton yields in representative areas of the Brazilian Cerrado. Data from 281 commercial production plots were taken to train (167 plots) and test (114 plots) linear regression models relating seed cotton yield and nine well-known VIs averaged in 15-days intervals. Among the evaluated VIs, EVI (Enhanced Vegetation Index) and TVI (Triangular Vegetation Index) showed the lowest root mean square errors (RMSE) and the highest determination coefficients. The best periods for in-season yield prediction were from the 90-105 to 135-150 days after sowing (DAS), i.e. phenological phases corresponding to boll development, open boll and fiber maturation, with lowest RMSE of about 750 kg ha-1 and R2=0.70. The best forecasts for early crop stages were provided by models at the peaks (maximum value of the VI time-series) for EVI and TVI, which occurred around 80-90 DAS. The proposed approach makes the yield predictability more inferable along the crop time series just by providing sowing dates, contour maps and its respective VIs.
... The height of the sugarcane stalk is a crucial factor for the production estimation of sugarcane farming. Several studies have been published to estimate the crop height from UAV images [20][21][22][23][24][25][26]. Generally, this was carried out through the difference between CSM and the GSM of the crop fields generated from UAV images. ...
... The study of Zhang et al. [24] also proved that the maize height could be precisely estimated through the difference between CSM and GSM generated from UAV images. A similar study was also carried out by Feng et al. [25], who indicated that the extracted cotton height from UAV images as an important input variable for the cotton production model could help to significantly improve the accuracy of the model. In our study, we also would like to use this approach to extract the sugarcane stalk height from our UAV stereo images for sugarcane yield estimation. ...
Article
Full-text available
The precise estimation of sugarcane yield at the field scale is urgently required for harvest planning and policy-oriented management. Sugarcane yield estimation from satellite remote sensing is available, but satellite image acquisition is affected by adverse weather conditions, which limits the applicability at the field scale. Secondly, existing approaches from remote sensing data using vegetation parameters such as NDVI (Normalized Difference Vegetation Index) and LAI (Leaf Area Index) have several limitations. In the case of sugarcane, crop yield is actually the weight of crop stalks in a unit of acreage. However, NDVI’s over-saturation during the vigorous growth period of crops results in significant limitations for sugarcane yield estimation using NDVI. A new sugarcane yield estimation is explored in this paper, which employs allometric variables indicating stalk magnitude (especially stalk height and density) rather than vegetation parameters indicating the leaf quantity of the crop. In this paper, UAV images with RGB bands were processed to create mosaic images of sugarcane fields and estimate allometric variables. Allometric equations were established using field sampling data to estimate sugarcane stalk height, diameter, and weight. Additionally, a planting density estimation model at the pixel scale of the plot was created using visible light vegetation indices from the UAV images and ground survey data. The optimal planting density estimation model was applied to estimate the number of plants at the pixel scale of the plot in this study. Then, the retrieved height, diameter, and density of sugarcane in the fields were combined with stalk weight data to create a model for estimating the sugarcane yield per plot. A separate dataset was used to validate the accuracy of the yield estimation. It was found that the approach presented in this study provided very accurate estimates of sugarcane yield. The average yield in the field was 93.83 Mg ha−1, slightly higher than the sampling yield. The root mean square error of the estimation was 6.25 Mg ha−1, which was 7.12% higher than the actual sampling yield. This study offers an alternative approach for precise sugarcane yield estimation at the field scale.
... Por outro lado, o aumento percentual foi de 34,8; 35,3 e 31,58 % nas áreas classificadas como baixo, médio e alto potencial, respectivamente. Esses resultados apontam no sentido de que nos níveis de menor potencial produtivo há um maior acréscimo proporcional no índice de vigor das plantas (NDVI) em razão do manejo vegetativo em taxa variada corroborando diversos autores (FENG et al., 2019;MARESMA et al., 2020;GARCÍA-MARTÍNEZ et al., 2020) Tsouros et al. (2019) apontam a possibilidade do uso de imagens de VANT para a identificação de falhas nas áreas agrícolas, como tamanho, localização e porcentagem de falhas, possibilitando sua correção. A utilização de VANTs e índices de vegetação são viáveis tanto econômica quanto agronomicamente (YEOM et al., 2019;HASSAN et al., 2019). ...
... However, counting cotton fruits by hand in the field is a laborious task that is usually limited to a few plants within a field or field plot (Ritchie et al., 2012). Traditional methods of yield prediction based on manual sampling (Huang et al., 2016) or visual inspection and experience (Feng et al., 2019) are prone to errors and are not practical for evaluating large numbers of plots commonly encountered in plant breeding programs. ...
Article
High-throughput phenotyping (HTP) has become a powerful tool for gaining insights into the genetic and environmental factors that affect cotton (Gossypium spp.) growth and yield. With the recent advances in the field of computer vision, namely the integration of deep learning algorithms, the accuracy and efficiency of HTP systems have improved dramatically, enabling them to automatically quantify such fundamental phenotypic traits as fruit identification and enumeration. However, there is currently no HTP system available for counting all the reproductive phases of cotton crop that can be deployed in agronomic field conditions throughout the growing season. This study presents CottonSense, an advanced HTP system that overcomes the challenges of deployment across multiple growth periods by effectively segmenting and enumerating cotton fruits at four stages of growth, including square, flower, closed boll, and open boll. Consequently, CottonSense enhances agronomic management through increased opportunities for data collection and analysis. Using RGB-D cameras, it captures and processes both two and three-dimensional data, facilitating a wider range of phenotypic trait extractions such as crop biomass and plant architecture. To segment the cotton fruits, a Mask-RCNN model is trained and optimized for faster inference using TensorRT. The model yields an average AP score of 79% in segmentation across the four fruit categories. Moreover, the model's accuracy in estimating total fruit count per image is validated by a strong agreement with the counts given by ten domain experts, as reflected by an R2 value of 0.94. Furthermore, to accurately count the segmented fruits over large populations of plants, an enumeration algorithm based on a tracking strategy is developed that achieves an R2 value of 0.93 when compared to hand-counted fruits in the field. The proposed HTP system, which is implemented entirely on an edge computing device, is cost-effective and power-efficient, making it an effective tool for high-yield cotton breeding and crop improvement. The code for CottonSense is publicly available at https://github.com/FeriBolour/CottonSense.
... Thus, in an attempt to come up with other innovative and efficient approaches to in-field data collection, others have come up with different methods/tools including the use of ultrasonic devices as well as the unmanned aerial vehicle (UAV) for such purposes (Turner, 2008;Adão et al., 2017;Chang et al., 2017). The UAV has in recent times been recognized and is receiving rapid attention as a useful tool for plant height estimations (Anthony et al., 2014;Han et al., 2018;Feng et al., 2019;Kawamura et al., 2020). The other advantage UAV offers is the fact that it is possible to incorporate normal red green blue (RGB) camera into its system which are familiar to many smallholder farmers in Africa who use mobile phones with RGB cameras to take photographs of their crops from time to time. ...
Article
This study aimed at estimating plant height of maize using low-cost unmanned aerial vehicle (UAV) based imagery. The experiment was laid out in a factorial arrangement in randomized complete block design. The factors were maize varieties (honampa, ahooden, ahoofe and abontem) and intra-row sowing spacing (20 cm, 30 cm and 40 cm). A constant inter-row spacing of 80 cm was used for all sub-plots. Data (both UAV-based imagery and manual data) were taken 21 days after sowing (DAS) -seedling/establishment stage-, 42 DAS (vegetative stage), 63 DAS (tasseling stage), and 84 DAS (physiological maturity stage). The results showed a high correlation between the UAV-based plant height and manual measurement with R2 (adjust) > 0.92 in all cases. For the intra-row sowing spacing, the 20 cm plant spacing was found to have had the best overall R2 (adjust) of 0.949, showing the strongest regression of UAV with the manual data among the treatments. Also, it was found that the best stage to estimate maize plant height is dependent on the variety being used since the honampa gave the best correlation coefficient (R = 0.85) at the tasseling stage while the ahooden, ahoofe and abontem gave optimum correlation coefficients (R) of 0.70, 0.90 and 0.69 respectively at the vegetative stage. Overall, the study shows potency of using UAV-based imagery for estimating plant height of maize. The findings of the study could be useful to agronomists and smallholder farmers in Ghana and other developing countries regarding the usefulness of the technology for in-field data collection and advisory services.
Chapter
Monitoring the development of trees and accurately estimating the yield are important to improve orchard management and production. Growers need to estimate the yield of trees at the early stage to make smart decisions for field management. However, methods to predict the yield at the individual tree level are currently not available due to the complexity and variability of each tree. This study aimed to evaluate the performance of an unmanned aerial vehicle (UAV)-based remote sensing system and machine learning (ML) approaches for a tree-level pomegranate yield estimation. Lightweight sensors, such as the multispectral camera, were mounted on the UAV platform to acquire high-resolution images. Eight characteristics were extracted, including the normalized difference vegetation index (NDVI), the green normalized vegetation index (GNDVI), the RedEdge normalized difference vegetation index (NDVIre), RedEdge triangulated vegetation index (RTVIcore), individual tree canopy size, the modified triangular vegetation index (MTVI2), the chlorophyll index-green (CIg), and the chlorophyll index-rededge (CIre). First, direct correlations were made and the correlation coefficient (R\({ }^2\)) was determined between these vegetation indices and tree yield. Then, machine learning approaches were applied with the extracted features to predict the yield at the individual tree level. The results showed that the decision tree classifier had the best prediction performance, with an accuracy of 85%. The study demonstrated the potential of using UAV-based remote sensing methods, coupled with ML algorithms, for estimating the pomegranate yield. Predicting the yield at the individual tree level will enable stakeholders to manage the orchard on different scales, thus improving field management efficiency.
Article
Full-text available
Estimating wheat yield accurately is crucial for efficient agricultural management. While canopy spectral information is widely used for this purpose, the incorporation of canopy volumetric features (CVFs) remains underexplored. This study bridges this gap by utilizing unmanned aerial vehicle (UAV) multispectral imaging to capture images and elevation data of wheat at key developmental stages—gestation and flowering stages. We innovatively leveraged the elevation differences between these stages to calculate canopy height, develop a novel CVF, and refine the wheat yield prediction model across various wheat varieties, nitrogen fertilizer levels, and planting densities. The integration of canopy volume information significantly enhanced the accuracy of our yield prediction model, as evidenced by an R ² of 0.8380, an RMSE of 313.3 kg/ha, and an nRMSE of 11.33%. This approach not only yielded more precise estimates than models relying solely on spectral data but also introduced a novel dimension to wheat yield estimation methodologies. Our findings suggest that incorporating canopy volume characteristics can substantially optimize wheat yield prediction models, presenting a groundbreaking perspective for agricultural yield estimation.
Article
Full-text available
Google Earth provides an open source, easy to access and cost free image data that support map interest community. Therefore, depending of this community on Google Earth, grows up day by day. More than simply providing locational information, Google Earth allows users to add their own content such as photos or descriptions of areas or landmarks. They can also extrapolate information from the satellite imagery obtained by digitizing areas of interest and exporting them for use elsewhere. As such, the application has found a strong following not only in explorers and navigators but also in classrooms all over the world. However, this popularity of Google Earth is not an indication of its accuracy. The aim of this research is to estimate the Google Earth horizontal and vertical accuracy in Khartoum State so as to evaluate this free source of data. This was carried out by comparing Google Earth measured coordinates of points with geodetic Global Positional System (GPS) receiver coordinates over sample of 16 check points located in Khartoum State. Since GPS provide accurate measurement of coordinates on the same ellipsoid as Google Earth, it was used to check the accuracy of Google Earth. Root Mean Square Error (RMSE) was computed for horizontal coordinates and was found to be 1.59m. For height measurement RMSE was computed to be 1.7m. For the research purposes and to pursue the changes occurred while Google Earth images updated, it was noted that the positional accuracy was changed and improved, but the elevation is still as it were before update.
Article
Full-text available
Biomass monitoring is one of the main pillars of precision farm management as it involves deeper knowledge about pest and weed status, soil quality, water stress, and yield prediction, among others. This research focuses on estimating crop biomass from high-resolution red, green, blue imaging obtained with an unmanned aerial vehicle. Onion, as one of the most cultivated vegetables, was studied for two seasons under non-controlled conditions in two commercial plots. Green canopy cover, crop height, and canopy volume (Vcanopy) were the predictor variables extracted from the geomatic products. Strong relationships were found between Vcanopy and dry leaf biomass and dry bulb biomass. Adjusted coefficient of determination (\({\text{R}}_{\text{adj}}^2\)) values were 0.76 and 0.95, respectively. Nevertheless, crop management practices and leaf depletion at vegetative stages significantly affect the accuracy of the canopy model. These results suggested that obtaining biomass using aerial images are a good alternative to other sensors and platforms as they have high spatial and temporal resolution to perform high-quality biomass monitoring.
Article
Full-text available
Non-destructive monitoring of crop development is of key interest for agronomy and crop breeding. Crop Surface Models (CSMs) representing the absolute height of the plant canopy are a tool for this. In this study, fresh and dry barley biomass per plot are estimated from CSM-derived plot-wise plant heights. The CSMs are generated in a semi-automated manner using Structure-from-Motion (SfM)/Multi-View-Stereo (MVS) software from oblique stereo RGB images. The images were acquired automatedly from consumer grade smart cameras mounted at an elevated position on a lifting hoist. Fresh and dry biomass were measured destructively at four dates each in 2014 and 2015. We used exponential and simple linear regression based on different calibration/validation splits. Coefficients of determination R 2 between 0.55 and 0.79 and root mean square errors (RMSE) between 97 and 234 g/m 2 are reached for the validation of predicted vs. observed dry biomass, while Willmott's refined index of model performance d r ranges between 0.59 and 0.77. For fresh biomass, R 2 values between 0.34 and 0.61 are reached, with root mean square errors (RMSEs) between 312 and 785 g/m 2 and d r between 0.39 and 0.66. We therefore established the possibility of using this novel low-cost system to estimate barley dry biomass over time.
Article
Full-text available
Accurate and timely detection of weeds between and within crop rows in the early growth stage is considered one of the main challenges in site-specific weed management (SSWM). In this context, a robust and innovative automatic object-based image analysis (OBIA) algorithm was developed on Unmanned Aerial Vehicle (UAV) images to design early post-emergence prescription maps. This novel algorithm makes the major contribution. The OBIA algorithm combined Digital Surface Models (DSMs), orthomosaics and machine learning techniques (Random Forest, RF). OBIA-based plant heights were accurately estimated and used as a feature in the automatic sample selection by the RF classifier; this was the second research contribution. RF randomly selected a class balanced training set, obtained the optimum features values and classified the image, requiring no manual training, making this procedure time-efficient and more accurate, since it removes errors due to a subjective manual task. The ability to discriminate weeds was significantly affected by the imagery spatial resolution and weed density, making the use of higher spatial resolution images more suitable. Finally, prescription maps for in-season post-emergence SSWM were created based on the weed maps-the third research contribution-which could help farmers in decision-making to optimize crop management by rationalization of the herbicide application. The short time involved in the process (image capture and analysis) would allow timely weed control during critical periods, crucial for preventing yield loss.
Article
Full-text available
Site-specific crop management is a promising approach to maximize crop yield with optimal use of rapidly depleting natural resources. Availability of high resolution crop data at critical growth stages is a key for real-time data-driven decisions during the production season. The goal of this study was to evaluate the possibility of using small unmanned aerial system (UAS)-based remote sensing technologies to monitor the crop stress of irrigated pinto beans (Phaseolus vulgaris L.) with varied irrigation and tillage treatments. A small UAS with onboard multispectral and infrared thermal imaging sensors was used to collect data from bean field plots on three growth stages in 2015 and 2016, respectively. Indicators including green normalized vegetation index (GNDVI), canopy cover (CC, ratio of ground covered by crop canopy to the total plot area) and canopy temperature (CT, °C) of crops were extracted from imaging data and correlated with ground-reference crop yield and leaf area index (LAI) estimated with a handheld ceptometer. Results show that GNDVI, CC and CT were able to differentiate crops with full and deficit irrigation treatments at each of the three growth stages in both years. Developed indicators were strongly correlated with to the crop yield with Pearson correlation coefficients (r) of approximate 0.71 and 0.72 for GNDVI and CC, respectively, in the early growth stage (54 days after planting) in both years. Canopy temperature showed even stronger correlation (r > 0.8) with yield at early growth stage. Performance of small UAS-based imagery-based indicators in crop stress monitoring and crop yield estimation was better than or comparable to that of the ground-based LAI estimates, indicating the potential of such remote sensing tool in rapid crop stress monitoring and management.
Article
Field-based next generation phenotyping has become of great interest to plant breeders and agricultural researchers in recent years, particularly for circumventing destructive or impractical phenotyping methods commonly used for certain traits. The non-destructive estimation of one such trait, above ground biomass (AGB), has been investigated repeatedly using 2D imagery, though little research has been conducted on 3D methods. The aims of the current study were to (i) investigate the use of readily-available consumer level digital cameras and software to estimate AGB, canopy height (CH) and harvest index (HI) of wheat plots, (ii) investigate the suitability of this data as a replacement for destructive sampling methods within a wheat breeding programme, and (iii) identify the point cloud density required for accurate estimation of AGB. To achieve this, a small plot trial of a single wheat cultivar was conducted in an irrigated nursery, at Roseworthy, South Australia. At physiological maturity plots were measured for CH and whole plots were harvested to attain AGB and threshed to measure grain yield and calculate HI. Prior to harvesting each plot was imaged using a digital camera, with these images being processed into 3D point clouds, which were subsequently used to estimate plot volume and CH. Strong correlations were observed between actual measurements of AGB, CH and HI to those estimated from point clouds. Images were processed in subset batches to determine an optimal number of images for processing. Stronger correlations between AGB and plot volume were observed when more images were processed, though as few as 48 images provided sufficiently accurate estimates of AGB. These methods were shown to be effective at estimating AGB, CH and HI and could be adopted by small scale research programmes. This study shows that a higher-throughput adaptation of this photogrammetry method could be used in phenotype intensive research such as plant breeding programmes.
Article
Estimating crop biophysical and biochemical parameters with high accuracy at low-cost is imperative for high-throughput phenotyping in precision agriculture. Although fusion of data from multiple sensors is a common application in remote sensing, less is known on the contribution of low-cost RGB, multispectral and thermal sensors to rapid crop phenotyping. This is due to the fact that (1) simultaneous collection of multi-sensor data using satellites are rare and (2) multi-sensor data collected during a single flight have not been accessible until recent developments in Unmanned Aerial Systems (UASs) and UAS-friendly sensors that allow efficient information fusion. The objective of this study was to evaluate the power of high spatial resolution RGB, multispectral and thermal data fusion to estimate soybean (Glycine max) biochemical parameters including chlorophyll content and nitrogen concentration, and biophysical parameters including Leaf Area Index (LAI), above ground fresh and dry biomass. Multiple low-cost sensors integrated on UASs were used to collect RGB, multispectral, and thermal images throughout the growing season at a site established near Columbia, Missouri, USA. From these images, vegetation indices were extracted, a Crop Surface Model (CSM) was advanced, and a model to extract the vegetation fraction was developed. Then, spectral indices/features were combined to model and predict crop biophysical and biochemical parameters using Partial Least Squares Regression (PLSR), Support Vector Regression (SVR), and Extreme Learning Machine based Regression (ELR) techniques. Results showed that: (1) For biochemical variable estimation, multispectral and thermal data fusion provided the best estimate for nitrogen concentration and chlorophyll (Chl) a content (RMSE of 9.9% and 17.1%, respectively) and RGB color information based indices and multispectral data fusion exhibited the largest RMSE 22.6%; the highest accuracy for Chl a + b content estimation was obtained by fusion of information from all three sensors with an RMSE of 11.6%. (2) Among the plant biophysical variables, LAI was best predicted by RGB and thermal data fusion while multispectral and thermal data fusion was found to be best for biomass estimation. (3) For estimation of the above mentioned plant traits of soybean from multi-sensor data fusion, ELR yields promising results compared to PLSR and SVR in this study. This research indicates that fusion of low-cost multiple sensor data within a machine learning framework can provide relatively accurate estimation of plant traits and provide valuable insight for high spatial precision in agriculture and plant stress assessment. Download link: https://authors.elsevier.com/c/1W01w3I9x1REBE DOI: DOI: 10.1016/j.isprsjprs.2017.10.011
Article
The Multiple Instance Hybrid Estimator for discriminative target characterization from imprecisely labeled hyperspectral data is presented. In many hyperspectral target detection problems, acquiring accurately labeled training data is difficult. Furthermore, each pixel containing target is likely to be a mixture of both target and non-target signatures (i.e., sub-pixel targets), making extracting a pure prototype signature for the target class from the data extremely difficult. The proposed approach addresses these problems by introducing a data mixing model and optimizing the response of the hybrid sub-pixel detector within a multiple instance learning framework. The proposed approach iterates between estimating a set of discriminative target and non-target signatures and solving a sparse unmixing problem. After learning target signatures, a signature based detector can then be applied on test data. Both simulated and real hyperspectral target detection experiments show the proposed algorithm is effective at learning discriminative target signatures and achieves superior performance over state-of-the-art comparison algorithms.
Article
Yield estimation is a critical task in crop management. Traditional methods are costly, time-consuming, and difficult to expand to a relatively large field. Remote sensing can provide quick coverage over a field at any scale. Satellite remote sensing is used for large-scale earth observation. Remote sensing with manned airplanes at relatively high altitudes (>500 m) has difficulty achieving the spatial resolution required for field-scale precision farming. Ground-based systems are typically used for point measurements and are restricted to field conditions. Unmanned aerial vehicles (UAVs) provide a unique platform for high-resolution remote sensing, and UAV-based remote sensing systems can be used to estimate crop yield in a cost-effective manner. The objective of this study was to develop and evaluate new methods for estimation of cotton yield for precision cotton farming. Experimental plots were laid out in a cotton field near Stoneville, Mississippi, in 2014. Nitrogen fertilizer was applied to the plots at five different rates to generate cotton yield variation. Two methods were employed to estimate cotton yield using very high-resolution digital images (2.7 cm pixel⁻¹) acquired from an inexpensive small multirotor UAV: (1) using three-dimensional point cloud data derived from multiple digital images of the cotton field to estimate cotton plant height and hence estimate yield, and (2) segmenting cotton boll signatures from the background of the digital images of the defoliated cotton field just prior to harvest and then estimating yield with the estimated cotton plot unit coverage. The results indicated that low-altitude remote sensing with an inexpensive small UAV can be used to estimate cotton yield accurately through estimation of plant height (R² = 0.43, compared with R² = 0.42 for yield estimation through manually measured plant height). The results further indicated that the method can offer reliable cotton yield estimation through estimation of cotton boll coverage in each plot with Laplacian image processing while considering a few plots with poor light condition as outliers (R² = 0.83). This study could benefit yield estimation of cotton, with similar methods used for other crops, in agricultural research and crop production. © 2016 American Society of Agricultural and Biological Engineers.
Article
Plant breeders and agronomists are increasingly interested in repeated plant height measurements over large experimental fields to study critical aspects of plant physiology, genetics and environmental conditions during plant growth. However, collecting such measurements using commonly used manual field measurements is inefficient. 3D point clouds generated from unmanned aerial systems (UAS) images using Structure from Motion (SfM) techniques offer a new option for efficiently deriving in-field crop height data. This study evaluated UAS/ SfM for multitemporal 3D crop modelling and developed and assessed a methodology for estimating plant height data from point clouds generated using SfM. High-resolution images in visible spectrum were collected weekly across 12 dates from April (planting) to July (harvest) 2016 over 288 maize (Zea mays L.) and 460 sorghum (Sorghum bicolor L.) plots using a DJI Phantom 3 Professional UAS. The study compared SfM point clouds with terrestrial lidar (TLS) at two dates to evaluate the ability of SfM point clouds to accurately capture ground surfaces and crop canopies, both of which are critical for plant height estimation. Extended plant height comparisons were carried out between SfM plant height (the 90th, 95th, 99th percentiles and maximum height) per plot and field plant height measurements at six dates throughout the growing season to test the repeatability and consistency of SfM estimates. High correlations were observed between SfM and TLS data (R2 = 0.88–0.97, RMSE = 0.01–0.02 m and R2 = 0.60–0.77 RMSE = 0.12–0.16 m for the ground surface and canopy comparison, respectively). Extended height comparisons also showed strong correlations (R2 = 0.42–0.91, RMSE = 0.11–0.19 m for maize and R2 = 0.61–0.85, RMSE = 0.12–0.24 m for sorghum). In general, the 90th, 95th and 99th percentile height metrics had higher correlations to field measurements than the maximum metric though differences among them were not statistically significant. The accuracy of SfM plant height estimates fluctuated over the growing period, likely impacted by the changing reflectance regime due to plant development. Overall, these results show a potential path to reducing laborious manual height measurement and enhancing plant research programs through UAS and SfM.