You are on page 1of 9

Computers and Electronics in Agriculture 147 (2018) 109–117

Contents lists available at ScienceDirect

Computers and Electronics in Agriculture


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

Original papers

Vineyard water status estimation using multispectral imagery from an UAV T


platform and machine learning algorithms for irrigation scheduling
management

Maria Romeroa,b, Yuchen Luoc, Baofeng Sud, Sigfredo Fuentese,
a
Agricultural Engineering University of Montpellier, School of Oenology, Faculty of Wine Science, 34060, France
b
University of Turin, School of Agricultural Science, Faculty of Agriculture, Forest and Food Science, 10124, Italy
c
Hochschule Geisenheim University, D-65366, Germany
d
College of Mechanical and Electronic Engineering, Northwest A&F University, Yangling 712100, Shaanxi, China
e
University of Melbourne, School of Agriculture and Food, Faculty of Veterinary and Agricultural Sciences, VIC 3010, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: Remote sensing can provide a fast and reliable alternative for traditional in situ water status measurement in
Remote sensing vineyards. Several vegetation indices (VIs) derived from aerial multispectral imagery were tested to estimate
Stem water potential midday stem water potential (Ψstem) of grapevines. The experimental trial was carried out in a vineyard in the
Artificial neural networks Shangri-La region, located in Yunnan province in China. Statistical methods and machine learning algorithms
Unmanned aerial vehicle
were used to evaluate the correlations between Ψstem and VIs. Results by simple regression between VIs in-
Water stress
dividually and Ψstem showed no significant relationships, with coefficient of determination (R2) for linear fitting
smaller than 0.3 for almost all the indices studied, except for the Optimal Soil Adjusted Vegetation Index
(OSAVI); R2 = 0.42 with statistical significance (p ≤ 0.001). However, results from a model obtained by fitting
using Artificial Neural Network (ANN), using all VIs calculated as inputs and real Ψstem from plants within the
study site (n = 90) as targets (Model 1), showed high correlation between the estimated water potential through
ANN (Ψstem ANN) and the actual measured Ψstem. Training, validation and testing data sets presented individual
correlations of R = 0.8, 0.72 and 0.62 respectively. The models obtained from the study site were then applied to
a wider area from the vineyard studied and compared to further Ψstem measured obtained from different sites
(n = 23) showing high correlation values between Ψstem ANN and real Ψstem (R2 = 0.83; slope = 1; p ≤ 0.001).
Finally, a pattern recognition ANN model (Model 2) was developed for irrigation scheduling purposes using the
same Ψstem measured in the study site as inputs and with the following thresholds as outputs: Ψstem below
−1.2 MPa considered as severe water stress (SS), Ψstem between −0.8 to −1.2 MPa as moderate stress (MS) and
Ψstem over −0.8 MPa with no water stress (NS). This model can be applied to analyze on a plant by plant basis to
identify sectors of stress within the vineyard for optimal irrigation management and to identify spatial variability
within the vineyards.

1. Introduction desired yield and quality. To manage irrigation scheduling there are
three questions to answer: (i) when to irrigate, (ii) how much to irrigate
Water is the key factor affecting quality and quantity of wine grapes. and (iii) where to apply water. The latter question can be only answered
Excessive water applications may result in increases of the vegetative with a spatial assessment of plant water status. Water potential (Ψ) is
growth and yield, but other parameters associated to quality traits, such one of the in situ methods, that has been traditionally effective to assess
as sugar content, acidity and pigment synthesis may be adversely af- crop water status in agriculture (Smith and Mullins, 2000; Choné, 2001;
fected. In contrast, severe water stress will drive to partial or complete Van Leeuwen et al., 2006; Jones, 2007). However, the most accurate
stomatal closure, causing a reduction on the photosynthetic activity measurements of Ψ can be achieved by using a pressure chamber
(Van Leeuwen et al., 2009). (Scholander et al., 1964), which applied to a large scale can be both
In regions where precipitations are scarce and concentrated in brief labor intensive and time consuming. Hence, adequate sampling and
periods along the year, irrigation plays a significant role to achieve repeatability are usually limited to a few sentinel plants and assumed to


Corresponding author.
E-mail address: sfuentes@unimelb.edu.au (S. Fuentes).

https://doi.org/10.1016/j.compag.2018.02.013
Received 12 December 2017; Received in revised form 26 January 2018; Accepted 10 February 2018
0168-1699/ © 2018 Elsevier B.V. All rights reserved.
M. Romero et al. Computers and Electronics in Agriculture 147 (2018) 109–117

Fig. 1. Location of the study in the Southern of China and reference plot used for this study. Yellow dots denote the locations where plant measurements were performed in a grid for each
cultivar: Chardonnay (n = 35) in the bottom section and Cabernet sauvignon (n = 55) at the top section. (For interpretation of the references to color in this figure legend, the reader is
referred to the web version of this article.)

be representative of the whole field. These assumptionscomplicate the water deficit in the photosynthesis process, specially, the epoxidation
implementation of efficient irrigation scheduling practices and any states of xanthophyll cycle pigments and the chlorophyll fluorescence
other management practices focused to control or uniform yield or emission (Moya et al., 2004). Even though there is not water absorption
quality within vineyards. The latter, mainly due to the heterogeneity of in the visible spectra domain, several VIs based on this region have
soils and microclimate found in most of vineyards around the world. shown high correlation levels with plant water status and are com-
Therefore, more efficient monitoring systems that have both, high monly used as its proxy to improve vineyard management (Govaerts
temporal and spatial resolutions are required. and Verhulst, 2010; Pôças et al., 2015; Rodríguez-Pérez et al., 2007).
Technological advances in remote sensing have provided non-in- The large amount of data collected through remote sensing allows to
vasive, time and cost-effective techniques that can detect the spatial describe almost the whole plant physiology and even changes within
variability of plant water status at a wider range of temporal scales plants depending on the resolution of sensors. Nevertheless, the re-
compared to any manual method (Acevedo-Opazo et al., 2008; Dorigo sponse of the plant to different environmental changes not always
et al., 2007). Over the last 30 years, spectral reflectance data have been corresponds to linear relationships, for this reason sometimes tradi-
used more and more in the studies of vegetation, due to the connection tional statistical approach are not powerful enough to obtain accurate
between the spectral proprieties and their biochemical and biophysical plant water status estimations. In this case, the use of Artificial Neural
attributes, e.g., biomass, leaf pigment contents, canopy water content, Network (ANN) to deal with large and complex task become a good
crop coefficient and crop evapotranspiration (Penuelas et al., 1997; alternative due to its ability to model both linear and nonlinear systems.
Zarco-Tejada et al., 2005; Rodríguez-Pérez et al., 2007;Baluja et al., The learning performed by a network runs automatically and it is based
2012; Ferreira et al., 2012; Hunt et al., 2012; Pôças et al., 2015; Costa on the selection of appropriate value of weight (Riad et al., 2004;
et al., 2016). Spectral data are commonly used as mathematical com- Samborska et al., 2014). The ANN models have shown to be powerful
binations of two or more spectral bands, which are named as spectral prediction tools for support decision making in the area of agriculture,
indices. The rising availability of sensors with the ability to provide several examples could be mentioned such as: rainfall-runoff process
reflectance information over a spectral range, has contributed to in- (Hsu et al., 1995), prediction of evaporation (Sudheer et al., 2002), soil
crease the interest in using vegetation indices (VI) based on the visible, erosion and its relationship with leaking of nutrients in plants (Kim and
red-edge and near infrared (NIR) regions of the electromagnetic light Gilley, 2008), estimation of plant water uptake (Qiao et al., 2010),
spectrum. The VIs based on the visible and red-edge regions have been plant classification using leaf recognition (Wu et al., 2007), crop yield
reported to be able to detect crop water status at the canopy level prediction (Khairunniza-Bejo et al., 2014) and vegetation mapping
(Rossini et al., 2013). This is based on the knowledge of the effect of (Carpenter et al., 1999). Nevertheless, in terms of modelling plant water

110
M. Romero et al. Computers and Electronics in Agriculture 147 (2018) 109–117

Table 1
Vegetation indices considered for this study with their respective equations and references.

Number Indices Abrev. Equation Reference

1 Difference Vegetation Index DVI NIR-RED Richardson and Everitt (1992)


2 Green Index GI GREEN/RED Chamard et al. (1991)
3 Modified Soil Adjusted Vegetation Index MSAVI 2NIR+1-(√2(NIR+1)2-√8(NIR-RED))/2 Qi et al. (1994)
4 Normalized Difference Vegetation Index NDVI (NIR-RED)/(NIR+RED) Rouse et al., (1973)
5 Normalized Difference Greenness Vegetation Index NDGI (NIR-GREEN)/(NIR+GREEN) Gamon and Surfus (1999)
6 Normalized Difference Red Edge Index NDRE (NIR-REG)/(NIR+REG) Barnes et al. (2000)
7 Optimal Soil Adjusted Vegetation Index OSAVI (NIR-RED)/(NIR+RED+0.16) Rondeaux et al. (1996)
8 Red Green Ratio Index RGRI RED/GREEN Gamon and Surfus (1999)
9 Renormalized Difference Vegetation Index RDVI (NIR-RED)/√NIR+RED Roujean and Breon (1995)
10 Simple ratio Index SR NIR/RED Birth and McVey (1968)

response, the high input-output mapping ability performance by the a total area of 4.7 ha.
ANN has presented higher performance accuracy (R2 = 0.96) than
multiple lineal regression to predict steam water potential from soil
2.2. Plant water status and UAV-based spectral field measurements
moisture measurements (Martí et al., 2013). Preliminary assessments
on grapevines using multispectral information obtained from UAV
Vine water status was assessed by Ψstem measured using two
platforms to assess plant water status and ANN modelling techniques
Scholander pressure bomb (PSM 600 D. Albany, OR, USA) between
have been recently implemented (Poblete et al., 2017). However, the
12:00 and 14:00, after two hours after covering the leaves and no
latter study included as inputs for ANN modelling, specific multi-
transpiration adaptation inside plastic/aluminum foil bags. The phy-
spectral bands and not indices as predictors of stem water potential
siological measurements were conducted on 90 vines distributed within
(Ψstem), which may hinder accuracy.
a grid in the measurement plot which were considered as observed data
The aim of this study was to assess vine water status (Ψstem) through
(Ψstem obs) for modelling purposes. Each measured vine was separated
VIs calculated from multispectral imagery as a possible alternative so-
by three adjacent vines as buffer to attain maximum variability within
lution for the traditional pressure chamber measurement. Furthermore,
the data for modelling purposes. The experiment was carried out at
this study also aimed to implement machine learning modelling using
véraison during the 2016 growing season in three dates: October 7th,
all the relevant VIs reported in previous literature as linked to plant
October 22nd and November 4th.
water status as inputs.
The reflectance data of the whole experimental vineyard were col-
lected using a UAV platform (Phantom 3 professional, DJI Co. Ltd,
Shenzhen, China) equipped with a SEQUOIA multispectral sensors
2. Material and methods
(Parrot Co. Ltd, Paris, France). The flight height was 30 m above ground
level, which allowed obtaining imagery at 4 cms spatial resolution in
2.1. Vineyard site description and vine material
the VIS, GREEN, RED, RED EDGE and NIR region. Imagery was ac-
quired at 13:00. for all the bands in one manual flight per day of
The research was carried out in a vineyard located in the Yunnan
measurement.
province (28°33′47.91″N and 98°52′18.05″E; 2300 m.a.s.l.), China
(Fig. 1). Vineyards are dominantly on slopes and terraces, and this
makes it a unique landscape favoring the wine production style from 2.3. Image processing, data analysis and machine learning modelling
the region. The vineyard was originally established by local Tibetan
farmers, based on community family units. Thus, plots usually are The full size raw multispectral images were processed using Pix4d
smaller than half hectare each. The climate type of Dêqên County can mapper 3.0.1 (Pix4D S A, Switzerland) to obtain the ten vegetation
be described as site-specific. It varies along with the change of altitude, indices calculated (Table 1). The geometric correction and resampling
which contributes to the long vine growing season in the area. In at 10 cm spatial resolution were performed using Quantum GIS Geo-
general, it is a transition between a subtropical highland climate and graphic Information System 2.14.11 (Qgis, Free Software Foundation,
humid continental climate. The mean annual total precipitations can Boston, USA) using one measurement as a control point for the cali-
exceed 250 mm and more than 70% concentrates during the months of bration of the remaining dates. The pure canopy pixel extraction was
June, July and August. The soil texture is mainly sandy clay with clay performed for each index and date from plants measured for midday
content ranging from 10 to 25%. Ψstem using System for Automated Geoscientific Analyses version 3.0.0
The main cultivars used for modelling purposes were Cabernet (SAGA, Free Software Foundation, Boston, USA). From these data, ten
Sauvignon and Chardonnay both training on vertical shoot positioned VIs, which incorporate the GREEN, RED, RED EDGE and NIR region
and planted on 2004 on their own roots. The water was supplied using were calculated (Table 1). To compare Ψstem with the VIs calculated
furrow irrigation with a maximum furrow width not exceeding 150 cm. from data collected from the aerial imagery, linear and curvilineal
The plantation distance is mainly 1.2 m between vines and 1.5 m be- models were fitted to the data using Microsoft Excel 2016 (Microsoft,
tween rows. The area considered in this study has a total surface of USA, Washington).
2.8 ha. One plot planted with Cabernet Sauvignon and Chardonnay The machine learning modelling method used to process multi-
with no irrigation during the period of study was selected to be used as spectral and physiological data (Model 1) was implemented using the
a reference area to measure intensively vine water status and to gen- Matlab Neural Network Toolbox™ 10 (Mathworks Inc., Matick, MA,
erate the machine learning models. USA). All the ten VIs calculated were considered as inputs and midday
Once the machine learning models were developed, they were ap- Ψstem obs as targets for all plants measured in all measurement dates
plied to the whole valley, which have Cabernet Sauvignon and Merlot (n = 270). Model 1 (Fig. 2A) was an ANN fitting model constructed
cultivars trained on vertical shoot positioned system (VSP) and planted using 10 neurons for the hidden layer, data division was performed
on 2004 and 2012 respectively. The whole valley is irrigated using automatically by dividing the data into three stages: (i) training, which
furrow irrigation system with the same planting distance of 1.2 × 1.5 m is necessary to compute the gradient and to update the weights and
between plants and rows respectively. The whole valley vineyards have biases, (ii) validation, which is needed to minimize the error and avoid

111
M. Romero et al. Computers and Electronics in Agriculture 147 (2018) 109–117

testing using a default derivative function (Mathworks Inc. Matick, MA.


USA).

2.4. Statistical analysis

To assess and compare the accuracy and performance of the ANN


Model 1 obtained, statistical data such as the determination coefficients
(R2), standard error of estimates (SEE) and root mean square error
(RMSE) were obtained for the linearized model output. There was no
outlier analysis for any of the fitting models presented, hence models
include all the data (n = 90) from replicates for Ψstem measurements
from the three dates, which sum up a total of 270 data. The statistical
Fig. 2. (A) Model 1, considering a two-layer feedforward network with ten hidden neu-
rons and sigmoid functions to fit regression models for ten vegetation indices (VIs) as
significance was calculated using Turkey HDS range test.
inputs and stem water potential (Ψstem) as target; (B) Model 2 for pattern recognition
considering also ten VIs as inputs and three water stress categories: Non-stressed (NS)
vines (Ψstem ≤ 0.8 MPa), moderate stressed (MS) vines (0.8 MPa ≥ Ψstem ≤ 1.2 MPa) and 2.5. Assessment of model performances at the pixel by pixel, plant by plant
severe stressed (SS) vines (Ψstem ≥ 1.2 MPa). For hidden and output layers, w = weights and vineyard scales
and b = biases.

Once VIs were calculated for all the study area, multispectral image
overfitting by monitoring the validation error in the training stage, and processing algorithms were applied to obtain a concatenated matrix
(iii) testing, which is necessary to compare different models. Specifi- cube of size (M ∗ N ∗ VIL). In which M represent the number or pixel
cally, a random data division function in Matlab® was used, which rows within spectral images, N the number of columns and VIL the
randomly divides the data from samples into the three stages with 60% vegetation index image layers considered (n = 10). Models 1 and 2
(n = 162) for training, 20% (n = 54) for validation with a mean were then applied to process pixel by pixel from the M,N,VIL cube using
squared error performance algorithm and 20% (n = 54) for testing with a customized loop code to extract Ψstem per pixel (Model 1) and nu-
a default derivative function, the data used in each stage is not used in merical values corresponding to the three water stress levels considered
the other two, which makes every stage independent from each other. (Model 2), as 1 = NS, 2 = MS and 3 = SS. Values below 1 were ob-
Model 2 was constructed to train a pattern recognition neural net- tained in sections with no spectral indices available.
work model (Fig. 2B). The objective of the model was the classification The plant by plant assessment was performed after applying Model
of water status levels based on ten VIs measured as inputs (Table 1) into 2 to the M,N,VIL cube using the method proposed by (Su et al., 2016)
the three levels of vine water stress as targets, namely: Non-stressed which recognizes the grapevines rows and separate each plant in a box
(NS) vines (Ψstem ≤ 0.8 MPa), moderate stressed (MS) vines corresponded to the plantation space, which is 1.2 m between plants
(0.8 MPa ≥ Ψstem ≤ 1.2 MPa) and severe stressed (SS) vines and 1.5 m between rows. This procedure allows the automatic calcu-
(Ψstem ≥ 1.2 MPa). The ANN are capable of finding patterns in an in- lation of data within each region corresponding to each plant. Data
ductive way through algorithms based on the pre-existing data Beale extraction was obtained as the mode water stress threshold value in the
et al. (2010). The ANN used in this study was a standard two-layer middle section of each plant section. Once having the processed matrix
feedforward network with a sigmoid transfer function in the hidden representing water stress threshold values per plant, a mesh grid spline
layer and a linear transfer function in the output layer structure con- interpolation (based on a cubic spline using a not-a-knot end condi-
formed using the 10 components (VIs) obtained using the UAV tions) technique was used to map spatial differences of thresholds. The
(Table 1) as the input layer and three water stress levels as the targets. interpolated value at specific analysis points is based on a cubic inter-
The selected number of neurons in the hidden layer was ten as depicted polation on the values at neighboring grid points for each respective
in Fig. 2B. The validation and testing data were randomly divided using dimension (Mathworks Inc., Matick. MA. USA).
a random data division function in 60% (n = 162) for training using a
scale conjugate gradient training function, 20% (n = 54) for validation
using a cross-entropy performance function and 20% (n = 54) for

Fig. 3. Spatial distribution of Stem water potential (MPa) in the study plot during the three dates using kriging interpolation techniques from all Ψstem obs measurement points per date
(n = 90).

112
M. Romero et al. Computers and Electronics in Agriculture 147 (2018) 109–117

3. Results and discussion not higher than R2 = 0.25. The statistical analysis showed significant R
and R2 between Ψstem obs and NDGI, GI and OSAVI indices. In parti-
3.1. Vine water status spatial assessment cular, OSAVI showed the best performance, R2 = 0.42 (p ≤ 0.001).
OSAVI has been described by Rondeaux et al. (1996) as an adjusting
The spatial distribution of Ψstem obs of 90 sampled vines for three factor for the SAVI family of indices, comparing to some typical indices
measurement dates are shown in Fig. 3. The range of Ψstem obs values like NDVI or DVI. OSAVI has enhanced sensitivity to vegetation cover
found were from −0.5 MPa to −1.5 MPa, showing that the temporal and soil brightness. Moreover, it allows to minimize the NDVI sensi-
and spatial distribution of plant water status were from non-stress to tivity to variation in soil background; for this reason, it is used in zones
severe stress levels. In the first measurement date (October 7th), most of with spare vegetation where soil is visible through the canopy and also
the area was not stressed, largely due to irrigation received just five where NDVI saturates due to high vegetation density. In this study,
days before. Only 20% of measurements showed Ψstem obs values lower OSAVI was the index that best correlated with Ψstem obs, one of the
than −1.0 MPa, mainly concentrated in the southwestern edge of this reasons could be related to the high vine density and the narrow leaf
plot (green to red color). The second measurement date (October 22nd) area in the Southwestern zone of the plot, which favors the saturation of
showed that vines with Ψstem obs lower than −1.0 MPa reached 30% NDVI value. Indexes, such as NDVI, NDRE, NDGI or DVI were expected
and the stressed zone was extended from the edge to the center. In the to be somehow better correlated with plant water status as many au-
last measurement date (November 4th), around 40% of vines showed thors have previously reported (Carter, 1993; Penuelas et al., 1997;
Ψstem obs lower than −1.0 MPa, hence the stressed area increased sig- Rodríguez-Pérez et al., 2007; Eitel et al., 2011; Baluja et al., 2012;
nificantly. Ferreira et al., 2012; Calderón et al., 2013; Van Beek et al., 2013; Zarco-
Several studies (Choné, 2001; Williams and Araujo, 2002; Van Tejada et al., 2013; Pôças et al., 2015). However, poor correlations and
Leeuwen et al., 2009) have successfully demonstrated the accuracy of high SEE for all the indices were in contradiction to these studies, which
midday Ψstem to assess plant water status in vineyards. Besides, similar limited the possibility of using a standard spectral wavelength for-
results have been obtained in others crops, such as pear, citrus and mulation or a single VI for real-time prediction of water status.
apple. Van Beek et al. (2013) showed a decrease of Ψstem under dif- To verify if there was any improvement of the correlations under
ferent water status regime in pear orchards. However, there is an im- different stress levels, the data collected were divided by measurement
portant source of variability of Ψstem among the vines of the same row dates. Results in Table 3 showed that most of the VIs are poorly or not
and area due to heterogeneity of soil and microclimate, which could well correlated with Ψstem obs among the three dates. Besides, there was
lead to errors in irrigation management decisions when selecting sen- no significant improvements in correlations. It should be pointed out
tinel vines that are not representative for the majority of plants for each that in the last measurement date (November 4th), which corresponded
plot. Van Leeuwen et al. (2006) also showed high variability in vine to to severe water stress in considerable parts of the reference plot, a few
vine water status in the same plot. On the other hand, the measurement VIs showed a significant correlation (p ≤ 0.01) with Ψstem, however as
could be influenced by several factors such as microclimate changes it was mentioned before, the R2 was considerably lower. According to
which affect evapotranspiration of plants, position and condition of the these results presented GI, RDVI, OSAVI, MSAVI and DVI corresponded
sampling leaf, petiole cut, among others (Turner, 1988). to the indices that could predict better long-term water stress. More-
The pressure bomb method to measure plant water status is very over, GI and OSAVI displayed an increasing tendency of the correlation
accurate as described before, however it offers low spatial resolution in improvements over time. In general, it is possible to say that Red and
practical terms. This method is restricted to sentinel plants for man- NIR spectral region could serve as long-term predictors of water stress.
agement purposes and it is still considered an expensive and time- Several authors (Ferreira et al., 2012; Jones, 2007; Pôças et al.,
consuming, since it requires trained personnel to operate a heavy in- 2015; Williams and Araujo, 2002) have suggested the relevance of pre-
strument (chamber plust N2 gas cylinder and attachments). dawn water potential (Ψpw) as a better indicator of water status due to
the equilibrium between plant and soil. However, Pôças et al. (2015)
using optimized VI by narrow bands reported poor correlation with Ψpw
3.2. Vine water status assessment by UAV-based multispectral imagery and indices that incorporate information between Red and NIR region
(R2 ≤ 0.6). Nevertheless, in the same study, better correlations were
3.2.1. Analysis of the different predictors for vine water status found with indices in the visible domain spectrum like GI, NDGI and
Relationships between vine water status and data retrieved from RGRI with R2 ranging from 0.77 to 0.82. Furthermore, evidence has
multispectral imagery and respective VIs calculated from Ψstem obs been presented using night-time transpiration sensors (sap flow) that
monitored plants are presented in Table 2. Linear and polynomial re- have demonstrated that in Mediterranean conditions eucalyptus
lationships were found between the VIs calculated and Ψstem obs, from (Fuentes et al., 2007), almond (Fuentes et al., 2013) and grapevines
which polynomial fits had a better performance. Nevertheless, the R2 (Fuentes et al., 2014) open stomata at night-time and significant water
for most of the indices in this study were low and not statistically sig- loss can be achieved, which makes difficult or even impossible the
nificant, which fluctuated between R2 = 0.09–0.49 and in most cases

Table 2
Statistical analysis showing the correlations and equations between the different vegetation indices studied and Ψstem.

Index Equation (Ψstem=) R (n = 270) R2 b


(n = 270) SEE RMSE

NDVI 2 a
20.824x - 40.015x + 26.531 −0.43 0.20 2.61 2.57
NDGI 27.945x2 - 15.941x + 9.5242 −0.40 0.18* 2.65 2.61
NDRE 11.969x2 - 17.259x + 10.793 −0.13 0.05 2.87 2.80
GI 1.012x2 - 5.8603x + 14.154 −0.37 0.14** 2.69 2.67
RGRI 12.283x2 - 16.031x + 12.777 0.39 0.20 2.67 2.58
SR 0.0308x2 − 0.9932x + 14.446 −0.40 0.27 2.65 2.46
RDVI −3.6x2 + 6.8144x + 5.9953 −0.16 0.05 2.86 2.81
OSAVI 52.796x2 - 71.675x + 31.293 −0.60 0.42*** 2.32 8.02
MSAVI 14.42x2 - 24.391x + 17.078 −0.43 0.25 2.60 2.50
DVI 34.37x2 - 35.221x + 15.681 −0.41 0.23 2.64 2.53

a
Correspond to the respective index value; bSignificance level: *p ≤ 0.05; **
p ≤ 0.01; p ≤ 0.001; x denoted the specific index used for the fitting.
***

113
M. Romero et al. Computers and Electronics in Agriculture 147 (2018) 109–117

Table 3 set and calculated VIs of further random plants with respective Ψstem
Statistics analysis of correlations (R) between different vegetation indexes and Ψstem obs measured from another vineyard plot (n = 28). The linear regression
(MPa) for the three dates studied.
between Ψstem obs and Ψstem ANN, as expected had a significant
Index R (n = 90) (p ≤ 0.001) and high correlation (R = 0.83; slope = 1.0). In both cases,
the residual analyses showed independence, normality, and homo-
7-Oct 22-Oct 4-Nov geneity of the variance (data not shown). These results also showed that
more than 74% of the variance in the Ψstem obs can be explained by the
NDVI −0.54 −0.42 −0.35
NDGI −0.42 −0.27 −0.38 ANN model. A possible explanation for the good ANN results could be
NDRE −0.13 −0.07 −0.12 associated to the wavelength range in which each index is calculated.
GI −0.29 −0.36 −0.35** Most of the indices in the present research are built by bands in the Red
RGRI 0.41 0.50 0.46 and NIR spectra, commonly associated to physiological changes in
SR −0.40 −0.55 −0.40
plants (Kazmierski et al., 2011; Rossini et al., 2013; Gamon et al., 2015;
RDVI −0.56 −0.67 −0.60**
OSAVI −0.63 −0.59 −0.62** Pôças et al., 2015). In addition, Green and Red edge bands have also
MSAVI −0.50 −0.45 −0.37** been mentioned as water status indicators in vegetative tissue.
DVI −0.47 −0.43 −0.35** Zygielbaum et al. (2009) demonstrated the strong correlation of
greenness measure with crop water status, mainly because water stress
manifests an increase in green reflectance. Instead, Red edge also has
equilibrium between plant and soil water potentials, hence rendering
been associates to the chlorophyll content and considered as a possible
inaccuracies in the use of Ψpw in these conditions. All the latter studies
predictor of crop water status (Holer et al., 1983; Fillela and Penuelas,
were conducted under pressurized irrigation systems (mainly drip ir-
1994; Pinar and Curran, 1996).
rigation) in which non-uniform wetting patterns are achieved. This can
In fact, the wavelengths and their respectively indices used in this
be the same case with furrow irrigation systems used in this study,
study have the potential to be predictor of water status, but at the same
which result in non-uniform wetting patterns.
time each index presents some problems associated to reflectance and
On the other hand, similar research done in citrus groves by Waldo
radiance, e.g. DVI does not deal with difference between soil and ve-
and Schumann (2009) showed significant correlation between Ψstem
getation in shady area; NDVI could present oversaturation in high
and NDVI (R2 = 0.31). Furthermore, Rallo et al. (2014) reported good
density vineyards and does not eliminate atmospheric effect; SR reduces
prediction of Ψleaf in olive orchards using optimized indices like NDGI
the effects of atmosphere and topography and OSAVI reduces problems
and GI.
of saturation (Carter, 1993; Rondeaux et al., 1996). In this case, it is
possible to associate the ANN results to a compensation effect. That is to
3.2.2. Estimating Ψstem based on artificial neural networks (ANN) say, once the ten VI are combined, using ANN as a tool to evaluate the
The ANN uses a two-layer feedforward network and non-linear influence of each index over the target can significantly improve the
methods to assess patterns between the inputs and targets producing a estimation of Ψstem. Furthermore, even though some VIs indexes did not
linearized model for simplification (Model 1). All the VIs used as inputs correlate well with Ψstem obs it is not required to discard them for the
showed positive and linear/non-linear correlations. Model 1 showed a ANN model construction, since these modelling techniques can identify
strong and statistically significant performance when comparing ob- the pattern of the data and individual weights of inputs and fit the ANN
served stem water potential (Ψstem obs) versus estimated stem water model accordingly.
potential (Ψstem ANN). Specifically, the ANN regressions (Table 4), Nowadays, due to great computing power advancements, modelling
showed high correlations between Ψstem obs and the Ψstem ANN using ANN requires ubiquitous personal computers. The modelling
(R = 0.88) for the training data (n = 176). Additionally, the random technique can be performed by consultancy services and research in-
data used for the validation and testing also presented high correlations stitutions that have access to these software platforms, such as Matlab®.
with Ψstem (R = 0.73 and 0.62 respectively). The final linearized neural The product obtained by this technique for growers would be a software
fit regression combines the results obtained from the training, valida- package containing the ANN modelling algorithms to process auto-
tion and testing data. As expected, a high correlation was observed matically information from UAV multispectral imagery for irrigation
between Ψstem obs and Ψstem ANN (R = 0.84; slope = 1.0). It is important management purposes. This will require normal computer skills level
to note that the final model is not linear, since the linearization is a for users.
simplification of the complex model procedures to assess accuracy only.
The final model considers sigmoidal functions as presented in Fig. 2, 3.2.3. Using Ψstem estimation to manage vineyard irrigation
which represent most of processes found in nature. Overall, the analysis of canopy reflectance showed high potential
Model 1 was tested a second time, besides the model performance for Ψstem estimation. As an illustration on how Model 1 could be used
assessment presented before, to predict Ψstem using a multispectral data for irrigation management, Fig. 4A shows the vineyard map using Ψstem
ANN from Model 1 (RMSE = 0.15 MPa) compared to Ψstem obs (Fig. 4B).

Table 4 Model 1 rendered adequate accuracy, which is important for irrigation


Neural fitting using Leveng-Marquard algorithm (Model 1) with VIs as inputs and Ψstem as scheduling management, considering the criteria to determine levels of
target. Pattern recognition model (Model 2) using VIs as inputs and three water status stress. As previously presented, the criteria has as minimum 0.2 MPa
classification levels as targets: Non-stressed (NS) vines (Ψstem ≤ 0.8 MPa), moderate (Non-stressed (NS) vines (Ψstem ≤ 0.8 MPa), moderate stressed (MS)
stressed (MS) vines (0.8 MPa ≥ Ψstem ≤ 1.2 MPa) and severe stressed (SS) vines
(Ψstem ≥ 1.2 MPa). Na = non-applicable.
vines (0.8 MPa ≥ Ψstem ≤ 1.2 MPa) and severe stressed (SS) vines
(Ψstem ≥ 1.2 MPa). Hence the error of Model 1 is lower than this cri-
Stage Samples MSE R Accuracy (%) terion (0.15). The map created through the ANN Model 1 estimations
shows several zones in the vineyard requiring different irrigation
Training 162 2.68 0.88 Na
Validation 54 3.35 0.73 Na
management strategies. According to literature review, Ψstem below
Testing 54 3.68 0.62 Na −1.5 MPa should be considered as a cut-off between moderate and
Model 1 270 3.24 0.84 Na severe water stress (Choné, 2001; Van Leeuwen et al., 2009). However,
Training 162 Na Na 83.5 these thresholds could be considered as more appropriate to cultivars
Validation 54 Na Na 82.9
such as Shiraz (Acevedo-Opazo et al., 2010).
Testing 54 Na Na 82.9
Model 2 270 Na Na 83.3 For the present study, Ψstem below −1.2 MPa (orange to red colors)
were considered as cut off between moderate and severe water stress,

114
M. Romero et al. Computers and Electronics in Agriculture 147 (2018) 109–117

Fig. 4. Interpolated map of stem water potential (MPa) for reference plot on 22nd November: (A) Ψstem ANN map estimated through ANN model 1 and (B) observed Ψstem obs.

which have been shown to be more appropriated for cultivars such as


Chardonnay (Williams and Araujo, 2002). Furthermore, Ψstem between
−0.8 and −1.2 MPa (light blue to green color) corresponded to mod-
erate stress and finally Ψstem over −0.8 MPa (blue color) was con-
sidered with no water stress. Some studies done in Cabernet sauvignon
and Chardonnay considered plants without water stress when Ψstem was
over −0.6 MPa, instead when Ψstem fell to −1.2 MPa vines are con-
sidered under water deficit, which had a positive effect in the devel-
opment of aroma in cultivars such a Chardonnay (Deluc et al., 2009).
Using the previous classification and the ANN Model 2, it was possible
to create an irrigation management map for the whole vineyard. Fig. 5,
shows a vineyard water status classification at the pixel by pixel scale
using Model 2. Three main groups were used: NS vines (blue color), MS
(green to yellow color), SS (red color). By obtaining regular Ψstem maps
throughout the growing irrigation season using the methodology pro-
posed in this paper, irrigation scheduling using UAV-based remote
sensing becomes more practical for the assessment of spatial variability
of water status.
Finally, Fig. 6 shows an example of Model 2 applied to Cabernet
Sauvignon and Merlot plots (yellow oval from Fig. 5) using a plant by
plant data extraction procedure (Fig. 6A) and spline interpolation
techniques (Fig. 6B). The latter figure shows two defined spatial sectors
of NS, and SS with some plants surrounding SS at MS level. From
Fig. 6B, it can be seen a high spatial variability of vine water status,
which can give information to growers for site specific management
strategies. These management strategies can be focused to either
manage variability or to maintain a certain wine style, which many
times defines a region.

4. Conclusion

The use of multispectral sensors onboard of a UAV platform allowed


obtaining high-resolution images to monitor vine water status, which
has the potential to replace the traditional pressure chamber method,
especially for vineyards in extreme water stress conditions. Machine
Fig. 5. Stem water potential (MPa) estimate on a pixel by pixel basis through the ANN learning methods (ANN) demonstrated to be a powerful tool to mining
Model 2 for the Adong vineyard. Non-stressed (NS) pixels (Ψstem ≤ 0.8 MPa), moderate data from multi-dimensional maps. As it was shown in this study, once
pixels (MS) vines (0.8 MPa ≥ Ψstem ≤ 1.2 MPa) and severe stress (SS) pixels
the ANN models were trained using different VIs, high correlations
(Ψstem ≥ 1.2 MPa). The yellow broken lined oval represents the Cabernet sauvignon and
were found between the estimates (Ψstem ANN) and actual Ψstem obs
Merlot plots considered in Fig. 6. (For interpretation of the references to color in this
figure legend, the reader is referred to the web version of this article.) (Model 1). A pixel by pixel data information availability per plant may
be useful for physiological studies and more in-depth irrigation prac-
tices, such as identification of localized irrigation problems (broken
emitters or flooded furrows). However, this level of information offers

115
M. Romero et al. Computers and Electronics in Agriculture 147 (2018) 109–117

A B Stress Level
3 (SS)

2.5

2.0 (MS)

1.5

1 (NS)

0.5

0
Fig. 6. Stem water potential (MPa) estimated on a plant by plant basis (A) using Model 2 and interpolated by spline interpolation techniques (B) for sections of cultivars Cabernet
Sauvignon (top block) and Merlot (bottom block) for the Adong vineyard. Map B show mainly two sections of non-stressed (NS) vines (Ψstem ≤ 0.8 MPa, blue), and severe stress (SS) vines
(Ψstem ≥ 1.2 MPa, yellow) with some of the latter section surrounded with vines at moderate stress (MS) vines (0.8 MPa ≥ Ψstem ≤ t1.2 MPa, green). (For interpretation of the references
to color in this figure legend, the reader is referred to the web version of this article.)

little help for more practical irrigation scheduling strategies based on Chamard, P., Courel, M.F., Docousso, M., Guénégou, M.C., LeRhun, J., Levasseur, J.,
irrigation blocks. To address the latter problem, Model 2 showed to Togola, M., 1991. Utilisation des bandes spectrales du vert et du rouge pour une
meilleure évaluation des formations végétales actives In Télédétection et
perform accurately in the identification of regions within vineyards of Cartographie. AUPELF-UREF, Sherbrooke, QC pp. 203–209.
low, medium and high-water stress, which offer spatial information to Choné, X., 2001. Stem water potential is a sensitive indicator of grapevine water status.
growers for optimal irrigation scheduling practices and to identify Ann. Bot. 87, 477–483. http://dx.doi.org/10.1006/anbo.2000.1361.
Costa, J.M., Vaz, M., Escalona, J., Egipto, R., Lopes, C., Medrano, H., Chaves, M.M., 2016.
strategies to manage this variability more efficiently. This study con- Modern viticulture in southern Europe: Vulnerabilities and strategies for adaptation
firmed that an UAV platform equipped with multispectral sensors and to water scarcity. Agric. Water Manag. 164, 5–18. http://dx.doi.org/10.1016/j.
machine learning modelling strategies could be an efficient and af- agwat.2015.08.021.
Deluc, L.G., Quilici, D.R., Decendit, A., Grimplet, J., Wheatley, M.D., Schlauch, K.A.,
fordable option to assess vine water status and to map vineyard spatial Mérillon, J.-M., Cushman, J.C., Cramer, G.R., 2009. Water deficit alters differentially
variability to optimize irrigation management. Other applications such metabolic pathways affecting important flavor and quality traits in grape berries of
as biomass calculation, weeds monitoring and disease detection using Cabernet Sauvignon and Chardonnay. BMC Genomics 10, 212. http://dx.doi.org/10.
1186/1471-2164-10-212.
these techniques are worthy of further investigation. Nevertheless,
Dorigo, W.A., Zurita-Milla, R., de Wit, A.J.W., Brazile, J., Singh, R., Schaepman, M.E.,
problems associated to irradiance and reflectance still need to be in- 2007. A review on reflective remote sensing and data assimilation techniques for
vestigated for other potential applications. enhanced agroecosystem modeling. Int. J. Appl. Earth Obs. Geoinf. 9, 165–193.
http://dx.doi.org/10.1016/j.jag.2006.05.003.
Eitel, J.U.H., Vierling, L.A., Litvak, M.E., Long, D.S., Schulthess, U., Ager, A.A., Krofcheck,
Appendix A. Supplementary material D.J., Stoscheck, L., 2011. Broadband, red-edge information from satellites improves
early stress detection in a New Mexico conifer woodland. Remote Sens. Environ. 115,
Supplementary data associated with this article can be found, in the 3640–3646. http://dx.doi.org/10.1016/j.rse.2011.09.002.
Ferreira, M.I., Conceição, N., Silvestre, J., Fabião, M., 2012. Transpiration and water
online version, at http://dx.doi.org/10.1016/j.compag.2018.02.013. stress effects on water use, in relation to estimations from NDVI: application in a
vineyard in SE Portugal. Options Méditerranéennes. Ser. B Stud. Res. 67, 203–208.
References Fillela, I., Penuelas, J., 1994. The red edge position and shape as indicators of plant
chlorophyll content, biomass and hydric status. Int. J. Remote Sens. http://dx.doi.
org/10.1080/01431169408954177.
Acevedo-Opazo, C., Tisseyre, B., Ojeda, H., Guillaume, S., 2010. Spatial extrapolation of Fuentes, S., De Bei, R., Collins, M.J., Escalona, J.M., Medrano, H., Tyerman, S., 2014.
the vine (Vitis vinifera L.) water status: a first step towards a spatial prediction model. Night-time responses to water supply in grapevines (Vitis vinifera L.) under deficit
Irrig. Sci. 28, 143. irrigation and partial root-zone drying. Agric. Water Manag. 138, 1–9. http://dx.doi.
Acevedo-Opazo, C., Tisseyre, B., Ojeda, H., Ortega-Farias, S., Guillaume, S., 2008. Is it org/10.1016/j.agwat.2014.02.015.
possible to assess the spatial variability of vine water status? J. Int. des Sci. la Vigne Fuentes, S., Mahadevan, M., Bonada, M., Skewes, M.A., Cox, J.W., 2013. Night-time sap
du Vin. 42, 203–220. flow is parabolically linked to midday water potential for field-grown almond trees.
Baluja, J., Diago, M.P., Balda, P., Zorer, R., Meggio, F., Morales, F., Tardaguila, J., 2012. Irrig. Sci. 31, 1265–1276. http://dx.doi.org/10.1007/s00271-013-0403-3.
Assessment of vineyard water status variability by thermal and multispectral imagery Fuentes, S., Palmer, A.R., Taylor, D., Zeppel, M., Whitley, R., Eamus, D., 2007. An au-
using an unmanned aerial vehicle (UAV). Irrig. Sci. 30, 511–522. http://dx.doi.org/ tomated procedure for estimating the leaf area index (LAI) of woodland ecosystems
10.1007/s00271-012-0382-9. using digital imagery, MATLAB programming and its application to an examination
Barnes, E., Clarke, T., Richards, S., Colaizzi, P.D., Haberland, J., Kostrzewski, M., Waller, of the relationship between remotely sensed and field measurements of LAI. 5th Int.
P., Choi, C., Riley, E., Thompson, T., Lascano, R.J., Li, H., Moran, M.S., 2000. Work. Funct. Struct. Plant Model 1070–1079. http://dx.doi.org/10.1071/fp08045.
Coincident Detection of Crop Water Stress, Nitrogen Status and Canopy Density Using Gamon, J.A., Kovalchuck, O., Wong, C.Y.S., Harris, A., Garrity, S.R., 2015. Monitoring
Ground-Based Multispectral Data. Proc. Fifth Int. Conf. Precis. Agric. [CD Rom]. seasonal and diurnal changes in photosynthetic pigments with automated PRI and
Beale, M.H., Hagan, M.T., Demuth, H.B., 2010. Neural Network ToolboxTM User’s Guide. NDVI sensors. Biogeosciences 12, 4149–4159. http://dx.doi.org/10.5194/bg-12-
MathWorks, Inc., MathWorks. 4149-2015.
Birth, G.S., McVey, G.R., 1968. Measuring the Color of Growing Turf with a Reflectance Gamon, J.A., Surfus, J.S., 1999. Assessing leaf pigment content and activity with a re-
Spectrophotometer. flectometer. New Phytol. 143, 105–117. http://dx.doi.org/10.1046/j.1469-8137.
Calderón, R., Navas-Cortés, J.A., Lucena, C., Zarco-Tejada, P.J., 2013. High-resolution 1999.00424.x.
airborne hyperspectral and thermal imagery for early detection of Verticillium wilt of Govaerts, B., Verhulst, N., 2010. The normalized difference vegetation index (NDVI)
olive using fluorescence, temperature and narrow-band spectral indices. Remote Greenseeker (TM) handheld sensor: toward the integrated evaluation of crop man-
Sens. Environ. 139, 231–245. http://dx.doi.org/10.1016/j.rse.2013.07.031. agement. Part A-Concepts and case studies. Int. Maize Wheat Improv. Cent. 1–12.
Carpenter, G.A., Gopal, S., Macomber, S., Martens, S., Woodcock, C.E., Franklin, J., 1999. Holer, D.N.H., Dockray, M., Barber, J., 1983. The red edge of plant leaf reflectance. Int. J.
A neural network method for efficient vegetation mapping. Remote Sens. Environ. Remote Sens. 4, 273–288. http://dx.doi.org/10.1080/01431168308948546.
70, 326–338. http://dx.doi.org/10.1016/S0034-4257(99)00051-6. Hsu, K., Gupta, H.V., Sorooshian, S., 1995. Artificial neural network modeling of the
Carter, G.A., 1993. Responses of leaf spectral reflectance to plant stress. Am. J. Bot. 80, rainfall-runoff process. Water Resour. Res.
239–243. http://dx.doi.org/10.2307/2445346. Hunt, E.R., Doraiswamy, P.C., McMurtrey, J.E., Daughtry, C.S.T., Perry, E.M., Akhmedov,

116
M. Romero et al. Computers and Electronics in Agriculture 147 (2018) 109–117

B., 2012. A visible band index for remote sensing leaf chlorophyll content at the Roujean, J.L., Breon, F.M., 1995. Estimating PAR absorbed by vegetation from bidirec-
Canopy scale. Int. J. Appl. Earth Obs. Geoinf. 21, 103–112. http://dx.doi.org/10. tional reflectance measurements. Remote Sens. Environ. 51, 375–384. http://dx.doi.
1016/j.jag.2012.07.020. org/10.1016/0034-4257(94)00114-3.
Jones, H.G., 2007. Monitoring plant and soil water status: established and novel methods Rouse, J.W., Hass, R.H., Schell, J.A., Deering, D.W., 1973. Monitoring vegetation systems
revisited and their relevance to studies of drought tolerance. J. Exp. Bot. 58, in the great plains with ERTS. Third Earth Resour. Technol. Satell. Symp. 1, 309–317.
119–130. http://dx.doi.org/10.1093/jxb/erl118. https://doi.org/citeulike-article-id:12009708.
Kazmierski, M., Glemas, P., Rousseau, J., Tisseyre, B., 2011. Temporal stability of within- Samborska, I.A., Alexandrov, V., Sieczko, L., Kornatowska, B., Goltsev, V., Cetner, M.D.,
field patterns of ndvi in non irrigated mediterranean vineyards. J. Int. des Sci. la Kalaji, H.M., 2014. Artificial neural networks and their application in biological and
Vigne du Vin. 45, 61–73. agricultural research. J. NanoPhotoBioSciences 2, 14–30.
Khairunniza-Bejo, S., Mustaffha, S., Wan Ismail, I., 2014. Application of artificial neural Scholander, P.F., Hammel, H.T., Hemmingsen, E.A., Bradstreet, E.D., 1964. Hydrostatic
network in predicting crop yield: a review. J. Food Sci. Eng. 705, 283–291. http://dx. pressure and osmotic potential in leaves of mangroves and some other plants. Proc.
doi.org/10.1016/j.aca.2011.06.033. Natl. Acad. Sci. 52, 119–125.
Kim, M., Gilley, J.E., 2008. Artificial Neural Network estimation of soil erosion and nu- Smith, K. a, Mullins, C.E., 2000. Soil and Environmental Analysis. doi: 10.1201/
trient concentrations in runoff from land application areas. Comput. Electron. Agric. 9780203908600.
64, 268–275. Su, B.F., Xue, J.R., Xie, C.Y., Fang, Y.L., Song, Y.Y., Fuentes, S., 2016. Digital surface
Martí, P., Gasque, M., González-Altozano, P., 2013. An artificial neural network approach model applied to unmanned aerial vehicle based photogrammetry to assess potential
to the estimation of stem water potential from frequency domain reflectometry soil biotic or abiotic effects on grapevine canopies. Int. J. Agric. Biol. Eng. 9, 119–130.
moisture measurements and meteorological data. Comput. Electron. Agric. 91, http://dx.doi.org/10.3965/j.ijabe.20160906.2908.
75–86. http://dx.doi.org/10.1016/j.compag.2012.12.001. Sudheer, K.P., Gosain, A.K., Mohana Rangan, D., Saheb, S.M., 2002. Modelling eva-
Moya, I., Camenen, L., Evain, S., Goulas, Y., Cerovic, Z.G., Latouche, G., Flexas, J., Ounis, poration using an artificial neural network algorithm. Hydrol. Process. 16,
A., 2004. A new instrument for passive remote sensing. Remote Sens. Environ. 91, 3189–3202.
186–197. http://dx.doi.org/10.1016/j.rse.2004.02.012. Turner, N.C., 1988. Measurement of plant water status by the pressure chamber tech-
Penuelas, J., Pinol, J., Ogaya, R., Filella, I., 1997. Estimation of plant water concentration nique. Irrig. Sci. 9, 289–308. http://dx.doi.org/10.1007/BF00296704.
by the reflectance Water Index WI (R900/R970). Int. J. Remote Sens. 18, 2869–2875. Van Beek, J., Tits, L., Somers, B., Coppin, P., 2013. Stem water potential monitoring in
http://dx.doi.org/10.1080/014311697217396. pear orchards through worldview-2 Multispectral Imagery. Remote Sens. 5,
Pinar, A., Curran, P.J., 1996. Technical Note Grass chlorophyll and the reflectance red 6647–6666. http://dx.doi.org/10.3390/rs5126647.
edge. Int. J. Remote Sens. 17, 351–357. http://dx.doi.org/10.1080/ Van Leeuwen, C., Goutouly, J., Azaïs, C., Marguerit, E., Roby, J., Chone, X., Germain, C.,
01431169608949010. 2006. Intra-block variations of vine water status in time and space Variations intra-
Poblete, T., Ortega-Farías, S., Moreno, M., Bardeen, M., 2017. Artificial neural network to parcellaires temporelles et spatiales du régime hydrique de la vigne. Vine 64–69.
predict vine water status spatial variability using multispectral information obtained Van Leeuwen, C., Tregoat, O., Chon??, X., Bois, B., Pernet, D., Gaudill??re, J.P., 2009.
from an unmanned aerial vehicle (UAV). Sensors 17, 2488. http://dx.doi.org/10. Vine water status is a key factor in grape ripening and vintage quality for red bor-
3390/s17112488. deaux wine. How can it be assessed for vineyard management purposes? J. Int. des
Pôças, I., Rodrigues, A., Gonçalves, S., Costa, P.M., Gonçalves, I., Pereira, L.S., Cunha, M., Sci. la Vigne du Vin 43, 121–134. http://dx.doi.org/10.20870/oeno-one.2009.43.3.
2015. Predicting grapevine water status based on hyperspectral reflectance vegeta- 798.
tion indices. Remote Sens. 7, 16460–16479. http://dx.doi.org/10.3390/rs71215835. Waldo, L., Schumann, A., 2009. Alternative methods for determining crop water status for
Qi, J., Chehbouni, A., Huete, A.R., Kerr, Y.H., Sorooshian, S., 1994. A Modified Soil irrigation of citrus groves. Soil Water 63–71.
Adjusted Vegetation Index 126, 119–126. http://dx.doi.org/10.1016/0034-4257(94) Williams, L.E., Araujo, F.J., 2002. Correlations among Predawn Leaf, midday leaf, and
90134-1. midday stem water potential and their correlations with other measures of soil and
Qiao, D.M., Shi, H.B., Pang, H.B., Qi, X.B., Plauborg, F., 2010. Estimating plant root water plant water status in Vitis vinifera. J. Am. Soc. Hortic. Sci. 127, 448–454. http://dx.
uptake using a neural network approach. Agric. Water Manag. 98, 251–260. doi.org/10.1016/j.yqres.2009.07.010.
Rallo, G., Minacapilli, M., Ciraolo, G., Provenzano, G., 2014. Detecting crop water status Wu, S.G., Bao, F.S., Xu, E.Y., Wang, Y.-X., Chang, Y.-F., Xiang, Q.-L., 2007. A leaf re-
in mature olive groves using vegetation spectral measurements. Biosyst. Eng. 128, cognition algorithm for plant classification using probabilistic neural network, in:
52–68. http://dx.doi.org/10.1016/j.biosystemseng.2014.08.012. Signal Processing and Information Technology, 2007 IEEE International Symposium
Riad, S., Mania, J., Bouchaou, L., Najjar, Y., 2004. Rainfall-runoff model usingan artificial on. IEEE, pp. 11–16.
neural network approach. Math. Comput. Model. 40, 839–846. http://dx.doi.org/10. Zarco-Tejada, P.J., Berjón, A., López-Lozano, R., Miller, J.R., Martín, P., Cachorro, V.,
1016/j.mcm.2004.10.012. González, M.R., de Frutos, A., 2005. Assessing vineyard condition with hyperspectral
Richardson, A.J., Everitt, J.H., 1992. Using spectral vegetation indices to estimate ran- indices: Leaf and canopy reflectance simulation in a row-structured discontinuous
geland productivity. Geocarto Int. 7, 63–69. http://dx.doi.org/10.1080/ canopy. Remote Sens. Environ. 99, 271–287. http://dx.doi.org/10.1016/j.rse.2005.
10106049209354353. 09.002.
Rodríguez-Pérez, J.R., Riaño, D., Carlisle, E., Ustin, S., Smart, D.R., 2007. Evaluation of Zarco-Tejada, P.J., González-Dugo, V., Williams, L.E., Suárez, L., Berni, J.A.J.,
hyperspectral reflectance indexes to detect grapevine water status in vineyards. Am. Goldhamer, D., Fereres, E., 2013. A PRI-based water stress index combining struc-
J. Enol. Vitic. 58, 302–317. tural and chlorophyll effects: assessment using diurnal narrow-band airborne imagery
Rondeaux, G., Steven, M., Baret, F., 1996. Optimization of Soil-Adjusted Vegetaion and the CWSI thermal index. Remote Sens. Environ. 138, 38–50. http://dx.doi.org/
Indices. 10.1016/j.rse.2013.07.024.
Rossini, M., Fava, F., Cogliati, S., Meroni, M., Marchesi, A., Panigada, C., Giardino, C., Zygielbaum, A.I., Gitelson, A.A., Arkebauer, T.J., Rundquist, D.C., 2009. Non-destructive
Busetto, L., Migliavacca, M., Amaducci, S., Colombo, R., 2013. Assessing canopy PRI detection of water stress and estimation of relative water content in maize. Geophys.
from airborne imagery to map water stress in maize. ISPRS J. Photogramm. Remote Res. Lett. 36. http://dx.doi.org/10.1029/2009GL038906.
Sens. 86, 168–177. http://dx.doi.org/10.1016/j.isprsjprs.2013.10.002.

117

You might also like