You are on page 1of 11

Ecological Indicators 111 (2020) 105991

Contents lists available at ScienceDirect

Ecological Indicators
journal homepage: www.elsevier.com/locate/ecolind

Rainfed wheat (Triticum aestivum L.) yield prediction using economical, T


meteorological, and drought indicators through pooled panel data and
statistical downscaling

Nasrin Salehniaa,c, Narges Salehniab, , Ahmad Saradari Torshizib, Sohrab Kolsoumic
a
Faculty of Agriculture, Department of Water Engineering, Ferdowsi University of Mashhad, Mashhad, Iran
b
Department of Economics, Ferdowsi University of Mashhad, P.O. Box 9177-948951, Mashhad, Iran
c
AgriMetSoft, Roshd Center, Ferdowsi University of Mashhad, P.O. Box 9177-949207, Mashhad, Iran

A R T I C LE I N FO A B S T R A C T

Keywords: Agriculture productions play significant roles in economic development. Extreme weather events, especially
Precipitation drought under climate change conditions, can affect future crop production. Nowadays, researchers are trying to
SPEI apply modeling approaches for estimating future changes on amounts of crop yields. This study employed pooled
SPI panel data to simulate the most effective meteorological drought indices, economic and meteorological variables
Diurnal temperature
on rainfed wheat yield. The observation period was 1990–2016 for several meteorological data, besides SPI
Durbin Watson test
Chow test
(Standardized Precipitation Index) and SPEI (Standardized Precipitation Evapotranspiration Index) drought
indices in monthly and yearly scales. The available economic variables during the study period were yearly
guaranteed wheat prices (Rial/kg) and area under cultivation (ha). In this research, first, the most effective
variables were selected according to the efficiency criteria and stepwise regression. Then by using pooled panel
data, a relation was estimated between yield and the independent variables. Finally, with future downscaled
variables, the amount of wheat yield was determined for the next 20 years (2019–2038). The GFDL- ESM2M and
MIROC5 models under RCP45 and RCP85 were run, and MIROC5 under RCP45 was selected as the best model,
for the evaluation period. The results revealed that guaranteed wheat prices, yearly precipitation and sunshine
hours, the area under cultivation, and SPI of October were identified as the most effective variables on wheat
yield through the Panel model. By using the projection weather variables and the pooled panel model, we
achieved that the amount of rainfed wheat yield would be increased over two next decades at Mashhad,
Sabzevar, and Torbat H. locations.

1. Introduction 2007). Agriculture in the northeast of Iran has been usually affected by
drought events. Many researchers have assessed the possible impact of
Climate change and global warming have significant impacts on them on crop yields using basic simulation models and statistical
plant growth and crop yield productivity all over the world. One of the methods, and most recent assessments have yielded that arid and semi-
most critical factors causing the yield variability is climate (Chaves, arid regions are highly vulnerable to climate change (IPCC, 2014).
2001), and climate change will shift growing conditions and thus im- In arid and semi-arid areas, such as Iran which average annual
pact future crop production. Agriculture plays a crucial role in eco- precipitation is approximately 250 mm, crops are mostly rainfed, and
nomic development and poverty reduction (World Bank, 2005). Con- the sowing date of plants depends on the rain; hence rainfall variability
siderable attention has been dedicated to agricultural effects of climate is critical for agricultural yields. In this country, rainfed agriculture is
change. As a consequence of climate change, drought usually issues likely to be the most sensitive sector to climate change (Nassiri et al.,
from precipitation deficit. 2006). Wheat and rice are Iran’s two main food staples. Wheat is har-
Drought is a severe and complicated natural hazard that sig- vested on more than 6 million hectares; of these, about 2.170 million
nificantly impacts agricultural yields, and its occurrence will adversely hectares are irrigated, and approximately 3.830 million hectares are
affect wheat yield at the critical growth stages of wheat such as til- rainfed (Agriculture Statistics, 2013).
lering, jointing, anthesis, and grain filling (Xue et al., 2006; IPCC, Wheat has traditionally been one of the three most essential crops


Corresponding author.
E-mail addresses: salehnia61@gmail.com (N. Salehnia), n.salehnia@um.ac.ir (N. Salehnia).

https://doi.org/10.1016/j.ecolind.2019.105991
Received 22 May 2019; Received in revised form 25 November 2019; Accepted 4 December 2019
Available online 16 December 2019
1470-160X/ © 2019 Published by Elsevier Ltd.
N. Salehnia, et al. Ecological Indicators 111 (2020) 105991

Fig. 1. The study area with three synoptic stations located in Khorasan-e Razavi province, northeastern Iran, Middle East, with an area of 118,851 km2.

(corn, rice, and wheat) used in the world (FAOSTAT, 2010; Zhou, prefectures over the 1977–1996 period, have been examined by Chen
2010). It is crucial to simulate the yield of wheat under various stress and Chang (2005). The results of Chen and Chang (2005) not only il-
conditions like high air temperature and low precipitation to prepare lustrated the sensitivity of crop yield distributions in response to
response strategies to future climate change. The issue of how to in- weather changes in this area but also identified the atmospheric con-
crease wheat yields has been a critical research question for agrono- ditions, which control the yield distributions. The determinants of
mists and farmers. Crop yield is a function of weather variables and wheat output per hectare in Ethiopia during 2011– 2013 by using a
economic factors (Cabas et al., 2010). To investigate the meteorological panel data approach were explored. The model identified specific
and economic variables affecting rainfed wheat yield, different ap- contributors to wheat yields that include farm management techniques,
proaches, models, and methods have been applied. Several simulation weather (e.g., rainfall), water availability, and policy intervention
models have been developed for estimating wheat yield, such as BISm (Mann and Warner, 2017).
model by Snyder et al. (2004), AquaCrop developed by FAO (FAO, This study focuses on determining the most effective meteorological
2012), WOFOST, and DSSAT by Jones et al., 2003. Different meth- and economic variables on rainfed wheat yield. In this regards, three
odologies for simulating impacts of climate change on crop production specific goals were identified: 1) associations between drought indices,
were revealed by White et al. (2011), they reviewed 221 peer-reviewed meteorological-economic variables, and wheat yield in northeast of Iran
papers that used crop simulation models to examine diverse aspects of known as Khorasan-e Razavi province, through pooled panel data, 2)
how climate change might affect agricultural systems. Finally, they comparison of the performance of statistical downscaling method over
could categorize 221 types of researches to six subject areas namely, two CMIP5 models under two RCPs scenarios during the evaluation
target crops and regions; the crop model(s) used and their character- period (2006–2016), then selecting the best CMIP5 model under the
istics; sources and application of data on [CO2] and climate; impact RCP scenario for each station and obtaining the future projections of
parameters evaluated; assessment of variability or risk; and adaptation the selected weather variables, and 3) assessment future wheat yield
strategies. amount over the next decades through the selected CMIP5 model and
Many other studies attempted to determine the most effective the chosen RCP scenario.
weather variables on crop yields, and predicted the amount of crop
yields, under climate change conditions. We have summarized some of 2. Materials and methods
these researches which have implemented in all different countries
around the world (Supplementary File 1). 2.1. Study sites and weather data
The primary model used in this study was Panel data. Panel data
have become widely available in both the developed and developing In this study, we examined the climate and economic variables af-
countries. Panel data refer to a data set containing observations of fecting wheat yield at three major agricultural locations (Mashhad,
various phenomena over multiple periods (Chen and Chang, 2005; Sabzevar, and Torbat Heydarieh (Torbat H.)) in Khorasan-e Razavi
Hsiao, 2014). Unlike regression data, with panel data, we observe many province located in northeastern Iran (Fig. 1). This region has a semi-
subjects over time, and unlike time-series, with panel data we find arid climate and lies between 33° and 38°N latitude and 55° and 61°E
many topics, therefore, it has two dimensions: spatial (cross-sectional) longitude, with an area of 118,851 km2. Table 1 presents the phy-
and temporal (time series) (Darmofal, 2015). This concept is performed siocharacteristics of the study locations, during the baseline weather
in the various field of researches, especially in crop science. Coelli record (1990–2016). The data were collected from the meteorological
(1996) used panel data on rice production in Vietnam to investigate the station at each location. Homogenization and quality control of weather
total factor productivity (TFP) growth. The impact of weather on the data were performed by the national meteorological organization of
yields of seven major crops in Taiwan based on pooled panel data for 15 Iran (www.weather.ir) before the release of such data to users.

2
N. Salehnia, et al. Ecological Indicators 111 (2020) 105991

Table 1 amount was obtained as Eq. (1):


The annual average of weather variables during the baseline period
0.408 × Δ × (Rn − G ) + γ × [(890 T + 273)] × U2 × VPD
(1990–2016) for the study sites (http://www.irimo.ir/). ETO =
Δ + γ × (1 + 0.34 × U2) (1)
Location LA LN Elev. Total Ave. Ave. Ave.
(m) Pre. (mm) Tmean Wind Humidity where ETO is the reference crop evapotranspiration (mm/day). Δ shows
(°C) speed (%) the slope of the vapor pressure curve (kPa°C−1), Rn is the net radiation
(m/s)
at the crop surface (MJM−2d−1). The variable G is the soil heat flux
Mashhad 36°16′ N 59°38′ E 999 246 15.1 1.2 51 density (MJM−2d−1), γ is the psychometric constant (kPa°C−1), 890
Sabzevar 36°12′ N 57°43′ E 972 191 17.7 1.6 40 refers to the conversion factor, T is the average air temperature at 2 m
Torbat H. 35°16′ N 59°13′ E 1451 257 14.1 1.2 45 height (°C), U2 presents the wind speed measured at a 2 m height
(ms−1), and VPD is the vapor pressure deficit (kPa).
LA: latitude, LN: longitude, Elev.: elevation, Ave.: Average, Tmean: mean
For computing SPEI, we used the classical approximation of
temperature (°C), Pre.: Precipitation.
Abramowitz and Stegun (1965) and also Vicente-Serrano et al. (2010).
This index is accumulated at different time scales. The SPEI is based on
Historical wheat yield (kg/ha), the area under cultivation (ha), and
a monthly climatic water balance (precipitation minus ETO), which is
guaranteed wheat prices (Rial/kg) which are produced under rainfed
adjusted using a three-parameter log-logistic distribution. The deficit or
conditions during 1990–2016, for the three locations were obtained
surplus accumulation of a climate water balance (D) at different time
from Iran’s Ministry of Agriculture (https://www.maj.ir/). According to
scales is determined by the difference between the precipitation (P) and
the climate of the case study, the past reviewed researches
ETO for the day i:
(Supplementary File 1), and available reports of meteorology organi-
zation of Iran (http://www.irimo.ir/), some weather variables play an Di = Pi − EToi (2)
essential role in the amount of wheat yield, so in this research we
The probability density function of a three-parameter log-logistic
collected these data, namely monthly weather data included the max-
distributed variable is expressed as Eq. (3):
imum and minimum temperature (◦C), diurnal temperature range (DTR
−2
(◦C)), sunshine hours (Nhour) (h), yearly total precipitation (mm), as β x − γ β−1 ⎡ x − γ β⎤
f (x ) = ⎛ ⎞ × 1+⎛ ⎞
well the monthly amount of precipitation (mm), relative humidity (%), α⎝ α ⎠ ⎢ ⎝ α ⎠ ⎥
⎣ ⎦ (3)
wind speed (m/h), and mean temperature (◦C) for Jan, May, and Oct
during 1990–2016. where α , β , and γ are scale, shape, and origin parameters, respectively,
The DTR is the difference between the daily maximum and for D values in the range (γ > D < ∞). The parameters of the Pearson
minimum temperatures (DTR = Tmax-Tmin). It is often used in climate III distribution can be obtained as Eqs. (4)–(7):
studies since the only use of mean temperature cannot reveal very 2w1 − w0
significant temperature change patterns by averaging the trends β=
6w1 − w0 − 6w2 (4)
(Holder et al., 2006). The SPI (Standardized Precipitation Index) was
calculated over Jan, May, and Oct and SPEI (Standardized Precipitation (w0 − 2w1) β
α=
Evapotranspiration Index) was estimated at a yearly scale. For calcu- Γ(1 + 1 β )Γ(1 − 1 β ) (5)
lating the SPEI, we used ETo (reference crop evapotranspiration).
Therefore the FAO Penman-Monteith method (Allen et al., 1998) was 1 + 1⎞ ⎛1 − 1⎞
γ = w0 − α Γ ⎛⎜ ⎟Γ⎜ ⎟
used (see description below). All the input variables are presented in ⎝ β ⎠ ⎝ β ⎠ (6)
Table 2. To analyze future wheat yield, the period 2019–2038 was di-
vided in two decades, namely 2019–2028, and 2029–2038. where Γ (β ) is the gamma function of β . The probability distribution
function of the D series is presented by:
β −1
⎡ α ⎞ ⎤
2.2. Calculation of the SPEI, SPI, and evapotranspiration F (x ) = ⎢1 + ⎜⎛
γ⎠ ⎥

⎝ x − (7)
⎣ ⎦
SPEI is a meteorological drought index that consists of the pre-
In the last step, with the value of F(x), the SPEI can be estimated as
cipitation data (mm) and the evapotranspiration data (mm/day). The
the standardized values of F(x). The SPEI equation is calculated by Eq.
evapotranspiration rate from a reference surface is called the reference
(8):
crop evapotranspiration or reference evapotranspiration and is denoted
as ETo. We followed the FAO Penman-Monteith approach to calculate C0 + C1 W + C2 W 2
SPEI = W −
the reference evapotranspiration (ETO) (Allen et al., 1998). The ETo 1 + d1 W + d2 W 2 + d3 W 3 (8)
where W = −2Ln (P ) for P ≤ 0.5 and P is the probability of exceeding
Table 2
All the input variables at three stations. a determined D value. The constants are C0 = 2.515517,
C1 = 0.802853, C2 = 0.010328, d1 = 1.432788, d2 = 0.189269 and
Variable Scale Period d3 = 0.001308.
Precipitation (mm) Monthly (May, Jun., and Oct.) 1990–2016 The SPI is widely used as a meteorological drought index (Angelidis
-Yearly et al., 2012). Since continuous long-term data of at least 30 years is
Wind speed (m/s) Monthly (May, Jun., and Oct.) 1990–2016 required to compute SPI (Mckee et al., 1993), therefore we used
Relative Humidity (%) Monthly (May, Jun., and Oct.) 1990–2016 30 years (1987–2016) of precipitation over case study area for calcu-
Mean of temperature (°C) Monthly (May, Jun., and Oct.) 1990–2016
SPI Monthly (May, Jun., and Oct.) 1990–2016
lating it, then we have separated the monthly SPI data over 1990–2016.
DTR (°C) Yearly 1990–2016 SPI is usually computed by fitting the gamma probability distribution
Nhour (hour) Yearly 1990–2016 (Edwards and McKee, 1997) to the observed precipitation data, as
SPEI Yearly 1990–2016 follows:
The area under cultivation (ha) Yearly 1990–2016
Guaranteed wheat prices (Rial/ Yearly 1990–2016 1
G (x ) = x α − 1e−x β (x > 0)
kg) β α Γ(α ) (9)
DTR: Diurnal Temperature Range. where x > 0 is the amount of precipitation, α > 0 is a shape

3
N. Salehnia, et al. Ecological Indicators 111 (2020) 105991

parameter, β > 0 is a scale parameter, and, Γ(α) is the gamma func-

The related information, the details of characteristics of each model, and the processes for downloading original models are available at: http://www.ipcc-data.org/sim/gcm_monthly/AR5/Reference-Archive.html.
tion, and α, β, and Γ(α) computed as following (Eqs. (10) and (11)),
where n is the number of observations:

Horizontal resolution

1 ⎛ 4A ⎞ x − ∑ ln(x )
α= ⎜1 + 1+ ⎟, β = , whereA = ln(x ) −
4A ⎝ 3 ⎠ α n (10)

(lat × lon)

1.40*1.40

2.5*2.0
Γ(α ) = ∫0 y α − 1e−ydy (11)
SPI values range from −2 (extremely dry) to +2.0 (extremely wet),

Grid size

256*128
also SPI is in “normal condition” when 0.99 > SPI > −0.99. Details

144*90
about the SPI computation can be found in McKee et al. (1993), and
Guttman (1999). For computing the drought indices, the Drought

Model abbreviation
Monitor and Prediction (DMAP) software tool (AgriMetSoft, 2018a,b)
was used.

GFDL-ESM2M

MIROC5
2.3. GCMs and statistical downscaling method

We applied statistical downscaling for finer resolution of the

NOAA GFDL
models. In this study, we used the Delta method (Bennett et al., 2014).

Institute ID
The Delta technique was selected because it is widely used for down-

MIROC
scaling GCMs outputs (Maraun et al. 2010; Themeßl et al. 2011). Delta
method was calculated through Eqs. (12) and (13), as follows:

The Model for Interdisciplinary Research on Climate


μ (PObs )
PPr . = PGCMrcp ×

NOAA/Geophysical Fluid Dynamics Laboratory


μ (PGCMhist ) (12)

Model with MOM4 ocean component (USA)


TPr . = TGCMrcp + μ (TObs ) − μ (TGCMhist ) (13)
where T and P are the values of temperature and precipitation, re-
spectively. The subscript rcp refers to the GCM’s RCP outputs over the
future period, and subscript Obs represents the observed values of the
synoptic station. GCMhist represents the data of historical run of the
CMIP5 used model. Subscripts Obs, and Pr. refer to observation, and
prediction data, respectively, and μ is the mean of data during the in-
Earth System

tent period (Yazd et al., 2019).


Model name

The used CMIP5 models in this study were GFDL-ESM2M


(2.02° × 2.5°) and MIROC5 (1.4008° × 1.4062°). The GCMs models
under different RCP scenarios can download from http://www.ipcc-
data.org/sim/gcm_monthly/AR5/Reference-Archive.html. The brief
The University of Tokyo, National Institute for Environmental Studies, and Japan Agency for Marine-
National Oceanic and Atmospheric Administration (NOAA) Geophysical Fluid Dynamics Laboratory

information’s models are presented in Table 3. The models were run


under the Representative Concentration Pathway (RCP) scenarios,
RCP4.5 and RCP8.5, which refer to radiative forcing in 2100 of 4.5 and
8.5 Wm−2, and CO2 equivalent mixing ratios of 650 and 1370 ppm,
respectively (Moss et al., 2010).

2.4. Panel data and stepwise method

The model for rainfed wheat yield is specified with a logarithmic


functional (Lobell and Burke, 2010) form as Eq. (14):

lnYit = β0 + β1 lnx1it + β2 lnx 2it + β3 lnx 3it + ⋯+βn lnxnit + Uit i


= 1.0. 3; and t = 1990⋯2016 (14)
where the dependent variable for estimation rainfed wheat yield for
each station, in the panel data model is Yit . Explanatory variables are
x1it , x2it , x3it (refer to all the independent variables), i is the number of
stations, and a constant value as Uit . The random disturbance term, Uit
Earth Science and Technology

has mean 0 and is assumed uncorrelated with the independent vari-


Defined details of used GCMs.

ables. Parameters β1, β2, β3, and βn are to be estimated, and β0 re-
presents the model intercept. Data for three stations covering the
Modeling center (group)

1990–2016 period results in 81 (3 (number of stations) * 27 (number of


years)) observations.
In this research, all the variables were assessed through stepwise
regression (Salehnia et al., 2019) for identifying the ones with greater
importance. Then selected variables were analyzed by Panel data
Table 3

method for choosing the most effective ones. Stepwise regression is a


method of regression modeling which can be applied to predict a

4
N. Salehnia, et al. Ecological Indicators 111 (2020) 105991

dependent variable, using a set of independent variables (Lobell and period (2006–2016), the prediction performance of the models was
Burke, 2010; Das et al., 2018) (21 variables in Table 2). Linear re- compared based on the mean absolute error (MAE) (MacLean, 2005),
gression is performed by adding or removing independent variables on and Nash–Sutcliffe efficiency (NSE), as defined in Eqs. (21) and (22).
each iteration. Stepwise regression procedure was used for the selection The mean absolute error (MAE) Nash–Sutcliffe efficiency (NSE),
of significant variables (O’Gorman and Woolson, 1991).
n
1
MAE =
n
× ∑ |Oi − Pi|
2.5. Statistical tests i=1 (21)

2.5.1. Significance test for coefficient n


∑i = 1 (Oi − Pi )2
To perform significant test, we compute the t-statistic for each of the NSE = 1 − −
n
coefficients that derive from the ratio of the coefficient to its standard ∑i = 1 (Oi − O )2 (22)
deviation, and its value was compared with t(α, n-k), where α is the
confidence level. where n is the number of samples, P is the predicted weather value, and
O is the observed weather data value. The MAE is the average of all
βk absolute errors and represents the average differences between the
tn − k =
se (β ) k (15) model predicted and observed values. The NSE is a normalized statistic
that shows the proportion of explained variance relative to the 1:1 line
in a comparison of two estimates of the same variable. The range of NSE
2.5.2. Chow test lies between 1.0 (perfect fit) and −∞.
The Chow test determines if the regression coefficients are different Also, we applied the Taylor diagram, which can represent three
for split data sets. It tests whether one regression line or two separate different statistics simultaneously (i.e., the centered RMS (Root Mean
regression lines best fit a split set of data. If the two parts can be re- Square) difference, the Pearson correlation, and the standard devia-
presented by one single regression line, we say that the regression can tion). Taylor diagrams (Taylor, 2001) provide a way of graphically
be pooled (Chow, 1960). With two sets of observations: group1 and summarizing how closely a model matches observations. Through the
group2, consider two regression equations (Eqs. (16) and (17)) as fol- Taylor diagram, we can decide which of the models and scenarios are
lows: the proper one. Besides Taylor diagrams, we utilized boxplot. Boxplot
Group1(samplesizeofn1): y = α1 + β11 x1 + β12 x2 + ⋯+β1k xk + u (16) presents future data versus the observation data. This type of graphical
representation was selected because, in addition to the median, the
Group2(samplesizeofn2): y = α2 + β21 x1 + β22 x2 + ⋯+β2k xk + v (17) boxplot depicts the extreme values (minimum and maximum) of data.
In the boxplot, the Interquartile Range (IQR, i.e. 25th–75th quantiles) is
The null hypothesis of the model is
represented by boxes, the whiskers extend to quantile 5 and quantile
H0: β11 = β21, β12 = β22, ⋯, β1k = β2k , α1 = α2 . The residual sums of
95, the black horizontal line indicates the median, and values beyond
squares for Group1 and Group2 respectively refer to RSS1 and RSS2,
the whiskers are outliers. For drawing boxplots, Data Tool (https://
RSS1 + RSS2 = URSS, and k is the number of regressors, n1 and n2 are
agrimetsoft.com/data-tool.aspx) was used. All the calculations and
respectively the numbers of observations in the first of the two sub-
statistical analyses in this research were computed by using EViews
samples, and the other sub-samples. The restricted Residual Sums of
version 11 and Stata version 14 software tools.
Squares (RRSS) being that of the OLS (ordinary least squares) on the
pooled model and the Unrestricted Residual Sums of Squares (URSS)
being that of the LSDV (Least Squares Dummy Variable) regression. 3. Results and discussion
When the null hypothesis is correct, the F-ratio was performed ac-
cording to the Eq. (18), as Chow-test (Chow, 1960): 3.1. Assessing the results of selecting the effective variables
(RRSS − URSS ) (k + 1)
Chow = Out of the 21 variables analyzed in Table 2, six variables showed
URSS (n1 + n2 − 2k − 2) (18)
significant correlation with yield (P < 0.05) by using stepwise re-
gression. These significant variables were further subjected to stepwise
2.5.3. Breusch-Pagan (BP) test regression analysis and those variables, which could be related to yield
The BP (Breusch and Pagan) test is used to check for hetero- variability reasonably (Precipitation (mm), Nhour, SPI of Oct., mean of
scedasticity in a linear regression model (Breusch and Pagan, 1979). It temperature of Oct. (°C), area under cultivation (ha), and guaranteed
is based on models of the type Eq. (19) for the variances of the ob- wheat prices (Rial/kg)) were used in the Panel data model.
servations, wherez rt explain the difference in the variances. The null
hypothesis is equivalent to the (p − 1) parameter restrictions
δ1 = ⋯=δr = 0 , and α is the confidence level. 3.2. Analyzing the results of statistic tests

δt2 = E (ut2) = h (α 0 + α1 z1t +⋯+αr z rt ) (19)


Table 4 shows the result of chow test, as can be seen in this table,
the zero hypothesis of the pooled data is accepted at the 5% error level
2.5.4. Durbin Watson test and therefore the research data should be estimated by the pooled
The Durbin Watson is a test where the residuals from a linear re- method. It means that the spatial and temporal dimensions of data can
gression or multiple regression are independent. The test statistic is be omitted and the estimation is satisfied by OLS. In this case, the ex-
presented in Eq. (20), as following, where n is the number of observa- planatory variables are detailed as they include all of the cross-section
tions et are residuals from an ordinary least squares regression: characteristics in a single case study.
n
∑t = 2 (et − et − 1)2 Table 4
d= n
∑t = 2 (et )2 (20) The result of the Chow test.
Effects Test Statistic d.f. Prob.

2.5.5. Efficiency criteria Cross-section F 0.773558 (2,72) 0.4652


For selecting the best model and scenario during the evaluation

5
N. Salehnia, et al. Ecological Indicators 111 (2020) 105991

Table 5 two models under two RCPs, while precipitation showed a much lower
The coefficients and statistical criteria from the Panel data model. Dependent Pearson correlation (0.29 up to 0.55). In general, the best model for
Variable: LNY, Method: Panel Least Squares, Sample: 1990 2016, Periods in- Nhour and precipitation was MIROC5 under RCP4.5 with the Delta
cluded: 27, Cross-sections included: 3, Total panel (balanced) observations: 81. method, due to both of them had the best standard deviation (0.99 and
Variable Coefficient Std. Error t-Statistic Prob. 0.94) and RMS (1.4 and 0.73 mm) values compared with the other
cases, respectively.
C 10.65775 5.269710 2.022456 0.04670*
As well, in Sabzevar (Fig. 2(b)) and Torbat H. (Fig. 2(c)), the best
LNP 0.103875 0.038625 2.689336 0.0088**
LNPre 0.574854 0.136955 4.197391 0.0001**
model was MIROC5 under RCP4.5, for both weather variables (Nhour
LNNhour −1.191619 0.641554 −1.857395 0.067*** and precipitation). Indeed, for Nhour values in Sabzevar, the range of
LNCult 0.096387 0.055896 1.724405 0.088*** Pearson correlation was about 0.87–0.90, NSE = 0.57–0.59,
LNSPIOct −0.442108 0.118708 −3.724335 0.0004** MAE = 1.19–1.23 h, RMS = 1.44–1.48 h, and the best model was
LNTmeanOct 0.195143 0.337535 0.578143 0.56412
MIROC5 under RCP4.5. Whereas, for precipitation in the selected
R-squared 0.519420 Mean dependent 5.881358
var model (MIROC5 under RCP4.5) with observed data have presented the
Adjusted R- 0.475589 S.D. dependent var 0.488735 Pearson correlation = 0.40, RMS = 0.62 mm, and MAE = 0.40 mm.
squared For Torbat H. location (Fig. 2(b)), the best criteria’s for precipitation
S.E. of regression 0.373622 Akaike info 0.951312
were consist of NSE = 0.13, RMS = 0.7 mm, MAE = 0.47 mm, and the
criterion
Sum squared resid 10.32992 Schwarz criterion 1.158240
Pearson correlation = 0.56, and for Nhour were NSE = 0.64,
Log-likelihood −31.52813 Hannan-Quinn 1.034334 RMS = 1.52 h, MAE = 1.24 h, and the Pearson correlation = 0.89.
criteria. However, GFDL-ESM2M did not show a good representation (under
F-statistic 10.48166 Durbin-Watson stat 2.042527 both RCPs) of precipitation and Nhour data, in comparison to MIROC5.
Prob (F-statistic) 0.000000

* Significant at 5% level. 3.4. Future projections of precipitation, sunshine hours, and SPI
** Significant at 1% level.
*** Significant at 10% level. As shown in Fig. 3, for tracking the changes of total monthly pre-
cipitation and average monthly Nhour, these variables were plotted
3.2.1. Panel data model results during the recent 20 years (1997–2016) against 20 years in the future
The Eq. (23) developed by using the Panel data model, relative to period (2019–2038). In Fig. 3(a), in Mashhad station, there was not any
this equation, the coefficients are presented in Table 5. The results of remarkable change in Jan. and Feb., but there were noticeable differ-
Table 5 indicate that among of six variables only the last one (Tmea- ences in the other months, as, during summer and fall, the average
noct) did not show a significant relationship with yield (P > 0.05, and amount of Nhour would be decreased in both seasons with −7% and
relative to the significance test for coefficient), in Panel data model. So, −17%, respectively. Whereas during Mar. and spring, the amount of
we deleted this variable for the rest of analyzing. Nhour would increase by at least 14 h (27% and 12%, respectively)
compared to the baseline average Nhour, 20 years later. In the Sabzevar
lnYit = β0 + β1 lncultit + β2 lnPreit + β3 lnPit + β4 lnNhourit + β5 and Torbat H. stations, there was a similar condition in this variable
lnSPIoctit + β6 lnTmeanoctit + Uit i = 1⋯3; t = 1990⋯2016 (Fig. 3(c) and (e)). The whole amount of Nhour’s changes during
2019–2038 against baseline period had very slight variations (around
(23)
0), and it would not be changed in yearly scale, but there was a shift in
The Breusch-Pagan/Cook-Weisberg test for heteroscedasticity was the monthly scale in the amounts of Nhour, over the three stations
revealed that Ho: Constant variance, chi2 (6) = 4.87, and (Fig. 3).
Prob > chi2 = 0.5606. So, in this model, the H0 hypothesis is ac- For total precipitation, during 2019–2038 the amount of pre-
cepted; that is, the variation is identical. As well, the result of the cipitation will increase over spring, end of summer and in Oct., in three
Durbin Watson test (2.0425) was confirmed that the lack of self-cor- stations, with MIROC5 under RCP45 (Fig. 3(b), (d), and (f)). The sea-
relation (Table 5). sonal changes project the most significant increase in spring and Oct.,
Overall, regarding the results of Eq. (23) with increasing the guar- and the smallest in summer. For example, these changes in Apr. will be
anteed wheat prices, precipitation and area under cultivation, the around 56%, 80%, and 64%, respectively for Mashhad, Sabzevar, and
wheat yield would improve, but the increase in the Nhour, and the Torbat H., for 20 future years. Also, in Oct., during 20 years later, there
SPIoct. would decrease the amount of yield. The average temperature of will be incremental changes in total precipitation, while these changes
Oct. also would not have a significant effect on the yield amount. The are 150 mm, 140 mm, and 152 mm, respectively for Mashhad, Sab-
results show that with the assumption of other factors being constant, zevar, and Torbat H. Overall, 10%, 4%, and 8% increase in precipita-
one percent increase in guaranteed wheat prices, yearly precipitation tion relative to the baseline period for Mashhad, Sabzevar, and Torbat
and area under cultivation then wheat yield increased by 0.1%, 0.57%, H., respectively, using MIROC5 under RCP45. In general, the total
and 0.09% respectively. Increased one percent in Nhour and SPIoct yearly amount of precipitation will increase by around 8% in the two
respectively reduces 1.19% and 0.44% in wheat yield. In addition, the future decades in compare to the 20-years of the baseline period. In the
result of F-statistics reveals that the whole regression is significant; as climate of study area (arid and semi-arid), the rainfall in October, April,
this model is a micro model, the adjusted R-squared (0.48) is accep- and May is very beneficial for wheat growth, so through MIROC5′s
table, too. Akaike, Schwarz and Hannan-Quinn criteria represent that projections, it seems that the amount of rain-fed wheat will increase
the lagged form of the variables should be short-run. during 2019–2038.
For the Mashhad station (Fig. 4(a)), the projection IQR, median
3.3. Statistical downscaling and selecting the best model (−0.23) and mean (0.02) of future SPI is lower than the baseline
period, so it seems that the variation and deviation of SPI in Oct during
For selecting the best model and scenario during the evaluation 20 years later will be lower than the baseline period, whereas the ex-
period (2006–2016), a Taylor diagram was drawn. Fig. 2(a) illustrates treme wet and extreme drought conditions will be higher than the
the monthly evaluation of the precipitation and Nhour over Mashhad baseline period. For the Sabzevar station (Fig. 4(b)), the IQR is shrunk
station. There was a high correlation using observed and downscaled and with a strong skewness to the lower values of SPI – according to the
projected data analysis for Nhour (0.87 up to 0.91), the ranges of NSE median- and presents a stronger shift to extreme drier conditions, in
and MAE were 0.75 up to 0.80 h, and 0.97 up to 1 h, respectively for comparison to the baseline period. The median of SPI data in future

6
N. Salehnia, et al. Ecological Indicators 111 (2020) 105991

Fig. 2. Taylor diagram of stations with two GCMs, under RCP4.5 and RCP8.5 with Delta method for precipitation and Nhour for evaluation period, (a) Mashhad, (b)
Sabzevar, and (d) Torbat H.

tends to wet condition than a baseline period, while the mean of SPI period), so in this Section (3.5), the simulated wheat yield was calcu-
tends to the normal condition. The depicted boxplot (Fig. 4(c)) for the lated by the Panel data model’s equation. These results are presented in
Torbat H. station shows the differences between future SPI and baseline Fig. 6 for three stations. The results indicate that in Mashhad station
one. Relative to the boxplots in Fig. 4(c) for future SPI, we can reveal (Fig. 6(a)), the amount of yield would increase in a first and second
that it is shifted to lower values, present a skewness to more negative decade (2019–2028 and 2029–2038) in comparison to the baseline
values, and a significant shrink of the IQR. Finally, in the future period decades (1997–2006 and 2007–2016), and overall mean wheat yields
(2019–2038), the boxplots of the three locations depict that minimum would be increased up to 28% with selected variables. Fig. 6(b) reveals
values of SPI in Oct. move to more negative and drier values of SPI, that in Sabzevar location, the Panel data model forecasts a possible
whereas the variations of IQRs tend to shorter values. It seems that the ~12% reduction in the wheat yield for the first future decade (320 kg/
severe and extreme drought would possibly increase in comparison to ha) relative to the first decade of baseline (360 kg/ha), but there is a
the baseline period. However, the occurrence of wet periods tends to noticeable increment around ~60% for the second decade, through
increase as the median future periods of SPI in Oct., further than zero MIROC5 under RCP45. For Torbat H., there is not any change for the
for three locations (Fig. 4). first future decade vs. baseline one, but in the second decade, there
would be possible an 18% increment for wheat yield amount, in com-
3.5. Simulation wheat yield in future parison to the second baseline decade, with MIROC5 under RCP45.
Consequently, the mean total wheat yields during the future period in
The results show that the Exponential plot has the best outcome, comparison to baseline periods possibly would increase to 30%, 20%,
since R2 = 0.99. Finally, for the future price, the achieved equation in and 10% respectively for Mashhad, Sabzevar, and Torbat H. locations.
Fig. 5 was performed for future years. MIROC5 projected the selected Eyshi Rezaie and Bannayan (2012) obtained similar outputs since they
weather variables under RCP45 (according to the result of evaluation stated that the rainfed wheat would increase 15% during 2010–2039 in

7
N. Salehnia, et al. Ecological Indicators 111 (2020) 105991

Fig. 3. Historical Observations (1997–2016) and downscaled prediction (2019–2038) data of average Nhour and total precipitation period over the three stations.

comparison with the baseline with HadCM3 model under the A2 sce- scenario, we yielded that total precipitation would probably increase in
nario, in Khorasan province. spring and early fall, and gradually decreased Nhour and shift its values
throughout the study region. These changes may lead to an increase in
4. Conclusion rainfed wheat yield, which is demonstrated through the prediction of
rainfed wheat yields by MIROC5 data under RCP45. In conclusion, we
In this study, we have assessed the potential impacts of weather showed that mean total wheat yields during the future period vs.
variables, drought indices, and economic variables on wheat yield at baseline periods would increase to 30%, 20%, and 10% respectively for
three major agricultural locations (Mashhad, Sabzevar, and Torbat H.) Mashhad, Sabzevar, and Torbat H. locations.
in northeastern Iran. We applied different approaches, such as Pooled
Panel data model and Delta statistical downscaling method for pre- CRediT authorship contribution statement
diction rainfed wheat yield. Through the achieved Panel’s equation
with selected downscaled weather data of future, the mean value of the Nasrin Salehnia: Writing - original draft, Methodology,
area under cultivation, and exponential relation for amount guaranteed Conceptualization, Software, Formal analysis, Writing - review &
wheat prices, the amount of future wheat yields were estimated for the editing, Investigation, Resources, Data curation, Validation,
three locations. According to the MIROC5 outputs under RCP45 Visualization. Narges Salehnia: Project administration, Supervision,

8
N. Salehnia, et al. Ecological Indicators 111 (2020) 105991

Fig. 4. The amount of SPI for Oct. during baseline period (1997–2016) versus future period (2019–2038), over three stations. The black horizontal line indicates the
median, and the multiplication sign (×) refers to the mean value.

Writing - review & editing, Funding acquisition. Ahmad Saradari Acknowledgment


Torshizi: Investigation. Sohrab Kolsoumi: Software.
This work was financially supported by the Ferdowsi University of
Mashhad, Iran with no. 41045, so the authors are thankful for the
Declaration of Competing Interest support of this university. As well, the authors are grateful for the
thoughtful and constructive comments provided by one the anonymous
The authors declare that they have no known competing financial reviewers, which prompted significant improvements to the manu-
interests or personal relationships that could have appeared to influ- script.
ence the work reported in this paper.

Fig. 5. The Exponential equation for the amount of Guaranteed Wheat Prices (Rial/kg), during 1990–2017.

9
N. Salehnia, et al. Ecological Indicators 111 (2020) 105991

Change 2014 (19), 211–227. https://doi.org/10.1007/s11027-012-9435-x.


Bennett, C., Grose, R., Corney, P., White, J., Holz, K., Katzfey, J., Post, A., Bindoff, L.,
2014. Performance of an empirical bias-correction of a high-resolution climate da-
taset. Int. J. Climatol. 34, 2189–2204. https://doi.org/10.1002/joc.3830.
Breusch, T.S., Pagan, A.R., 1979. A simple test for heteroskedasticity and random coef-
ficient variation. Econometrica 47 (5), 1287–1294 JSTOR 1911963. MR 0545960.
Cabas, J., Weersink, A., Olale, E., 2010. Crop yield response to economic, site and climatic
variables. Clim. Change 101, 599–616. https://doi.org/10.1007/s10584-009-9754-4.
Cerri, C., Eduardo, P., Sparovek, G., Bernoux, M., Willian, E., Easterling, J., Melillo, M.,
Cerri Carlos, C., 2007. Tropical agriculture and global warming: impacts and miti-
gation options. Sci. Agric. (Piracicaba, Braz) 64 (l), 83–99. https://doi.org/10.1590/
S0103-90162007000100013.
Chaves, J.P., 2001. An International Analysis of Agricultural Productivity. Agricultural
Investment and Productivity in Developing Countries. FAO Economic and
Development Paper No: 148.
Chen, C.C., Chang, C.C., 2005. The impact of weather on crop yield distribution in
Taiwan: some new evidence from panel data models and implications for crop in-
surance. Agric. Econ. 33 (supplement), 503–511. https://doi.org/10.1111/j.1574-
0864.2005.00097.x.
Chow, G.C., 1960. Tests of equality between sets of coefficients in two linear regressions.
Econometrica 28 (3), 591–605. https://www.jstor.org/stable/1910133.
Coelli, T.J., 1996. A guide to DEAP version 2.1: A Data Envelopment Analysis (Computer)
program. CEPA Working Papers No. 8/96. ISSN 1327-435X. ISBN 1 86389 4969.
Department of Econometrics, University of New England, Armidale, Australia.
Darmofal, D., 2015. Time-Series Cross-Sectional and Panel Data Models. University of
South Carolina, Cambridge University Press, pp. 141–157. https://doi.org/10.1017/
CBO9781139051293.009.
Das, B., Nair, B., Reddy, V.K., et al., 2018. Evaluation of multiple linear, neural network
and penalised regression models for prediction of rice yield based on weather para-
meters for west coast of India. Int. J. Biometeorol. 62, 1809. https://doi.org/10.
1007/s00484-018-1583-6.
Dong, Z., Pan, Z., He, Q., et al., 2018. Vulnerability assessment of spring wheat pro-
duction to climate change in the Inner Mongolia region of China. Ecol. Indic. 85,
67–78. https://doi.org/10.1016/j.ecolind.2017.10.008.
Durbin, J., Watson, G.S., 1951. Testing for serial correlation in least squares regression, II.
Biometrika 38 (1–2), 159–179. https://doi.org/10.1093/biomet/38.1-2.159. JSTOR
2332325.
Edwards, D.C., McKee, T.B., 1997. Characteristics of 20th century drought in the United
States at multiple time scales”. Climatology Rep. 97–2. Department of Atmospheric
Science, Colorado State University, Fort Collins, Colorado.
Eyshi Rezaie, E., Bannayan, M., 2012. Rainfed wheat yields under climate change in
northeastern Iran. Meteorol. Appl. 19, 346–354. https://doi.org/10.1002/met.268.
FAO, 2012. Downloads for AquaCrop (Version 4.0) Standard Window Program and Plug-
in Program. http://www.fao.org/nr/water/aquacrop.html.
FAOSTAT, 2010. Food and Agriculture Organization of the United Nations statistics.
http://faostat.fao.org/.
Gbetibouo, G.A., Hassan, R.M., 2004. Measuring the economic impact of climate change
on major South African field crops: a Ricardian approach. Global Planet. Change 47,
143–152. https://doi.org/10.1016/j.gloplacha.2004.10.009.
Ghassemi, H., Harrison, G., Mohammad, K., 2002. An accelerated nutrition transition in
Fig. 6. The predicted amount of wheat yield vs. baseline wheat yield (kg/ha)
Iran. Public Health Nutr. 5 (1A), 55–149. https://doi.org/10.1079/PHN2001287.
for three stations, over two decades, based on MIROC RCP45. Golkar Hamzee Yazd, H.R., Salehnia, N., Kolsoumi, S., Hoogenboom, G., 2019. Prediction
of climate variables by comparing the k-nearest neighbor method and MIROC5
outputs in an arid environment. Clim. Res. J. 77, 99–114. https://doi.org/10.3354/
Appendix A. Supplementary data cr01545.
Guttman, N., 1999. Accepting the standardized precipitation index: a calculation algo-
Supplementary data to this article can be found online at https:// rithm. J. Am. Water Resour. Assoc. 35 (2), 311–322. https://doi.org/10.1111/j.1752-
1688.1999.tb03592.x.
doi.org/10.1016/j.ecolind.2019.105991. Hernandez-Barrera, S., Rodriguez-Puebla, C., Challinor, A.J., 2017. Effects of diurnal
temperature range and drought on wheat yield in Spain. Theor. Appl. Climatol. 129
References (1–2), 503–519. https://doi.org/10.1007/s00704-016-1779-9.
Holder, C., Boyles, R., Robinson, P., Raman, S., Fishel, G., 2006. Calculating a normal
temperature range that reflects daily temperature variability. Bull. Am. Meteorol.
Abramowitz, M., Stegun, I.A., 1965. Handbook of Mathematical Functions, with Soc. 87, 769–774.
Formulas, Graphs, and Mathematical Tables. Dover Publications, pp. 1046. Hsiao, C., 2014. Analysis of Panel Data, third ed. Cambridge University Press, New York.
Agriculture Statistics, 2013. Vol. 2, Horticultural Crops, Iranian Ministry of Agriculture IPCC, 2007. Summary for Policy Makers. Climate Change 2007: The Physical Science
(in Persian). Basis. Contribution of Working Group I to the Fourth Assessment Report. Cambridge
AgriMetSoft, 2018. Drought Monitor and Prediction (Version 1.0) [Computer software]. University Press, Cambridge.
Available at: https://agrimetsoft.com/Drought Monitoring And Prediction.aspx. IPCC, 2014. Summary for Policy makers, in: Climate Change 2014: Impacts, Adaptation,
AgriMetSoft, 2018. Data Tool [Excel Add-ins]. Available at: https://agrimetsoft.com/ and Vulnerability. Contribution of Working Group II to the Fifth Assessment Report of
data-tool.aspx. the Intergovernmental Panel on Climate Change.
Allen, R.G., Pereira, L.S., Raes, D., Smith, M., 1998. Crop Evapotranspiration: Guidelines Jing-Song, S., Guang-Sheng, Z., Xing-Hua, S., 2012. Climatic suitability of the distribution
for Computing Crop Water Requirements. FAO Irrigation and Drainage Paper 56. of the winter wheat cultivation zone in China. Europ. J. Agronomy 43, 77–86.
FAO, Rome. https://doi.org/10.1016/j.eja.2012.05.009.
Angelidis, P., Maris, F., Kotsovinos, N., et al., 2012. Computation of drought index SPI Jones, J.W., Hoogenboom, G., Porter, C.H., Boote, K.J., Batchelor, W.D., Hunt, L.A.,
with alternative distribution functions. Water Resour. Manage. 26, 2453. https://doi. Wilkens, P.W., Singh, U., Gijsman, A.J., Ritchie, J.T., 2003. The DSSAT cropping
org/10.1007/s11269-012-0026-0. system model. Eur. J. Agron. 18 (3–4), 235. https://doi.org/10.1016/S1161-
Arshad, M., Amjath-Babu, T.S., Aravindakshan, S., Krupnik, T.J., Toussaint, V., Kächele, 0301(02)00107-7.
H., Müller, K., 2018. Climatic variability and thermal stress in Pakistan’s rice and Knox, J., Hess, T., Daccache, A., Wheeler, T., 2012. Climate change impacts on crop
wheat systems: a stochastic frontier and quantile regression analysis of economic productivity in Africa and South Asia. Environ. Res. Lett. 7, 034032. https://doi.org/
efficiency. Ecol. Indic. 89, 496–506. https://doi.org/10.1016/j.ecolind.2017.12.014. 10.1088/1748-9326/7/3/034032.
Asseng, S., Pannell, D.J., 2013. Adapting dryland agriculture to climate change: farming Koocheki, A., Nassiri, M., Jamali, J.B., Marashi, H., 2006. Evaluation of the effects of
implications and research and development needs in Western Australia. Clim. Change climate change on growth characteristics and yield of rainfed wheat in Iran. Agric.
118 (2), 167–181. https://doi.org/10.1007/s10584-012-0623-1. Sci. 20, 83–95 (Abstract in English).
Bannayan, M., Eyshi Rezaei, E., 2014. Future production of rainfed wheat in Iran Kristensen, K., Schelde, K., Olesen, J.E., 2011. Winter wheat yield response to climate
(Khorasan province): climate change scenario analysis. Mitig. Adapt. Strateg. Glob. variability in Denmark. J. Agric. Sci. 149, 33–47. https://doi.org/10.1017/

10
N. Salehnia, et al. Ecological Indicators 111 (2020) 105991

S0021859610000675. scenarios. Clim. Res. 7, 271–281. https://doi.org/10.3354/cr007271.


Lobell, D.B., Burke, M.B., 2010. On the use of statistical models to predict crop yield Sendhil, R., Jha, A., Kumar, A., Singh, S., 2018. Extent of vulnerability in wheat produ-
responses to climate change. Agric. Forest Meteorol. 150 (11), 1443–1452. https:// cing agro-ecologies of India: Tracking from indicators of cross-section and multi-di-
doi.org/10.1016/j.agrformet.2010.07.008. mension data. Ecol. Indic. 89, 771–780. https://doi.org/10.1016/j.ecolind.2018.02.
Lobell, D.B., Ortiz-Monasterio, J.I., 2007. Impacts of day versus night temperatures on 053.
spring wheat yields: a comparison of empirical and CERES model predictions in three Snyder, R.L., Organ, M., Bali, K., Eching, S., 2004. Basic irrigation scheduling BIS. http://
locations. Agron. J. 99 (2), 469–477. https://doi.org/10.2134/agronj2006.0209. www.waterplan.water.ca.gov/landwateruse/wateruse/Ag/CUP/California_Climate_
Lobell, D.B., Ortiz-Monasterio, J.I., Asner, P.G., Matson, P.A., Naylor, R.L., Falcon, W.P., Data_010804.xls.
2005. Analysis of wheat yield and climatic trends in Mexico. Field Crops Res. 94 Taylor, K.E., 2001. Summarizing multiple aspects of model performance in a single dia-
(2–3), 250–256. https://doi.org/10.1016/j.fcr.2005.01.007. gram. J. Geophys. Res. Atmos. 106 (D7), 7183–7192. https://doi.org/10.1029/2000J
MacLean, A., 2005. Statistical Evaluation of WATFLOOD (Ms). University of Waterloo, D9007 19. (ISSN 2156-2202).
Ontario, Canada. Themeßl, M.J., Gobiet, A., Leuprecht, A., 2011. Empirical-statistical downscaling and
Mann, M.L., Warner, J.M., 2017. Ethiopian wheat yield and yield gap estimation: A error correction of daily precipitation from regional climate models. Int. J. Climatol.
spatially explicit small area integrated data approach. Field Crops Res. 201, 60–74. 31 (10), 1530–1544. https://doi.org/10.1002/joc.2168/full.
https://doi.org/10.1016/j.fcr.2016.10.014. Trnka, M., Dubrovsky, M., Semeradova, D., Zalud, Z., 2004. Projections of uncertainties in
Maraun, D., Wetterhall, F., Ireson, A.M., Chandler, R.E., Kendon, E.J., Widmann, M., climate change scenarios into expected winter wheat yields. Theor. Appl. Climatol.
Brienen, S., et al., 2010. Precipitation downscaling under climate change: recent 77 (3–4), 229–249. https://doi.org/10.1007/s00704-004-0035-x.
developments to bridge the gap between dynamical models and the end user. Rev. Tubiello, F.N., Rosenzweig, C., Goldberg, R.A., Jagtap, S., Jones, J.W., 2002. Effects of
Geophys. 48 (RG3003), 380. https://doi.org/10.1029/2009RG000314. climate change on US crop production: simulation results using two different GCM
McKee, T.B., Doesken, N.J., Kliest, J., 1993. The relationship of drought frequency and scenarios. Part I: Wheat, potato, maize, and citrus. Clim. Res. 20, 259–270. https://
duration to time scales. In: Proc. of the 8th Conference on Applied Climatology. doi.org/10.3354/cr020259.
American Meteorological Society, Boston, pp. 179–184. Ullah, A., Salehnia, N., Kolsoumi, S., Ahmad, A., Khaliq, T., 2018. Prediction of effective
Ministry of Agriculture-Jahad, Islamic Republic of Iran, 2018. https://www.maj.ir/. climate change indicators using statistical downscaling approach and impact as-
Moss, R.H., et al., 2010. The next generation of scenarios for climate change research and sessment on pearl millet (Pennisetum glaucum L.) yield through Genetic Algorithm in
assessment. Nature 463, 747–756. https://doi.org/10.1038/nature08823. Punjab, Pakistan. Ecol. Indic. 90, 569–576. https://doi.org/10.1016/j.ecolind.2018.
Nassiri, M., Koocheki, A., Kamali, G.A., Shahandeh, H., 2006. Potential impact of climate 03.053.
change on rainfed wheat production in Iran. Arch. Agron. Soil Sci. 52 (1), 113–124. Vicente-Serrano, S.M., Begueria, S., Lopez-Moreno, J.I., 2010. A multi-scalar drought
https://doi.org/10.1080/03650340600560053. index sensitive to global warming: the standardized precipitation evapotranspiration
O’Gorman, T.W., Woolson, R.F., 1991. Variable selection to discriminant between two index. J. Clim. 23 (7), 1696–1718. https://doi.org/10.1175/2009JCLI2909.1.
groups: stepwise logistic regression or stepwise discriminant analysis? Am. Statist. White, J.W., Hoogenboom, G., Kimball, B.A., Wall, G.W., 2011. Methodologies for si-
Assoc. 45 (3), 187–193. https://www.jstor.org/stable/2684288. mulating impacts of climate change on crop production. Field Crops Res. 124 (3),
Ozdogan, M., 2011. Modeling the impacts of climate change on wheat yields in 357–368. https://doi.org/10.1016/j.fcr.2011.07.001.
Northwestern Turkey. Agric. Ecosyst. Environ. 141 (1–2), 1–12. https://doi.org/10. World Bank, 2005. Agricultural Growth for the Poor: An Agenda for Development. World
1016/j.agee.2011.02.001. Bank, Washington, DC.
Rodriguez-Puebla, C., Ayuso, S., Frias, M., Garcia Casado, L., 2007. Effects of climate Xu, X., Gao, P., Zhu, X., Guo, W., Ding, J., Li, C., Zhu, M., Wu, X., 2019. Design of an
variation on winter cereal production in Spain. Clim. Res. 34 (3), 223–232. https:// integrated climatic assessment indicator (ICAI) for wheat production: a case study in
doi.org/10.3354/cr00700. Jiangsu Province, China. Ecol. Indic. 101, 943–953. https://doi.org/10.1016/j.
Ruiz-Ramos, M., Sanchez, E., Gallardo, C., Minguez, M.I., 2011. Impacts of projected ecolind.2019.01.059.
maximum temperature extremes for C21 by an ensemble of regional climate models Xue, Q., Zhu, Z., Musick, J.T., Stewart, B.A., Dusek, D.A., 2006. Physiological mechanisms
on cereal cropping systems in the Iberian Peninsula. Nat. Hazards Earth Syst. Sci. 11 contributing to the increased water-use efficiency in winter wheat under deficit ir-
(12), 3275–3291. https://doi.org/10.5194/nhess-11-3275-2011. rigation. Plant Physiol. 163 (2), 154–164. https://doi.org/10.1016/j.jplph.2005.04.
Salehnia, N., Salehnia, N., Ansari, H., Kolsoumi, S., Bannayan, M., 2019. Climate data 026.
clustering effects on arid and semi-arid rainfed wheat yield: a comparison of artificial Zargar, A., Sadiq, R., Naser, B., Khan, F., 2011. A review of drought indices. Environ. Rev.
intelligence and K-Means approaches. Int. J. Biometeorol. 63 (7), 861–872. https:// 19, 333–349. https://doi.org/10.1139/a11-013.
doi.org/10.1007/s00484-019-01699-w. Zhou, M., 2010. Barley production and consumption. In: Zhang, G., Li, C. (Eds.), Genetics
Semenov, M.A., Wolf, J., Evans, L.G., Eckersten, H., Iglesias, A., 1996. Comparison of and Improvement of Barley Malt Quality. Springer-Verlag Berlin Heidelberg, pp.
wheat simulation models under climate change, II. Application of climate change 1–17 https://doi.org/10.1007/978-3-642-01279-2_1.

11

You might also like