You are on page 1of 26

Uncorrected Proof

© 2023 The Authors Journal of Water and Climate Change Vol 00 No 0, 1 doi: 10.2166/wcc.2023.361

Drought characterization and severity analysis using GRACE-TWS and MODIS datasets:
a case study from the Awash River Basin (ARB), Ethiopia

Yitea Seneshaw Getahun a, *, Ming-Hsu Lib, Yi-Ying Chenc and Tewoderos Addisu Yateb
a
Natural Resources Management, Debre Berhan University, National Central University, Taoyuan 320, Taiwan
b
Graduate Institute of Hydrological and Oceanic Sciences, National Central University, Taoyuan, Taiwan
c
Academia Sinica, Research Centre for Environmental Changes, Taipei 115, Taiwan
*Corresponding author. E-mail: yiseneshaw@gmail.com

YSG, 0000-0001-6479-1954

ABSTRACT

Drought is the utmost highly devastating phenomenon in Ethiopia because of societies that are reliant on rainfall-dependent agriculture; thus,
it is crucial to characterize drought at the basin scale using new developments in remote sensing products and recently proposed drought
indices. This study aimed to quantify drought in the Awash River Basin (ARB) using drought indices of the Standardized Precipitation Index
(SPI), Standardized Precipitation Evapotranspiration Index (SPEI), Evaporative Demand Drought Index (EDDI), Evaporative Stress Index (ESI),
and Water Storage Deficit Index (WSDI). The Mann–Kendall test analysis of annual and seasonal terrestrial water storage (TWS) showed a sig-
nificant increase from 2002 to 2017 that is beneficial for developmental activities in the ARB. GRACE showed a record high extreme drought
that persisted for 15 months was noticed from 2005/01 to 2006/03 with a total water storage deficit of 411.8 mm with a peak shortage of
46.24 mm in 2005/03, representing severe terrestrial water shortage in the ARB. This GRACE-TWS-based quantified extreme water shortage
in 2005/30 can be used as a threshold for adaptation. Overall, this study provides a reliable outcome that will be vital for the establishment of
climate change adaptation pathways in the future for viable water resources management to minimize the disastrous impacts of drought in
the AR.

Key words: drought characteristics, drought indices, drought severity, GRACE, MODIS, TWSA

HIGHLIGHTS

• TWS showed a significantly increasing trend in the basin with an annual rate of 3.6 mm/year.
• A significant TWS deficit was observed before 2009, and a significant increase afterward.
• The SPI, SPEI, ESI, and EDDI indices highly correlated at 3- and 6-month time scales.
• GRACE-based, i.e. a total water deficit of 411.8 mm is a useful threshold for climate change adaptation.

This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and
redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 2

GRAPHICAL ABSTRACT

1. INTRODUCTION
Drought is one of the most complex climate extreme events that affect the agricultural productivity and socio-economic devel-
opment of a region (Van Loon 2015; Mera 2018; Pendergrass et al. 2020). East Africa including Ethiopia has been
experiencing more frequent, long-lasting, and disastrous droughts that impact the lives of millions (Gebremeskel et al.
2019; Nguvava et al. 2019; Thomas et al. 2019). The chain of mountains and rugged topography in East Africa caused unusual
climate variability and considerably lower total annual rainfall than the rest of equatorial Africa, which plays a vital role in
recurring droughts (Liebmann et al. 2014). Additionally, the warmer climate in the future is capable of increasing the duration
and intensity of droughts in various places across the globe (Pendergrass et al. 2020). Most of the droughts in Ethiopia have
been linked with El Niño events (El Kenawy et al. 2016; Suryabhagavan 2017; MacDonald et al. 2019). Droughts in Ethiopia
have resulted in crop failure, death of livestock, human losses, conflict, diseases, epidemics outbreaks, and water shortage
(Gebremeskel et al. 2019; Nguvava et al. 2019). Hence, timely characterization of historical droughts in Ethiopia using
recent advancements in remote sensing is crucial. Drought in Ethiopia used to occur every 3–5 and 6–8 years in the northern

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 3

part as well as every 8–10 years for the entire portion of the country, but recently more frequent droughts have occurred even
in 3 or 2 consecutive years (Mera 2018). The recent historical droughts in 2002, 2003, 2004, 2005, 2006, 2007, 2008, 2009,
2010, and 2011 have affected a number of 8.74, 13.46, 2.5, 5, 8, 6, 14, 14, 10, 12, and 12 million people with a cost of huma-
nitarian aid of 92.33, 496.41, 58.92, 285.5, 393.12, 276.0, 1,077.8, 707.75, 488.8, 822.5, and 187.9 million USD, respectively
(Shitarek 2012; Mohammed et al. 2017; World Bank 2017; USAID 2018). In Borena district alone, 247,000 and 26,000 ani-
mals were lost due to the droughts in 2006 and 2008, respectively (Shitarek 2012; Mohammed et al. 2017; World Bank 2017).
Drought assessment and monitoring have been carried out using remote sensing datasets, especially where ground-based
observed datasets are not sufficient (AghaKouchak et al. 2015). Remote sensing approaches are cost-effective and efficient to
quantify droughts at different scales (Sinha et al. 2019). Recent advancements in satellite remote sensing help to generate new
datasets that are used to propose new drought indices. For example, the Gravity Recovery and Climate Experiment (GRACE)
satellite was released in 2002/03 that used to measure fluctuations in the terrestrial gravity field on a monthly time scale used
to generate recent drought indices (Tapley et al. 2004; Thomas et al. 2014). First, the long-period mean gravity field should be
deducted from the monthly values, and then the monthly variation of the gravity field can be expressed in terms of mass
change unit that is correspondent with water height (Ditmar 2018). In other words, the monthly change in the gravity
field is primarily due to the water mass variation over the continent. Thus, the variation in the terrestrial gravity field is equiv-
alent to the fluctuation of landmass water storage (TWS), which is a new reliable means of assessing TWS variation (Xie et al.
2018). Moreover, TWS is a measure of integrated water storage in a vertical, which comprises surface water, groundwater,
vegetation water, and soil moisture.
The importance of utilizing GRACE data for drought quantification and characterization have verified in various places
worldwide (Yirdaw et al. 2008; Chen et al. 2009; Frappart et al. 2012; Long et al. 2013; Sinha et al. 2019; Yu et al. 2019).
That is to say, conventional drought assessment that relies on the prolonged deficit of rainfall (meteorological drought),
depletion of soil moisture that limits the availability of water to crops (agricultural drought), and a shortage of streamflow
(hydrological drought) were not sufficient to quantify the integrated land water storage over a larger basin (Sinha et al.
2019; Sun et al. 2019). Therefore, drought assessment as well as characterization using recently proposed drought indices
and satellite information is crucial to provide important information for resource management along with supplementing
the previous studies.
Furthermore, the global warming effect has not been well represented in drought indices of the commonly used Standar-
dized Precipitation Index (SPI) and Palmer Drought Severity Index (PDSI) (Beguería et al. 2014; Vicente-Serrano et al. 2018).
Thus, to incorporate the current warming effect, an improved version of the SPI that includes temperature, which is the Stan-
dardized Precipitation Evapotranspiration Index (SPEI), was proposed by Vicente-Serrano et al. (2010). Plus, actual
evapotranspiration (AET) and potential evapotranspiration (PET)-based drought indices, such as the Evaporative Stress
Index (ESI) and the newly proposed Evaporative Demand Drought Index (EDDI) have been utilized to capture rapidly evol-
ving drought events (Anderson et al. 2007; Hobbins et al. 2016). Both the ESI and EDDI are indirectly sensitive to soil
moisture anomalies that are used to characterize rapidly evolving drought events though droughts are generally slowly devel-
oping and receding phenomena (Otkin et al. 2015; McEvoy et al. 2016; Wilhite & Pulwarty 2017). These indices are
appropriate for drought early warning and preparedness although the ESI has a limitation to being implemented in various
regions (McEvoy et al. 2016; Yao et al. 2018).
The most recent droughts have been frequently affecting the rift valley, eastern, and southeastern parts of Ethiopia, where
90% (104,737 km2) of the ARB is located with the reduction of 30% crop production in 2015/16 drought event alone (Shiferaw
et al. 2014; Mera 2018; Thomas et al. 2019; Berhane & Tesfay 2020). The main economic activity of Ethiopia, including the
ARB, is reliant on rain-fed agriculture. In addition to rain-fed agriculture, livestock keeping is a significant source of livelihood
in the ARB, particularly for pastoralist societies living in the downstream part, which is also highly susceptible to drought. The
ARB is the most populated, utilized as well as developed river basin in the country, which has been seriously affected by water
scarcity and recurring droughts (Edossa et al. 2010; Murendo et al. 2010; Hailu et al. 2017). The impact of drought on agricul-
ture and livestock production has also been disastrous in the ARB. Drought characterization including magnitude, duration,
severity, and intensity is vital for drought prediction (Wilhite & Pulwarty 2017). Therefore, it is essential to characterize drought
in the ARB based on its severity, frequency, duration, and magnitude which could help to minimize production losses.
Various indices of drought have been proposed for drought characterization and monitoring. However, some indices are
area-specific, and their performance should be verified prior to applying them to different areas (Bayissa 2018). Drought indi-
ces have their own merits and demerits so drought monitoring using multiple indices can better characterize drought events

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 4

than using a single index (Bayissa 2018; Yao et al. 2018; Davarpanah et al. 2021). Moreover, depending on data availability
multiple indices can be used to better characterize drought events than using an index as each drought type can be sensitive to
various parameters such as temperature, rainfall, evapotranspiration, and PET. (Yao et al. 2018; Davarpanah et al. 2021). In
this study, the most frequently used drought indices such as the SPEI, SPI, and ESI as well as the recently proposed ones like
the EDDI and the Water Storage Deficit Index (WSDI) were applied. The objectives of this study are (1) to evaluate the
spatiotemporal variability of TWSA and estimate the Water Storage Deficit (WSD) to use it as a means to characterize
droughts in the ARB; (2) to compare the WSDI with other drought indices, namely the SPI, SPEI, EDDI, and ESI; and
(3) to examine the applicability of recently proposed drought indices over the ARB by comparing them with the conventional
drought index of the SPI at different time scales.
Most previous studies in the ARB have focused on drought assessment including its causes (Edossa et al. 2010; Masih et al.
2014; Mera 2018; MacDonald et al. 2019). These studies have applied commonly used drought indices that are not sufficient to
assess the integrated land water storage and evaporative demand of the atmosphere. Characterization of drought using multiple
drought indices that incorporate the significance of temperature and new developments in satellite remote sensing is crucial for
ARBs that have inadequate observed datasets. Moreover, the various consequences of droughts within Ethiopia are specific to
different places so characterizing drought on a basin scale is important. Although it is difficult to explicitly quantify and fully
estimate the impacts of droughts, their characterization by including more information is necessary. Therefore, it is essential
to utilize the new developments in remote sensing data and proposed drought indices for the better characterization of droughts
in the ARB that could be helpful to enhance water resources management and minimize potential fatalities.

2. MATERIALS AND METHODS


2.1. Study area
The ARB is the study area that is located in the east-central part of Ethiopia. It is located between 7°520 –12°N and 37°570 –43°
250 E. The river has a length of 1,200 km, and a basin area of about 116,374 km2 with an irrigation potential of 205,400 ha
(Dost et al. 2013). The mean elevation of the ARB varies from 215 to 4,185 m and the river originates from the central high-
lands and ends in Lake Abbe near the border with Djibouti (Figure 1(b)). About 24 million population (18 million people and

Figure 1 | Map of the study area: (a) Africa, Ethiopia (ARB) and (b) meteorological stations, Lakes, DEM, major rivers in the ARB.

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 5

6 million livestock) are estimated to live in the ARB, making it the most highly exploited basin in the country due to the pres-
ence of main industries, bigger cities, and higher populations in the ARB than in other basins of the country (Adeba et al.
2015; Hailu et al. 2017).
The water shortage has been severe in recent years due to climate change, population growth, and urbanization
which has affected the development activities in the ARB (Dost 2013; Adeba et al. 2015; Hailu et al. 2017). The
three seasons in the ARB based on rainfall distribution are (i) the main rainy season (MRS), nationally called
‘Kiremt’ from June to September (JJAS), (ii) the slight rainy season (mRS), nationally known as ‘Belg’ from February
to May (FMAM), and (iii) the dry season, also called as ‘Bega’ from October to January (ONDJ) (Degefu et al.
2017). These seasons are mainly due to the rotation of the earth around the sun, and the monsoon system in the
area that direct the atmospheric circulation. The rugged topography also plays a significant role in the regional rainfall
distribution (Viste 2012).

2.2. Datasets
2.2.1. GRACE data
The monthly variations in the earth’s gravity field because of mass redistribution mainly associated with the moment of water
on and through landmass have been detected by GRACE since 2002 and expressed as terrestrial water storage anomaly
(TWSA). The most recent gridded monthly mean TWSA given by the Center for Space Research (CSR), the Jet Propulsion
Laboratory (JPL), and the GeoForschungsZentrum Potsdam (GFZ) were downloaded from http://grace.jpl.nasa.gov. The
downloaded .tif format data were processed using R and ArcGIS. The missed data in 2002/01, 2002/02, 2002/03, 2002/
07, 2003/06, 2011/01, 2011/06, 2012/10, 2013/03, 2013/08, 2013/09, 2014/02, 2014/07, 2014/12, 2015/10, 2015/11,
2016/04, 2016/09, 2016/10, and 2017/12 were filled using linear interpolation (Yang et al. 2017; Yu et al. 2019). The
missed data were about 20 months from 192, which is about 10% of all. The monthly GRACE TWSA dataset from
2002 to 2017 was used for this study (Cammalleri et al. 2019; Yu et al. 2019; Yimer et al. 2022). GRACE data have
a 1° spatial resolution with 25 cells covering the ARB and a temporal resolution of 1 month. The mean of TWSA pro-
vided by GFZ, CSR, and JPL processing centers was suggested as the most effective way to minimize the possible noises
in the gravity field though there were no significant changes in TWSA data retrieved from three centers (Humphrey
et al. 2016). Moreover, the GRACE datasets were preprocessed to remove the signals by filtering and truncation (Land-
erer & Swenson 2012; Sakumura et al. 2014; Ma et al. 2017). The TWSA of GRACE is the sum of soil moisture,
groundwater, surface water, and vegetation water for a semi-arid region as there is no snowpack (Sun et al. 2018)
and expressed as,

TWS ¼ SW þ SM þ GW þ VW (1)

where SW, SM, GW, and VW represent surface water, soil moisture, groundwater, and vegetation water, respectively.

2.2.2. MODIS data


The AET and PET Moderate Resolution Imaging Spectroradiometer (MODIS) Terra data were obtained from the National
Aeronautics and Space Administration (NASA) and the United States Geological Survey (USGS) of https://e4ftl01.cr.usgs.
gov/MOLT/MOD16A2.006/. The MODIS Terra dataset, i.e. MOD16A2, an 8-day temporal scale (8-day average) was used
for this study with a spatial resolution of 500 m (Running et al. 2017a, 2017b). The AET and PET data used for this study
were obtained from 2002 to 2017. The Penman–Monteith equation was used to generate the MOD16A2 product of the
AET, while the Priestley–Taylor equation is used to estimate the PET (Mu et al. 2007; Mu et al. 2011; Seleshi 2018). The
detailed MODIS flow chart of generating AET and PET products of MOD16A2 is presented in the user’s guide (Priestley
& Taylor 1972; Running et al. 2017a, 2017b).

2.2.3. Observed meteorological datasets


The observed rainfall and temperature datasets for 15 years (2002–2016) were acquired from 35 stations of the National
Meteorological Institute (NMI), Ethiopia (Figure 1). The data obtained from NMI were preprocessed and quality checked.

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 6

2.3. Drought indicators


2.3.1. GRACE-based WSDI
WSD is determined as the difference between monthly TWSA and the long-term mean of TWSA over the same month
(Thomas et al. 2014; Sun et al. 2018; Yu et al. 2019), is expressed as follows:

WSDk, p ¼ TWSAk, p  TWSAp (2)

where WSDk,p and TWSAk,p are the water storage deficit and GRACE-based TWSA time series in mm for the pth month in
the year k, respectively; TWSAp is the long-period average of TWSA for the same month (the pth month in a year). A negative
WSD indicates a terrestrial water storage deficit as compared to its monthly mean values, whereas a positive value shows a
water storage surplus. If a negative WSD lasts for three or more consecutive months, it is considered a drought event (Thomas
et al. 2014). Drought events based on WSD could be better characterized and be compared with other drought indices via the
WSDI. The WSDI is computed by normalizing WSD as follows:
WSDk,p  m
WSDIk,p ¼ (3)
s
where WSDIk,p represents the WSDI time series for the pth month in the year k; μ is the average of WSD time series, and σ is
the standard deviation of WSD. The WSD was standardized to provide the WSDI for better characterization and comparison
of GRACE-derived WSD with other standardized drought indices. Based on Thomas et al. (2014), the average WSD during
the drought months and number of consecutive drought months was used to assess the drought severity based on WSD,
which can be described as follows:
Sev(t) ¼ Mm(t)Dd(t) (4)

where Sev(t) and Mm(t) represents the severity, and the average WSD of drought events at duration t, respectively. Dd(t) rep-
resents the number of consecutive months between the starting and ending of the drought event t. This drought severity
measurement indicates the total terrestrial water shortage for the drought event defined by WSD, whereas WSDI indicates
the relative monthly water deficit for the drought event.

2.3.2. Standardized Precipitation Index


McKee et al. (1993) proposed the SPI to characterize the rainfall deficit in multiple time scales that are used to detect and monitor
drought for a particular area (Munagapati et al. 2018). Several studies in Ethiopia have used the SPI to evaluate meteorological
drought (Viste & Sorteberg 2013; Suryabhagavan 2017; Zeleke et al. 2017; Teweldebirhan et al. 2019). Since the SPI is normal-
ized to the required location as well as time, it is commonly accepted and applied worldwide for different purposes (Masih et al.
2014). The SPI is not only flexible and uses rainfall data only but is also suitable for complex topography and climatic conditions
(El Kenawy et al. 2016). The SPI normalizes the variation of rainfall at multiple accumulation periods (1–48 months) at a par-
ticular location using the long-term average. The shorter time scale of SPI from 1 to 6 months is used to monitor the
occurrence of meteorological droughts and the longer time scale of SPI from 12 to 24 months accumulation period are often
used as proxies for agricultural as well as hydrological droughts (Munagapati et al. 2018; Cammalleri et al. 2019). In addition
to drought severity, other characteristics of drought, namely, duration, intensity, and magnitude, were also computed using
SPI approaches. A drought event begins when SPI reaches 1.0 and stops when SPI turns positive. The positive sum of the
SPI values for all the months within a drought event can be identified as a drought magnitude (DM) (WMO 2012; Suryabhagavan
2017). In other words, the sum of the negative deviation during a drought is considered a DM. DM measures the cumulative water
deficit for a given drought event (Thompson 2017). It can be expressed as follows:

X
n
DM ¼ SPIij (5)
i¼1

To avoid the negative value of DM, Equation (5) can be rewritten as

!
X
n
DM ¼ SPIij (1) (6)
i¼1

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 7

where DM denotes the drought magnitude, and n represents the number of consecutive months per drought event at j time-
scale. The ratio of drought magnitude over drought duration can be defined as drought intensity (DL) and expressed as

DM
DL ¼ (7)
Dd(t)

where Dd(t) represents the duration of drought.

2.3.3. Standardized Precipitation Evapotranspiration Index


Vicente-Serrano et al. (2010) proposed the SPEI which is highly responsive to global warming and has been used to charac-
terize droughts. The computation of the SPEI is like the SPI, except that the SPEI includes temperature, which can be
computed as the difference between rainfall and PET (rainfall-PET) (Vicente-Serrano 2010; Homdee et al. 2016). The
SPEI is calculated by adjusting a data series of (rainfall-PET) to fit with the log-logistic distribution. The difference between
rainfall and PET is referred to as the climatic water balance with the atmospheric evaporative demand (Beguería et al. 2014).
The drought severity is associated with evaporative water demand because of evapotranspiration as the temperature rises
could better be identified using the SPEI (Homdee et al. 2016).

2.3.4. Evaporative Stress Index


A remote sensing-based drought index ESI is computed as a ratio of AET to PET (Anderson et al. 2011, 2016; Sur et al. 2015).
It is defined as

AET
ESI ¼ (8)
PET

ESI is sensitive to moisture stress that indicates terrestrial water availability (Zhang et al. 2019). The ESI value ranges from
zero to one that is associated with dryness or wetness conditions of the land surface. The higher (positive) ESI means the
atmospheric demand of evapotranspiration is met by available soil moisture and vegetation, whereas the lower ESI means
the land surface meets little or none of the atmospheric ET demand (Christian et al. 2019). However, after normalization
the ESI values range from  3.5 to þ3.5 and the positive ESI value indicates less or no drought occurrence, whereas the nega-
tive ESI value shows the occurrence of drought for a given location and period (McEvoy et al. 2016; Vicente-Serrano et al.
2018). Normalization by PET helps to minimize the AET variation due to the seasonal difference of available energy as well
as vegetation coverage, which further emphasize on the currently available moisture to vegetation regardless of previous
moisture conditions (Anderson et al. 2016; Farahmand 2016).

2.3.5. Evaporative Demand Drought Index


EDDI is a newly proposed drought index only relying on PET (Hobbins et al. 2016). It measures the drying potential of the
atmosphere that induces vegetation stress on the ground (Hobbins et al. 2016; McEvoy et al. 2019). In other words, it uses to
monitor the atmospheric evaporative demand that leads to the onset and development of drought when extreme atmospheric
anomalies like high temperatures or lack of rainfall persist for several weeks (Christian et al. 2019). Similar to the rainfall
(SPI), SPEI, and AET/PET(ESI), the monthly PET was adjusted to a normal distribution, and the cumulative probabilities
were normalized. Positive EDDI value denotes a drier condition than normal that might lead to drought, whereas the nega-
tive values indicate a wetter condition (Hobbins et al. 2016). Like ESI, SPEI and WSDI, EDDI was accumulated to 1-, 3-, 6-, 9-
, and 12-month time scale.

2.3.6. Drought severity classes


These drought severity classes are used to describe the magnitude of drought events over the ARB.

2.3.7. Mann–Kendall test


Mann–Kendall (MK) test was applied for trend analysis of annual and seasonal TWS using R studio software. The MK test has
been used in many studies across the world and it indicates the presence of increasing or decreasing trends and whether the
trend is statistically significant or not (Spearman 1904; Onyutha 2021). The MK test statistic ‘B’ is computed based on (Mann

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 8

1945; Kendall 1975; Getahun et al. 2021) using the equation:

X
n1 X
n
B¼ sgn(Dj  Di ) (9)
i¼1 j¼iþ1

The MK test is used to a time series Di that is ranked from i ¼ 1, 2, …, n  1 and Dj, which is ranked from j ¼ i þ 1, 2, …., n.
Each of the data point Di is considered as a reference point which is compared with the remaining of the data point’s Dj in
such a way that:

8
< þ1 if(Dj  Di ) . 0
Sgn(Dj  Di ) ¼ 0 if(Dj  Di ) ¼ 0 (10)
:
1 if(Dj  Di ) , 0

where Di and Dj are the yearly or seasonal values in years i and j ( j . i) correspondingly.MK has been recognized that when
the number of observations is more than 10 (n . 10), the statistic ‘B’ is nearly normally distributed with the average and var-
iance becomes 0 (Kendall 1975). In this case, the variance (Var (B) or (σ 2)) is considered as:

" #
1 Xm
Var(B) ¼ n(n  1)(2n þ 5)  ti (ti  1)(2ti þ 5) (11)
18 t¼1

where n is the number of observations, and m is the number of tied groups (a set of ti sample time series data owning the same
values). The standardized statistics test Z is given below:

8
> B1
>
> 1 if B . 0
>
>
< [Var(B)]2
Z¼ 0 if B ¼ 0 (12)
>
>
>
> Bþ1
>
: 1 if B , 0
[Var(B)]2

where Z indicates a normal distribution, a positive Z and a negative Z show increasing and decreasing trends for the series,
respectively. Accordingly, the null hypothesis (H0) indicates that the observed data (d1, …, dn) comprise a sample of n random
variables that are independent as well as identically distributed. For the alternative hypothesis (H1) that is a two-sided test the
distributions of di and dj are not identical for all i, j  n with i ≠ j. For H0 (no trend) is considered if Z , Zα/2, whereas
rejected (H1) if Z . Zα/2, where α is the significance level and Zα/2 the standard normal distribution for α/2. For this
study, the MK test with 95% confidence level (α ¼ 0.05) was considered with Zα/2 ¼ 1.645). In the MK test, the negative

Table 1 | Drought severity classes of drought indice values

Drought condition SPI/SPEI/ESI EDDI WSDI

Extreme drought  2 2.0  3


Severe drought 1.5 to 1.99 1.5–1.99  2 to 3
Moderate drought 1.0 to 1.49 1.0–1.49  1 to 2
Near normal Mild drought 0.99 to 0.5 0.5 to 0.99 0.99 to 0 Near normal
Normal 0.5  0  0.5 0.5  0  0.5
Mild wet 0.5–0.99 0.99 to 0.5 0.99–0
Moderate wet 1.0–1.49 1.0 to 1.49 1–2
Very wet 1.5–1.99 1.5 to 1.99 2–3
Extremely wet 2.0  2 3
Note: McKee et al. 1993; Li et al. 2015; Hobbins et al. 2016; Yu et al. 2019.

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 9

value of the slope indicates a decreasing trend, while the positive value of the slope illustrates an increasing trend. The mag-
nitude of the trend itself cannot be estimated by the MK test. Hence, Sen’s slope estimation test calculates both the slope and
intercept. The magnitude of the trend is estimated by Theil (1950); Sen (1968) slope estimator methods. The slope (βij) of all
data pairs is calculated as (Sen 1968). Overall, βij between any two values of a time series y will be projected from:

 
dj  di
bij ¼ Median (13)
ji

for k ¼ 1, 2, …, N, that dj and di are categorized as data values at time j and i ( j . i) harmoniously, and N is the number of all
pairs dj and di. The considered pairwise slopes are ordered from smallest to largest and relabeled as β(1), β(2)… β(n). The
median slope (Q) of these N all paired values aredenoted as Sen’s estimator of slope that is computed as Qmed? β([Nþ1]/2)
if N becomes odd, and it is taken as Qmed? (β([N/2]) þ β([Nþ2])/2)/2 if N becomes even. A positive value of β shows an increas-
ing trend and a negative value of β provides a decreasing trend.

2.3.8. Correlation coefficient


The correlation analysis among drought indices and ENSO (SSTA) was calculated by R studio software using Pearson cor-
relation coefficient (r) as defined by (Pearson 1920; Spiegel 1998; Getahun et al. 2021). A two-tailed significance level of 5%
was used for this study, assuming that both variables are normally distributed. Suppose that the two correlating variables are
drought indices and ENSO, Y (one drought index or ENSO) and W (another drought index), both owing n values Y1, Y2, Y3,
…. and W1, W2, W3, …, correspondingly, and then r is defined as follows:

P  
(yi  X)(W i  W)
r ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P (14)
2P  2
(yi  y) (Wi  W)

 and W
where the summation continues along all n probable values of Y and W in this sample. Let, Y  are the average of Y and
W in this sample and r varying between 1  r  1. When r close to 1 shows that as one drought index increases, there is an
exact predictable decrease in another drought index and on the other hand, when r close to 1 implies that as one drought
index increase, there is an exact predictable increase in another index. However, when r close to 0 indicates that as one
drought index increases or decreases, another cannot be predicted.

3. RESULTS AND DISCUSSION


3.1. Spatial-temporal variation, and trends of seasonal and annual TWSA
As indicated in Figure 2, the ARB’s MRS and annual water storage seem slowly increased over time. In 2004, both the yearly
and MRS terrestrial water storage severely reduced over the entire basin mostly with the rate of 3.5 to 5 cm/year. The MRS
showed a more considerable increase of TWS mainly in the western, northwestern, as well as southwestern parts of the basin,

Figure 2 | Annual and seasonal spatial-temporal variations of TWS in the ARB from 2002 to 2017.

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 10

corresponding to the distinct seasonality that reflects high rainfall in the MRS. On the other hand, a smaller area in the down-
stream and eastern portion of the basin exhibited a decrease in TWS with a rate of 0 to 2.5 cm/year in 2005, 2006, 2013, and
2017. A reduction in annual TWS was observed in the first 8 years (2002–2009) and 2012 with a lesser rate of decline as the
years passed. Using the MK test, a basin average of TWSA from 2002 to 2016 was analyzed and showed a significant increas-
ing trend of annual and MRS TWSA was noticed in the basin, as indicated in Figure 3 and Table 2. The increasing rate of
yearly and MRS TWSA was 3.6 and 3.1 mm/year, respectively. In parallel with TWSA, the MK test was performed on the
average basin rainfall, and an insignificant increase in annual and MRS rainfall was detected in the ARB.
The minor rainy and dry season spatial variations of TWSA over the ARB are illustrated in Figure 2. As displayed in
Figure 2, the mRS TWS notably reduced across the basin with a reduced rate of 0 to 13.5 cm/year, except for 2014 and
2016. The maximum reduction of mRS TWS was observed in the western, northwestern, and southwestern portions of the
basin as compared to the eastern and downstream of the ARB. The highest reduction rate mainly varying from 6.5 to
13.5 cm/year was noticed in the western part of the ARB, whereas the reduction rate in the downstream portion of the
basin varied from 0 to 5.5 cm/year. Besides the evident seasonal variation of TWS that it increases in the MRS and reduces
in the mRS and dry seasons corresponding to the rainfall pattern, the overexploitation of groundwater for irrigation and other
purposes was a reason for the depletion of TWS in the mRS.
As shown in Figure 2 above, the dry season reduction of TWS was less severe than the minor rainy season. Additionally, the
number of years that showed a decrease in TWS was less as compared to the mRS. Overall, a Belg season (FMAM) TWS
decrease before 2010 and increase after that can be linked with ENSO that more La Niña years before 2010 as La Niña
decreases the FMAM rainfall in the basin. After 2010 there were more El Niño years plus high temperatures in the Indian
ocean (Dipole Moment of Indian Ocean (DMI)) that increase the FMAM rainfall in the basin (Getahun et al. 2021). More-
over, studies revealed that water storage in the African landmass is getting higher after some time due to climate change/

Figure 3 | Annual and seasonal average changes in terrestrial water storage in the ARB from 2002 to 2016.

Table 2 | The MK test of basin mean seasonal and annual rainfall and TWSA from 2002 to 2016

Rainfall TWSA

Seasons P-value Sen’s slope Trend P-value Sen’s slope Trend

JJAS 0.84 þ0.048 No 0.003 þ3.1 þ


FMAM 0.92 0.029 No 0.00012 þ4.5 þ
ONDJ 0.78 þ0.655 No 0.022 þ3.1 þ
Annual 0.62 þ0.105 No 0.0006 þ3.6 þ

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 11

variability (Reager et al. 2016; Hasan et al. 2019; Hasan & Tarhule 2019). The dry season TWS reduction rate is less than
3.5 cm/year in most of the years, except for 2004. Generally, as time went by, there was a clear increase of TWS in the
ARB for both mRS and dry seasons. Similar to the MRS and annual series the MK test detected a significantly increasing
trend of TWSA for the minor rainy and dry seasons with an increasing rate of 4.5 and 3.1 mm/year, respectively as indicated
in Figure 3 and Table 2. Correspondingly, the dry season rainfall trend was detected to increase insignificantly, while the mRS
rainfall was identified to decrease insignificantly. Besides, the groundwater abstraction, an insignificant decreasing trend of
rainfall may contribute to the severe reduction of TWS in the minor rainy season (mRS).
In summary, the ARB experienced a significant increasing trend of annual and seasonal TWSA from 2002 to 2017, while
the annual and seasonal rainfall exhibited an insignificant trend. However, the mRS hasexperienced a severe reduction of
TWS as compared to other seasons.

3.2. GRACE-based WSD estimation


The GRACE-derived WSD was computed based on Equation (2) of the method section. As indicated in Figure 4, a significant
WSD was exhibited before 2009, and afterward, a predominantly water storage surplus was noticed in most of the months.
Despite the interruption of six intermittent months (Jun 2002, Jul 2002, Jan 2003, Apr 2004, Apr 2006), the most extensive
water storage deficits that occurred from 2002 to 2006 coincide with the rainfall decrease in the basin. Cumulative WSD
from 2002 to 2017 showed a continual deficit of water storage, implying that the ARB has been experiencing a long-lasting
severe terrestrial water deficit. In addition to the WSD pattern, Figure 4 indicates the comparison between cumulative WSD
and cumulative rainfall anomaly, which both matches reasonably well and follow the same way. Similar to the cumulative
WSD, the cumulative rainfall anomaly revealed a continuous deficit of rainfall from 2002 to 2016. The cumulative WSD
showed a slight fluctuation as compared to the cumulative rainfall anomaly. The correlation between cumulative rainfall
anomaly and cumulative WSD was 0.86, which indicated that rainfall is the central role player that influences terrestrial
water storage in the ARB.
The drought severity results presented in Table 3 were calculated using Equation (4) of the method section. As indicated in
Table 3, among the thirteen drought events detected in this study period, the maximum drought duration was 15 consecutive
months (from Jan-2005 to Mar-2006) with the highest drought severity that recorded a total water deficit of 411.8 mm. The
second long-lasting drought periods that lasted for 11 months were from 2003/03–2003/12 and 2009/04–2010/02 with a
severity of 300.1 and 162.0 mm, respectively. On the other hand, because of the short duration, the highest average
total water deficit of 48.14 mm, and the second-highest drought severity was observed from 2004/05–2004/12 which
lasted for 8 months with a total water deficit of 385.1 mm. The study period’s highest peak deficits were recorded in

Figure 4 | WSD, cumulative WSD, cumulative rainfall anomaly for the ARB from 2002 to 2016/2017, the accumulated curve is the
accumulation of normalized (z-scored) monthly average rainfall and WSD.

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 12

Table 3 | Grace-based drought events and water deficit from 2002 to 2017

Time period
Duration Average Total Peak Peak
Years Months (consecutive months) deficit (mm) deficit (mm) deficit (mm) deficit month

2002 Jan-May 5 17.16 85.8 32.19 Apr


Oct-Dec 3 37.05 111.2 71.8 Dec
2003 Feb-Dec 11 27.28 300.1 62.91 Feb
2004 Jan-Mar 3 40.30 120.9 56.44 Mar
May-Dec 8 48.14 385.1 66.05 Aug
2005–2006 Jan-2005 to Mar-2006 15 27.45  411.8 46.24 Mar-2005
2006 May-Oct 6 19.08 114.5 38.48 Jul
2007 Jan-May 5 16.89 84.5 47.16 May
2008 Jan-Jun 6 21.46 128.8 72.32 Jun
2009–2010 Apr-2009 to Feb-2010 11 14.73 162.0 35.94 Sep-2009
2012 Apr-Jul 4 21.49 86.0 38.32 Jun
2015 Oct-Dec 3 1.96 5.9 5.3 Dec
2016 Sep-Nov 3 5.33 16.0 12.14 Nov
Note: Bold value indicates highest water deficit.

June 2008 and December 2002 lasted 6, and 3 months with a magnitude of 72.3 and 71.8 mm, respectively. As shown in
Table 3, the drought event in the basin is so severe that either two-drought events per year or long-lasting drought events for
two consecutive years were detected during the study periods.
The TWS deficit during the annual (2002–2009), the dry season (2002–2005) and all seasons in 2004 may be strongly
related to climate change/variability (ENSO and DMI) the delay effect of the earliest major El Niño of 1997 go along
with by three La Niña events in 1998, 1999, and 2000, which El Niño reduces the MRS rainfall, and La Niña reduces the
mRS rainfall (Getahun et al. 2021). Besides Getahun et al. (2021) indicated that the warming in the Indian Ocean resulted
in a reduction of rainfall in the ARB. GRACE-TWS change is a result of seasonal water mass movement that can be affected
both by natural variability and human intervention are responsible for drought and identifying their individual influence of
climatic variability/change, land use land cover and regional hydrology along with how they affect each other is highly chal-
lenging (Vishwakarma 2020). The mRS rainfall contribution to the annual rainfall of the ARB was 42%; the downstream part
is highly influenced by the mRS rainfall. Moreover, the years 2002 and 2004 were strong El Niño years that could cause per-
sistent drought in the basin. The mRS TWS deficit in the ARB for the whole study period is due to the excessive water
abstraction for agricultural water demand and other uses (Adeba et al. 2015; Mersha et al. 2018). These authors also men-
tioned that the water withdrawal in the basin from December to April has been excessive and inappropriate, which was
not based on proper water resources management. Moreover, the rainfall distribution and high evapotranspiration in the
basin could cause the TWS deficit (Xu et al. 2019; Yu et al. 2019).
According to the MK trend test, a significant increasing trend of annual and seasonal TWS was detected in the ARB that in
line with a study carried out in the Sahel region (Okavango River basin, Lake Victoria, Lake Tana, the Blue Nile/Nile River
basin) (Ahmed et al. 2014; Hasan et al. 2019; Hasan & Tarhule 2019). On the contrary, an insignificant trend of rainfall, Lake
Tana level, and TWS were identified in the Upper Blue Nile (Seyoum 2018). Overall, the significantly increasing trend of TWS
in the ARB could be explained by global changes (i.e. climate change/variability, land cover change, and other human activi-
ties) rather than rainfall patterns alone. The construction of dams as well as reservoirs could increase the TWS in the ARB
(Berhe et al. 2013; Ahmed et al. 2014; Xu et al. 2019). The water flux from the neighboring basins and an insignificant
increase in rainfall in the MRS may also increase the TWS in the ARB. A study in the Sahel found that the land use/
cover change to agricultural land exhibited a rapid increase in land water storage (groundwater recharge) via diffuse recharge
(Ibrahim et al. 2014; Werth et al. 2017). They called it the ‘Sahelian paradox’ since it contradicts the fact that the agricultural
land expansion due to deforestation reduces the infiltration capacity and increases runoff that reduces recharge. However,
different water conservation structures and ponds could enhance groundwater recharge in agricultural lands. The expansion
of farmlands in the ARB could increase the TWS regardless of rainfall reduction (Getahun & Van Lanen 2015; Ayele et al.

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 13

2018). Reager et al. (2016) also found an increase in terrestrial water storage around the globe, whereas the sea level rise is
slowing down (decrease) which could be induced by climate change and variability. This global sea level decrease could also
be another reason why the TWS increased in the ARB.

3.3. WSDI comparison with ENSO and other drought indices


In this study, drought indices of WSDI, SPI, SPEI, ESI, and EDDI were compared to each other to characterize drought
events in the ARB. Additionally, the influence of ENSO on drought events was examined. The SPI, SPEI, ESI, and EDDI
values were computed for the 1, 3, 6, 9, and 12-month time scales, as indicated in Figures 5–10. The longer time scale
such as 6, 9, and 12-month drought indices characterize the hydrological droughts, important to monitor water resources
(Li et al. 2015). The rainy season in the ARB is from JJAS when the basin gets enough amount of rainfall. Thus, the rainfall
shortage in accumulative 3- or 4-month can cause water stress, and soil moisture deficit that leads to crop failure so that the
basin may be more sensitive to the time scale of 3- or 6-month drought indices.
Figures 5 and 6 show the time series of the EDDI, ESI, SPI, and SPEI at a time scale of 1 month as well as WSDI and
Nino3.4 SSTA in the ARB over the study period. There were more extreme and severe drought events under SPI than
other drought indices; this suggests that extreme droughts are mainly controlled by rainfall than temperature. The overall
number of drought events in a 1-month time scale under SPI, ESI, EDDI, and SPEI were 12, 10, 8, and 9, respectively.
The El Niño event in 1997 goes along with the three La Niña events in 1998, 1999, and 2000 that ensure 4 successive
years of rainfall decrease in the lower lowland parts of the ARB, where the mRS is the foremost contributor, besides the
major El Niño in 2002 that reduce rainfall in the MRS. These consecutive significant deficits of rainfall may cause severe
water depletion from 2002 to 2009 that correspond to severe drought events in 2002, 2003, and 2004. Furthermore, the
year 2004 was also a strong El Niño year that could reduce rainfall in the basin.
The years 2002, 2004, 2009, and 2015 were categorized as El Niño years (warming phase of ENSO), whereas 2007, 2008,
2010, 2011, and 2016 were classified as La Niña years (cooling phase of ENSO). In general, the warming phase of ENSO is
known to cause a reduction of the MRS rainfall and increase the mRS rainfall, whereas the opposite is true for the cooling
phase of ENSO. As can be seen, the El Niño years were identified as extreme or severe drought events by most of the drought
indices. As an illustration, most drought events in the basin are associated with El Niño episodes even if a few La Niña years
were also detected as drought years in the ARB.
The temporal variation of 3- and 6-month time scale SPI, ESI, SPEI, and EDDI drought indices were depicted in Figures 7
and 8. In essence, drought indices of the 6-month time scale showed fewer drought events with longer drought duration as
compared to the 3-month time scale. The 3-month ESI revealed that 2006, 2009, and 2012 were extreme or severe drought
years, and the remaining years showed moderate drought events, except for 2003, 2007, 2010, and 2016. Based on the
3-month SPEI, EDDI, and SPI the severe and extreme drought events occurred in 2002, 2008, and 2009. Additionally,
SPEI and EDDI discovered severe and extreme drought events in 2006, whereas the SPI detected extreme as well as

Figure 5 | Drought indices of the WSDI and SSTA Nino3.4 from 2002 to 2016 for the ARB.

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 14

Figure 6 | Drought indices of the EDDI, ESI, SPEI, and SPI at a time scale of 1 month from 2002 to 2016 for the ARB.

Figure 7 | Drought indices of the ESI, SPEI, EDDI, and SPI at a time scale of 3 months from 2002 to 2016 in the ARB.

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 15

Figure 8 | Drought indices of the ESI, SPEI, EDDI, and SPI at a time scale of 6 months from 2002 to 2016 in the ARB.

Figure 9 | Drought indices of the ESI, SPEI, EDDI, and SPI at a time scale of 9 months from 2002 to 2016 in the ARB.

severe drought events in 2013 (Figure 7). Like the ESI, the 3-months SPEI, EDDI, and SPI showed moderate drought in most
of the remaining years except for 2003, 2007, 2010, and 2016.
The 6-month analysis of SPI, SPEI, ESI, and EDDI showed that severe and extreme drought events mostly occurred in
2002, 2008, 2009, and 2012 (Figure 8). Moreover, two or three drought indices included 2003 and 2006 as extreme and
severe drought years, and most of the remaining years experienced moderate drought, except for 2005, 2007, 2010, and

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 16

Figure 10 | Drought indices of the ESI, SPEI, EDDI, and SPI at a time scale of 12 months from 2002 to 2016 in the ARB.

2014. The drought in 2006 was severe under ESI and SPEI, implying that the influence of temperature in this year was higher
than the rainfall. Overall, under the same time scale, the drought indices of ESI, SPI, SPEI, and EDDI showed a high degree
of similarity. The patterns of SPI and SPEI were much closer to each other than the remaining indices. In line with our
results, many other studies indicated that overall, the MRS rainfall in Ethiopia decreases during El Niño event, while La
Niña increases the MRS and the reverse occurs during the mRS (Korecha & Barnston 2007; Gleixner et al. 2016; Degefu
et al. 2017; Getahun et al. 2021).
Figures 9 and 10 show the drought indices of SPI, SPEI, EDDI, and ESI over 9 and 12-month time scales from 2002 to
2016. Based on SPI, SPEI, EDDI, and ESI drought indices of a 9-month time scale, severe and extreme drought events
were detected in 2002, 2003, 2008, and 2009, except that ESI showed moderate drought in 2002 as indicated in Figure 9.
Furthermore, SPI and SPEI drought indices of a 9-month time scale showed moderate drought in 2005, whereas indices
of SPEI and ESI indicated moderate drought in 2006. This indicates that the moderate drought in 2006 was highly influenced
by temperature rather than rainfall. Under a 9-month time scale of SPI, SPEI, EDDI, and ESI drought indices, moderate
droughts were also detected in 2011, 2012, 2013, 2015, and 2016.
As indicated in Figure 10, the 12-month drought indices of SPI, SPEI, and EDDI exhibited extreme and severe drought
events in 2003 and 2009. A time scale of 12-month ESI identified extreme and severe drought events in 2008, whereas
SPEI of the same time scale detected moderate drought in 2008. Using all drought indices of 12-month time scale moderate
drought events were detected in 2012 and 2015. As indicated in Figure 10, one or two-drought indices of 12-month time scale
detected moderate drought events in 2009, 2010, 2013, or 2016.

3.4. Correlation among the SPEI, SPI, EDDI, ESI, WSDI, and ENSO
As indicated in Figure 11, a correlation analysis between each drought indices and ENSO was carried out in the ARB. A high
correlation between ESI and SPEI drought indices that is greater than 0.9 was detected on 1-, 3-, and 6-month time scales
(Figure 11). Similarly, the correlation between drought indices (SPEI-SPI, EDDI-SPEI, and EDDI-ESI) was also high
(i.e. .+ 0.9) on 3- and 6-month time scales. Positive EDDI means that the atmosphere is very dry which leads to drought
and vegetation stress on the ground. The positive EDDI represents the drought itself and the likelihood of drought, whereas,
for other drought indices, the negative value represents the drought itself and the potential of drought.
Except for WSDI, all drought indices including ENSO showed a higher correlation on 3- and 6- month time scales than
other aggregation months. This indicates that the ARB drought indices at the time scale of 3- and 6-month can best explain

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 17

Figure 11 | Correlation of drought indices on 1-, 3-, 6-, 9-, and 12-month time scales at the ARB.

the agricultural and meteorological drought conditions in the basin and possibly the occurrence of agricultural drought. Most
importantly, since four of the drought indices such as SPI, SPEI, ESI, and EDDI have a high correlation to each other, it is
possible to use one of the indices in case of insufficient data availability.

3.5. Indices comparison based on severity and duration


The characteristics of drought namely, duration, magnitude, and intensity of ESI, EDDI, and SPEI drought indices were cal-
culated using the same approach as SPI, described in Equations (5) and (6. Regarding the WSDI drought index, drought
duration, severity, and other characteristics of drought were computed differently as described in Equations (2)–(4).
I. DM and severity using the WSDI
Considering WSDI, for a drought event to occur the negative values of WSDI should last for three or more successive
months, under those twelve drought events were detected in the ARB during 2002–2017 (Figure 12). Based on WSDI, the
maximum drought severity of 28.07 occurred from 2004/05 to 2006/03 for the duration of 23 months with a peak deficit
of 2.32 in 2004/08. This indicates an extreme shortage of total TWS in the ARB. The next most severe drought occurred
from 2003/02 to 2004/03 for the duration of 14 months, with a total severity of 14.83 and a peak deficit of 2.22 in
2003/02. As shown in Figure 12, the most persistent and severe drought events were indicated from 2002 to 2009, but after-
ward, the drought events were less in number and not severe.
II. DM and intensity using the SPI, ESI, EDDI, and SPEI
For a drought, an event to start using SPI, ESI, and SPEI drought indices, the SPI, ESI and SPEI values should reach 1.0
and ends when SPI, ESI, and SPEI values become positive. Whereas using EDDI for a drought event to start the EDDI value
should reach 1.0 and ends when the EDDI value becomes negative. About ten, ten, eight, and ten drought events were con-
firmed at a time scale of 3-month based on SPI, ESI, SPEI, and EDDI drought indices, respectively (Figures 13 and 14).
Considering a time scale of 6 months, each of the drought indices such as SPI, ESI, SPEI, and EDDI confirmed eight drought
events as shown in Figures 13 and 14. The drought indices of SPI, ESI, SPEI, and EDDI at a time scale of 3 months showed a
longer drought duration of 9, 10, 9, and 11 months either in 2008 or in 2009 with the drought intensity of 1.04, 1.05, 1.26,

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 18

Figure 12 | GRACE-based drought severity and water storage deficits in the ARB from 2002 to 2016. Note: reddish pink color indicates the
highest water deficit (411.8 mm) from 2005 to 2006. Please refer to the online version of this paper to see this figure in colour: http://dx.doi.
org/10.2166/wcc.2023.361.

Figure 13 | The EDDI-, ESI-, SPI-, and SPEI-based drought duration, magnitude and intensity at a 3-month time scale.

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 19

Figure 14 | The EDDI-, ESI-, SPI-, and SPEI-based drought duration, magnitude and intensity at a 6-month time scale.

and 1.05, respectively. On the other hand, under a 3-month time scale, the maximum drought intensity of 1.78 (SPI, a dur-
ation of 6 months), 1.38 (ESI, a duration of 7 months), 1.4 (SPEI, a duration of 7 months), and 1.32 (EDDI, a duration of 6
months) was occurred in 2002, except for ESI that occurred in 2009. The maximum DL under a 3-month time scale that
mostly occurred in 2002 was not necessarily correspondent with the longer drought duration shown in 2008/2009 indicating
that longer drought duration was less intense in 2008/2009. On a 6-month time scale, the long drought duration of 16 months
(SPI) with the intensity of 0.63, 14 months (ESI) with an intensity of 0.88, and 15 months (SPEI) with an intensity of 0.67
were indicated in 2012 or 2013. However, the EDDI’s longer drought duration of 9 months with a maximum intensity of 1.41
was identified in 2009, ultimately indicating an intense and longer drought duration (Figures 13 and 14). With the time
scale of 6 months, the maximum DL of 1.81 (SPI), 1.35 (ESI), and 1.28 (SPEI) was detected in 2002, whereas the
longer durations were revealed in 2012 or 2013. The drought indices of SPI, ESI, SPEI, and EDDI at a time scale of
3- and 6- months found that the most persistent droughts were observed in 2008–2009 and 2012–2013, while the maxi-
mum intensity was largely noticed in 2002 that could be associated with ENSO and global climate change (Davarpanah
et al. 2021; Getahun et al. 2021).
Figures 15 and 16 depicted the drought indices of SPI, ESI, SPEI, and EDDI on 9- and 12-month time scales. The number
of drought events using the drought indices of SPI, ESI, SPEI, and EDDI on a 9-month time scale was 6, 7, 7, and 5, respect-
ively. The longer drought duration of 27 months (SPI) with an intensity of 0.68, 16 months (ESI) with an intensity of 0.87,
16 months (SPEI) with an intensity of 0.67, and 11 months (EDDI) with an intensity of 0.72 were indicated under time scale
of 9-month from 2011 to 2013. Additionally, EDDI showed a longer drought duration of 11 months with an intensity of 1.16
in 2008 (Figures 15 and 16). Based on SPI, ESI, SPEI and EDDI, the maximum DL of 1.55 (duration of 10 months), 1.36
(duration of 10 months), 1.38 (duration of 9 months), and 1.41 (duration of 9 months) occurred in 2002, 2008, 2009 and 2009,
respectively. The number of drought years under a 12-month time scale was 5, 3, 5, and 4 using drought indices of SPI, ESI,

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 20

Figure 15 | The EDDI-, ESI-, SPI-, and SPEI-based drought duration, magnitude and intensity at a 9-month time scale.

SPEI, and EDDI, respectively. The longer drought duration was 26 months of SPI from 2011 to 2013, 24 months of ESI from
2008 to 2010, 13 months of SPEI from 2012 to 2013, and 26 months of EDDI from 2008 to 2010 (Figures 15 and 16).
Moreover, the drought onset (starting month) varies from one drought indices to another most with one or two months
back and forth. The drought onset also differs based on the monthly time scale change as indicated in Figures 12 and 13.
Overall, with the increase in the time scale of drought indices, the drought duration increased, while the number of drought
events considerably reduced. As indicated in Figures 12 and 13, in most of the cases, EDDI, SPI, and SPEI were capable to
detect a supposable drought onset month earlier, which is useful for drought estimation to be used for drought warning. Over-
all, it was possible to capture the most recent historical drought years in Ethiopia (2002, 2003, 2004, 2005, 2006, 2008, 2009,
2011, 2012, 2015, and 2016), particularly on 3- and 6-month time scales.
Several studies that have been investigating the cause of drought in Ethiopia and concluded that ENSO is the main cause of
most severe drought events in Ethiopia (Segele & Lamb 2005; Korecha & Barnston 2007; Fekadu 2015; Gleixner et al. 2016;
Degefu et al. 2017), which agrees with the findings in the ARB. Based on the suggestion that one index is not adequate for
accurately detecting and characterization of drought events (Zhong et al. 2020), different commonly used (SPI, SPEI, and
ESI) and newly proposed (WSDI, EDDI) drought indices were used for the better characterization of drought events in
the ARB. Since WSDI, EDDI, SPI, ESI, and SPEI are formulated differently using different variables and distribution func-
tions, some variation among drought indices is expected. However, the results indicated that drought indices exhibit a similar
overall pattern of detecting drought events, except for the variation in the onset and duration of droughts. A study in Upper
Blue Nile detected early onset of droughts using SPEI and SPI indices that correspond with our findings (Bayissa et al. 2018).
The most severe prolonged droughts of 2002–2003 were reported by several studies in Ethiopia including the ARB which
coincides with our results (Suryabhagavan 2017; Mohammed et al. 2017; Bayissa et al. 2018; Yadeta et al. 2020). In general,
drought indices of (SPI, ESI, SPEI, and EDDI) are able to detect most of the drought years such as 2002, 2003, 2004, 2005,

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 21

Figure 16 | The EDDI-, ESI-, SPI-, and SPEI-based drought duration, magnitude and intensity at a 12-month time scale.

2006, 2008, 2009, 2011, 2012, 2015, and 2016, especially on 3- and 6-month time scales (Viste & Sorteberg 2013; Masih et al.
2014; Mays 2014; Suryabhagavan 2017; USAID 2018; MacDonald et al. 2019). Studies across the globe also indicated that
the ongoing climate change induces drought is becoming intense, sever and frequent in recent decades (Shamshirband et al.
2020; Sahana et al. 2021; Band et al. 2022). Hence, regular assessment of drought using deferent drought indices is vital to
minimize the impact of drought to the most suspectable sector of agriculture. GRACE-based drought index is an excellent tool
to characterize hydrological or water resources related droughts: however, comparing/validating WSDI/WSD with other
conventional drought indices may not be a good approach as the hydroclimatic element used for each index is different
(Vishwakarma 2020). Vishwakarma (2020) stated that the severity of drought is not only dependent on climate anomalies
(rainfall deficit and high temperature) but is also on the landmass hydrological process and conditions. The drought indices
of SPI and SPEI well capture the drought events in Ethiopia and have been used in many parts of the country (Tefera et al.
2019). Overall, our results show that the severity and the length of the drought period coincides well with the major droughts
reported in Ethiopia (Mohammed et al. 2017; World Bank 2017; USAID 2018; MacDonald et al. 2019).

4. CONCLUSIONS
The spatial-temporal TWSA showed a clear seasonality that the MRS TWSA increased across the basin following rainfall pat-
terns, whereas the TWSA of the mRS and dry seasons decreased across the basin in most of the years. The excessive water
abstraction for irrigation and other uses was another reason for the depletion of TWS in the mRS and dry seasons besides the
seasonality. The terrestrial water storage in the ARB was significantly increased and the increasing rates of the annual and
MRS rainfalls were 3.6 and 3.1 mm/year, respectively. The SPI time scale of 9 months was better correlated with TWSA
in the ARB, indicating that SPI-9 can be used for assessing groundwater drought in the absence of groundwater measure-
ments. The WSD analysis from 2002 to 2017 showed a significant water shortage before 2009; afterward, the water

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 22

storage increased over time. The most severe water deficit occurred from 2002 to 2006 that coinciding with the rainfall short-
age in the basin. Of thirteen identified drought events using the WSD, the most severe drought was from 2005/01 to 2006/03
lasting for 15 months with a WSD of 411.8 mm and a peak shortage of 46.24 mm in 2005/03. The drought in 2004 was
also so severe that drought occurred twice per year from Jan-Mar and May-Dec with a total WSD of 120.9 and 385.1 mm,
respectively. Based on the results of the WSDI, severe and moderate drought events occurred from 2002 to 2009 and in 2012
within which 2002, 2003, 2004, and 2008 were relatively severe drought years. Based on the results of the WSDI, the most
severe drought of 28.07 shortage occurred from 2004/05 to 2006/03 for a duration of 23 consecutive months with a peak
shortage of 2.32 in 2004/08, representing an extreme terrestrial water shortage in the ARB. The second severe drought event
occurred from 2003/02 to 2004/03 for a duration of 14 consecutive months, with a total severity of 14.83. The severe deficits
from 2002 to 2006 might be explained by the lag effects of the earliest El Niño events in 1997 go along with three La Niña
events in 1998, 1999, and 2000 noting that the La Niña event will induce rainfall reductions in the mRS. Furthermore, the El
Niño events in 2002 and 2004 years induced reductions in the MRS rainfall that lead to persistent drought over the years.
Regardless of longer or shorter time scales, SPI, ESI, SPEI, and EDDI identified drought events in 2008–2009 and 2012–
2013, and the most severe drought in 2002. Drought indices of SPI, ESI, SPEI, and EDDI are capable to capture past drought
years in Ethiopia (i.e. 2002, 2003, 2004, 2005, 2006, 2008, 2009, 2011, 2012, 2015, and 2016), particularly on 3- and 6-month
time scales. The drought indices of SPI, SPEI, ESI, and EDDI with a time scale of 3 and or 6 months are the best indicators of
meteorological/agricultural droughts in the basin, whereas indices of SPI, and WSDI with a time scale of 9 months are best
indicators of groundwater drought. WSD severity is well quantified with drought indices of WSDI, while the magnitude of
water deficit is quantified with WSD. The correlation among most drought indices was higher on 3- and 6-month time
scales that a correlation of greater than 0.9 was observed among ESI vs. SPEI, SPEI vs. SPI, EDDI vs. SPEI, and EDDI
vs. ESI. Furthermore, the correlation between ENSO and drought indices was also relatively high at time scales of 3- and
6 months. Thus, drought indices with 3- and 6-month showed great potential to capture agricultural and meteorological
droughts in the ARB, and it is possible to use any one of these indices whenever insufficient data is available. Alternatively,
correlations between the WSDI and other drought indices were higher on 9- or 12- month time scales. The ARB is highly
sensitive to rainfall and a slight variation could cause a substantial influence on the socio-economic activities in the basin.
Although the significant increase of TWS over time is advantageous for irrigation and economic activities in the ARB, the
society relies on rain-fed agriculture and the drought is becoming more frequent and severe; so, it is vital to characterize/
quantify historical droughts to develop future adaptation pathways to minimize the possible impacts of drought in the basin.

ACKNOWLEDGEMENTS
We greatly recognize groups producing SST and GRACE datasets such as NOAA/NWS/CPC, NOAA/OAR/ESRL PSD, the
Center for Space Research (CSR), the Jet Propulsion Laboratory (JPL), and the GeoForschungsZentrum Potsdam (GFZ). We
acknowledge and appreciate the institutions specifically the National Aeronautics and Space Administration (NASA) and the
United States Geological Survey (USGS). We also acknowledge the Ethiopian National Meteorology Institute for giving us
the observed meteorological datasets.

AUTHORS’ CONTRIBUTIONS
Y.S.G. carried out the process and did data analysis, data interpretation, and drafted the first version of the manuscript. M.-
H.L. advised how to analyse the data, edited the paper, revised the paper, and shaped the manuscript. Y.-Y.C. revised, edited,
and amended the manuscript. T.A.Y. revised and amended the manuscript.

DATA AVAILABILITY STATEMENT


All relevant data are included in the paper or its Supplementary Information.

CONFLICT OF INTEREST
The authors declare there is no conflict.

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 23

REFERENCES

Adeba, D., Kansal, M. L. & Sen, S. 2015 Assessment of water scarcity and its impacts on sustainable development in Awash basin, Ethiopia.
Sustainable Water Resources Management 1 (1), 71–87. https://doi.org/10.1007/s40899-015-0006-7.
Ahmed, M., Sultan, M., Wahr, J. & Yan, E. 2014 The use of GRACE data to monitor natural and anthropogenic induced variations in water
availability across Africa. Earth-Science Reviews. Elsevier. https://doi.org/10.1016/j.earscirev.2014.05.009.
Anderson, M. C., Hain, C., Wardlow, B., Pimstein, A., Mecikalski, J. R. & Kustas, W. P. 2011 Evaluation of drought indices based on thermal
remote sensing of evapotranspiration over the continental United States. Journal of Climate 24 (8), 2025–2044. https://doi.org/10.1175/
2010JCLI3812.1.
Anderson, M. C., Zolin, C. A., Sentelhas, P. C., Hain, C. R., Semmens, K., Tugrul Yilmaz, M. & … Tetrault, R. 2016 The evaporative stress
index as an indicator of agricultural drought in Brazil: an assessment based on crop yield impacts. Remote Sensing of Environment 174,
82–99. https://doi.org/10.1016/j.rse.2015.11.034.
Ayele, G. T., Tebeje, A. K., Demissie, S. S., Belete, M. A., Jemberrie, M. A., Teshome, W. M. & … Teshale, E. Z. 2018 Time series land cover
mapping and change detection analysis using geographic information system and remote sensing, Northern Ethiopia. https://doi.org/10.
1177/1178622117751603.
Bachmair, S., Stahl, K., Collins, K., Hannaford, J., Acreman, M., Svoboda, M. & Overton, I. C. 2016 Drought indicators revisited: the need for
a wider consideration of environment and society. Wiley Interdisciplinary Reviews: Water 3 (4), 516–536. https://doi.org/10.1002/wat2.
1154.
Band, S. S., Karami, H., Jeong, Y. W., Moslemzadeh, M., Farzin, S., Chau, K. W. & … Mosavi, A. 2022 Evaluation of time series models in
simulating different monthly scales of drought index for improving their forecast accuracy. Frontiers in Earth Science 10, 839527.
Available from: https://www.frontiersin.org/articles/10.3389/feart.2022.839527/full
Bayissa, Y. A. 2018 Developing an Impact-Based Combined Drought Index for Monitoring Crop Yield Anomalies in the Upper Blue Nile
Basin, Ethiopia. Ph.D. Thesis, CRC Press, Delft University of Technology, The Netherlands. (accessed 4 July 2020).
Beguería, S., Vicente-Serrano, S. M., Reig, F. & Latorre, B. 2014 Standardized precipitation evapotranspiration index (SPEI) revisited:
parameter fitting, evapotranspiration models, tools, datasets and drought monitoring. International Journal of Climatology 34 (10),
3001–3023. https://doi.org/10.1002/joc.3887.
Berhane, A. & Tesfay, T. 2020 Impact of El Niño and La Niña on Agriculture in Ethiopia: Implications for El Niño and La Niña Adaptation
and Food Security in Ethiopia. Available from: https://www.preprints.org/manuscript/202009.0074/v1
Berhe, F. T., Melesse, A. M., Hailu, D. & Sileshi, Y. 2013 MODSIM-based water allocation modeling of Awash River Basin, Ethiopia. Catena
109, 118–128. https://doi.org/10.1016/j.catena.2013.04.007.
Cammalleri, C., Barbosa, P. & Vogt, J. V. 2019 Analyzing the relationship between multiple-timescale SPI and GRACE terrestrial water
storage in the framework of drought monitoring. Water 11 (8), 1672. https://doi.org/10.3390/w11081672.
Choi, M., Jacobs, J. M., Anderson, M. C. & Bosch, D. D. 2013 Evaluation of drought indices via remotely sensed data with hydrological
variables. Journal of Hydrology 476, 265–273. https://doi.org/10.1016/j.jhydrol.2012.10.042.
Christian, J. I., Basara, J. B., Otkin, J. A., Hunt, E. D., Wakefield, R. A., Flanagan, P. X. & Xiao, X. 2019 A methodology for flash drought
identification: application of flash drought frequency across the United States. Journal of Hydrometeorology 20 (5), 833–846. https://doi.
org/10.1175/JHM-D-18-0198.1.
Davarpanah, S., Erfanian, M. & Javan, K. 2021 Assessment of climate change impacts on drought and wet spells in Lake Urmia Basin. Pure
and Applied Geophysics 178 (2), 545–563. https://doi.org/10.1007/s00024-021-02656-8.
Degefu, M. A., Rowell, D. P. & Bewket, W. 2017 Teleconnections between Ethiopian rainfall variability and global SSTs: observations
and methods for model evaluation. Meteorology and Atmospheric Physics 129 (2), 173–186. https://doi.org/10.1007/s00703-016-
0466-9.
Dost, R., Obando, E. B. & Hoogeveen, W. 2013 Water Accounting Plus (WAþ) in the Awash River Basin Coping with Water Scarcity-
Developing National Water Audits Africa Client. FAO, Land and Water Division. Available from: http://www.wateraccounting.org/
files/projects/awash_basin.pdf. (accessed 15 July 2020).
Edossa, D. C., Babel, M. S. & DasGupta, A. 2010 Drought analysis in the Awash River Basin, Ethiopia. Water Resources Management 24 (7),
1441–1460. https://doi.org/10.1007/s11269-009-9508-0.
El Kenawy, A. M., McCabe, M. F., Vicente-Serrano, S. M., López-Moreno, J. I. & Robaa, S. M. 2016 Changes in the frequency and severity of
hydrological droughts over Ethiopia from 1960 to 2013. https://doi.org/10.18172/cig.2931.
Farahmand, A. 2016 Frameworks for Improving Multi-Index Drought Monitoring Using Remote Sensing Observations. Ph.D. Theses and
Dissertations, UC Irvine, USA. Available from: https://escholarship.org/uc/item/529g304 (accessed 15 May 2020).
Fekadu, K. 2015 Ethiopian seasonal rainfall variability and prediction using canonical correlation analysis (CCA). Earth Sciences 4 (3), 112.
https://doi.org/10.11648/j.earth.20150403.14.
Frappart, F., Papa, F., Santos Da Silva, J., Ramillien, G., Prigent, C., Seyler, F. & Calmant, S. 2012 Surface freshwater storage and dynamics in
the Amazon basin during the 2005 exceptional drought. Environmental Research Letters 7 (4). https://doi.org/10.1088/1748-9326/7/4/
044010.
Gebremeskel, G., Tang, Q., Sun, S., Huang, Z., Zhang, X. & Liu, X. 2019 Droughts in east Africa: causes, impacts and resilience. Earth-
Science Reviews. Elsevier B.V. https://doi.org/10.1016/j.earscirev.2019.04.015.

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 24

Getahun, Y. S. & Van Lanen, H. A. 2015 Assessing the impacts of land use-cover change on hydrology of Melka Kuntrie Subbasin in
Ethiopia, using a conceptual hydrological model. Hydrology:Current Research 6 (3), 210. https://doi.org/10.4172/2157-7587.1000210.
Getahun, Y. S., Li, M. H. & Pun, I. F. 2021 Trend and change-point detection analyses of rainfall and temperature over the Awash River basin
of Ethiopia. Heliyon 7 (9), e08024.
Gleixner, S., Keenlyside, N., Viste, E. & Korecha, D. 2016 The El Niño effect on Ethiopian summer rainfall. Climate Dynamics 49 (5–6),
1865–1883. https://doi.org/10.1007/s00382-016-3421-z.
Hailu, R., Tolossa, D. & Alemu, G. 2017 Water security: stakeholders’ arena in the Awash River Basin of Ethiopia. Sustainable Water
Resources Management 1, 19. https://doi.org/10.1007/s40899-017-0208-2.
Hasan, E. & Tarhule, A. 2019 Trend Dynamics of GRACE Terrestrial Water Storage in the Nile River Basin. Preprints, pp. 14–23. https://
doi.org/10.20944/preprints201909.0042.v1.
Hasan, E., Tarhule, A., Hong, Y. & Moore, B. 2019 Assessment of physical water scarcity in Africa using GRACE and TRMM satellite data.
Remote Sensing 11 (8), 904. https://doi.org/10.3390/rs11080904.
Hobbins, M. T., Wood, A., McEvoy, D. J., Huntington, J. L., Morton, C., Anderson, M. & Hain, C. 2016 The evaporative demand drought
index. Part I: linking drought evolution to variations in evaporative demand. Journal of Hydrometeorology 17 (6), 1745–1761. https://doi.
org/10.1175/JHM-D-15-0121.1.
Homdee, T., Pongput, K. & Kanae, S. 2016 A comparative performance analysis of three standardized climatic drought indices in the Chi
river basin, Thailand. Agriculture and Natural Resources 50 (3), 211–219. https://doi.org/10.1016/j.anres.2016.02.002.
Humphrey, V., Gudmundsson, L. & Seneviratne, S. I. 2016 Assessing global water storage variability from GRACE: trends, seasonal cycle,
subseasonal anomalies and extremes. Surveys in Geophysics. Springer Netherlands. https://doi.org/10.1007/s10712-016-9367-1.
Ibrahim, M., Favreau, G., Scanlon, B. R., Seidel, J. L., Le Coz, M., Demarty, J. & Cappelaere, B. 2014 Augmentation sur le long terme de la
recharge diffuse des aquifères suite à l’expension des cultures pluviales dans le Sahel, Afrique de l’Ouest. Hydrogeology Journal 22 (6),
1293–1305. https://doi.org/10.1007/s10040-014-1143-z.
Kendall, M. 1975 Rank Correlation Methods. Charles Griffin, London. Available from: http://www.worldcat.org/title/rank-
correlationmethods/oclc/489980698. (accessed 11 May 2019).
Korecha, D. & Barnston, A. G. 2007 Predictability of June–September rainfall in Ethiopia. Monthly Weather Review 135 (2), 628–650.
https://doi.org/10.1175/MWR3304.1.
Landerer, F. W. & Swenson, S. C. 2012 Accuracy of scaled GRACE terrestrial water storage estimates. Water Resources Research 48 (4).
https://doi.org/10.1029/2011WR011453.
Long, D., Scanlon, B. R., Longuevergne, L., Sun, A. Y., Fernando, D. N. & Save, H. 2013 GRACE satellite monitoring of large depletion in
water storage in response to the 2011 drought in Texas. Geophysical Research Letters 40 (13), 3395–3401. https://doi.org/10.1002/grl.
50655.
Ma, S., Wu, Q., Wang, J. & Zhang, S. 2017 Temporal evolution of regional drought detected from GRACE TWSA and CCI SM in Yunnan
Province, China. Remote Sensing 9 (11), 1124. https://doi.org/10.3390/rs9111124.
MacDonald, A. M., Bell, R. A., Kebede, S., Azagegn, T., Yehualaeshet, T., Pichon, F. & Calow, R. C. 2019 Groundwater and resilience to
drought in the Ethiopian highlands. Environmental Research Letters 14 (9), 095003. https://doi.org/10.1088/1748-9326/ab282f.
Mann, H. B. 1945 Nonparametric tests against trend. Econometrica 13, 245–259.
Masih, I., Maskey, S., Mussá, F. E. F. & Trambauer, P. 2014 A review of droughts on the African continent: a geospatial and long-term
perspective. Hydrology and Earth System Sciences 18 (9), 3635–3649. https://doi.org/10.5194/hess-18-3635-2014.
Mays, L. 2014 Integrated Urban Water Management: Arid and Semi-Arid Regions. CRC Press. https://doi.org/10.1201/9781482266207.
McEvoy, D. J., Huntington, J. L., Hobbins, M. T., Wood, A., Morton, C., Anderson, M. & Hain, C. 2016 The evaporative demand drought
index. Part II: cONUS-Wide assessment against common drought indicators. Journal of Hydrometeorology 17 (6), 1763–1779. https://
doi.org/10.1175/JHM-D-15-0122.1.
McEvoy, D., Hobbins, M., Brown, T., VanderMolen, K., Wall, T., Huntington, J. & Svoboda, M. 2019 Establishing relationships between
drought indices and wildfire danger outputs: a test case for the California-Nevada drought early warning system. Climate 7 (4), 52.
https://doi.org/10.3390/cli7040052.
McKee, T. B., Doesken, N. J. & Kleist, J. 1993 The Relationship of Drought Frequency and Duration to Time Scales. Available from: http://
citeseerx.ist.psu.edu/viewdoc/summary?doi ¼ 10.1.1.462.4342
Mera, G. A. 2018 Drought and its impacts in Ethiopia. Weather and Climate Extremes. Elsevier B.V. https://doi.org/10.1016/j.wace.2018.10.
002.
Mersha, A., Masih, I., de Fraiture, C., Wenninger, J. & Alamirew, T. 2018 Evaluating the impacts of IWRM policy actions on demand
satisfaction and downstream water availability in the upper awash basin, Ethiopia. Water 10 (7), 892. https://doi.org/10.3390/
w10070892.
Mohammed, Y., Yimer, F., Tadesse, M. & Tesfaye, K. 2017 Meteorological drought assessment in northeast highlands of Ethiopia. https://doi.
org/10.1108/IJCCSM-12-2016-0179.
Mu, Q., Heinsch, F. A., Zhao, M. & Running, S. W. 2007 Development of a global evapotranspiration algorithm based on MODIS and global
meteorology data. Remote Sensing of Environment 111 (4), 519–536. https://doi.org/10.1016/j.rse.2007.04.015.
Mu, Q., Zhao, M. & Running, S. W. 2011 Improvements to a MODIS global terrestrial evapotranspiration algorithm. Remote Sensing of
Environment 115 (8), 1781–1800. https://doi.org/10.1016/j.rse.2011.02.019.

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 25

Munagapati, H., Yadav, R. & Tiwari, V. M. 2018 Identifying water storage variation in Krishna Basin, India from in situ and satellite based
hydrological data. Journal of the Geological Society of India 92 (5), 607–615. https://doi.org/10.1007/s12594-018-1074-8.
Murendo, C., Keil, A. & Zeller, M. 2010 Drought Impacts and Related Risk Management by Smallholder Farmers in Developing Countries:
Evidence From Awash River Basin, Ethiopia. Research in Development Economics and Policy. Available from: https://ideas.repec.org/
p/ags/uhohdp/114750.html
Nguvava, M., Abiodun, B. J. & Otieno, F. 2019 Projecting drought characteristics over East African basins at specific global warming levels.
Atmospheric Research 228, 41–54. https://doi.org/10.1016/j.atmosres.2019.05.008.
Onyutha, C. 2021 Graphical-statistical method to explore variability of hydrological time series. Hydrology Research 52 (1), 266–283.
Available from: https://iwaponline.com/hr/article/52/1/266/78544/Graphical-statistical-method-to-explore
Pearson, K. 1920 Notes on the history of correlation. Biometrika 13 (1), 25–45. Available from: https://link.springer.com/article/10.1057/jt.
2009.5.
Pendergrass, A. G., Meehl, G. A., Pulwarty, R., Hobbins, M., Hoell, A., AghaKouchak, A., Bonfils, C. J. W., Gallant, A. J. E., Hoerling, M. &
Hoffmann, D. et al. 2020 Flash droughts present a new challenge for subseasonal-to-seasonal prediction. Nature Climate Change 10,
191–199. https://doi.org/10.1038/s41558-020-0709-0.
Priestley, C. H. B. & Taylor, R. J. 1972 On the assessment of surface heat flux and evaporation using large-scale parameters. https://doi.org/
10.1175/1520-0493(1972)100&lt;0081:OTAOSH . 2.3.CO;2.
Running, S., Mu, Q. & Zhao, M. 2017a MOD16A2: MODIS/Terra Net Evapotranspiration 8-Day L4 Global 500 m SIN Grid V006.
Distributed by NASA EOSDIS Land Processes DAAC. Available from: https://lpdaac.usgs.gov/products/mod16a2v006/ (accessed 12
February 2020).
Running, S. W., Mu, Q., Zhao, M. & Moreno, A. 2017b User’s guide MODIS global terrestrial evapotranspiration (ET) product (NASA
MOD16A2/A3) NASA earth observing system MODIS land algorithm. Available from: https://ladsweb.modaps.eosdis.nasa.gov/
missions-and-measurements/modis/MOD16_ET_User-Guide_2017.pdf (accessed 13 February 2020).
Sahana, V., Mondal, A. & Sreekumar, P. 2021 Drought vulnerability and risk assessment in India: sensitivity analysis and comparison of
aggregation techniques. Journal of Environmental Management 299, 113689. https://doi.org/10.1016/j.jenvman.2021.113689.
Sakumura, C., Bettadpur, S. & Bruinsma, S. 2014 Ensemble prediction and intercomparison analysis of GRACE time-variable gravity field
models. Geophysical Research Letters 41 (5), 1389–1397. https://doi.org/10.1002/2013GL058632.
Segele, Z. T. & Lamb, P. J. 2005 Characterization and variability of Kiremt rainy season over Ethiopia. Meteorology and Atmospheric Physics
89 (1–4), 153–180. https://doi.org/10.1007/s00703-005-0127-x.
Seleshi, Y. 2018 Calibration of the Priestley-Taylor evaporation model for Ethiopia. Zede Journal 36, 28–40. Available from: https://www.ajol.
info/index.php/zj/article/view/186799
Sen, P. K. 1968 Estimates of the regression coefficient based on Kendall’s tau. Journal of the American Statistical Association 63, 1379–
1389.
Seyoum, W. M. 2018 Characterizing water storage trends and regional climate influence using GRACE observation and satellite altimetry
data in the Upper Blue Nile River Basin. Journal of Hydrology 566, 274–284. https://doi.org/10.1016/j.jhydrol.2018.09.025.
Shamshirband, S., Hashemi, S., Salimi, H., Samadianfard, S., Asadi, E., Shadkani, S. & … Chau, K. W. 2020 Predicting standardized
streamflow index for hydrological drought using machine learning models. Engineering Applications of Computational Fluid Mechanics
14 (1), 339–350. Available from: https://www.tandfonline.com/doi/pdf/10.1080/19942060.2020.1715844.
Shiferaw, B., Tesfaye, K., Kassie, M., Abate, T., Prasanna, B. M. & Menkir, A. 2014 Managing vulnerability to drought and enhancing
livelihood resilience in sub-Saharan Africa: technological, institutional and policy options. Weather and Climate Extremes 3, 67–79.
https://doi.org/10.1016/j.wace.2014.04.004.
Sinha, D., Syed, T. H., Famiglietti, J. S., Reager, J. T. & Thomas, R. C. 2019 Characterizing drought in India using GRACE observations of
terrestrial water storage deficit. Journal of Hydrometeorology 18 (2), 381–396. https://doi.org/10.1175/JHM-D-16-0047.1.
Spearman, C. 1904 The proof and measurement of association between two things. The American Journal of Psychology 15 (1), 72–101.
https://doi.org/10.2307/1412159.
Spiegel, M. R. 1998 Theory and Problems of Statistics, Schaum’s Outline Series, 5th edn. McGraw-Hill, New York, NY, USA. Available from:
http://www.fulviofrisone.com/attachments/article/447/Schaum's%20Outline%20of%20Statistic.pdf
Sun, Z., Zhu, X., Pan, Y., Zhang, J. & Liu, X. 2018 Drought evaluation using the GRACE terrestrial water storage deficit over the Yangtze
River Basin, China. Science of the Total Environment 634, 727–738. https://doi.org/10.1016/j.scitotenv.2018.03.292.
Sur, C., Hur, J., Kim, K., Choi, W. & Choi, M. 2015 An evaluation of satellite-based drought indices on a regional scale. International Journal
of Remote Sensing 36 (22), 5593–5612. https://doi.org/10.1080/01431161.2015.1101653.
Suryabhagavan, K. V. 2017 GIS-based climate variability and drought characterization in Ethiopia over three decades. Weather and Climate
Extremes 15, 11–23. https://doi.org/10.1016/j.wace.2016.11.005.
Tapley, B. D., Bettadpur, S., Ries, J. C., Thompson, P. F. & Watkins, M. M. 2004 GRACE measurements of mass variability in the earth system.
Science 305 (5683), 503–505. https://doi.org/10.1126/science.1099192.
Tefera, A. S., Ayoade, J. O. & Bello, N. J. 2019 Comparative analyses of SPI and SPEI as drought assessment tools in Tigray Region, Northern
Ethiopia. SN Applied Sciences 1, 1265. https://doi.org/10.1007/s42452-019-1326-2.
Teweldebirhan, T. D., Uddameri, V., Forghanparast, F., Hernandez, E. A. & Ekwaro-Osire, S. 2019 Comparison of meteorological and
agriculture-related drought indicators across Ethiopia. Water 11 (11), 2218. https://doi.org/10.3390/w11112218.

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest
Uncorrected Proof

Journal of Water and Climate Change Vol 00 No 0, 26

Theil, H. 1950 A rank-invariant method of linear and polynomial analysis, part 3. In: Nederlandse Akademie van Wettenschappen.
Proceedings. Vol. 53, pp. 1397–1412.
Thomas, T., Nayak, P. C. & Ghosh, N. C. 2014 Spatiotemporal analysis of drought characteristics in the bundelkhand region of central India
using the standardized precipitation index. Journal of Hydrologic Engineering 20 (11). https://doi.org/10.1061/(ASCE)HE.1943-5584.
0001189
Thomas, E. A., Needoba, J., Kaberia, D., Butterworth, J., Adams, E. C., Oduor, P. & Nagel, C. 2019 Quantifying increased groundwater
demand from prolonged drought in the East African Rift Valley. Science of the Total Environment 666, 1265–1272. https://doi.org/10.
1016/j.scitotenv.2019.02.206.
USAID 2018 The United States Agency for International Development. Economics of Resilience to Drought Ethiopia Analysis. Available
from: https://www.agrilinks.org/sites/default/files/ethiopia_economics_of_resilience_final_jan_4_2018_-_branded.pdf (accessed 11 July
2020).
Van Loon, A. F. 2015 Hydrological drought explained. Wiley Interdisciplinary Reviews: Water 2 (4), 359–392. https://doi.org/10.1002/wat2.
1085.
Vicente-Serrano, S. M., Beguería, S. & López-Moreno, J. I. 2010 A multiscalar drought index sensitive to global warming: the standardized
precipitation evapotranspiration index. Journal of Climate 23 (7), 1696–1718. https://doi.org/10.1175/2009JCLI2909.1.
Vicente-Serrano, S. M., Miralles, D. G., Domínguez-Castro, F., Azorin-Molina, C., El Kenawy, A., Mcvicar, T. R. & Peña-Gallardo, M. 2018
Global assessment of the standardized evapotranspiration deficit index (SEDI) for drought analysis and monitoring. Journal of Climate
31 (14), 5371–5393. https://doi.org/10.1175/JCLI-D-17-0775.1.
Vishwakarma, B. D. 2020 Monitoring droughts from GRACE. Frontiers in Environmental Science 8, 584690. Available from: https://www.
frontiersin.org/articles/10.3389/fenvs.2020.584690/full
Viste, E. 2012 Moisture Transport and Precipitation in Ethiopia. Available from: https://folk.uib.no/evi003/Publications/Viste_
PhDthesis2012.pdf
Viste, E. & Sorteberg, A. 2013 Moisture transport into the Ethiopian highlands. International Journal of Climatology 33 (1), 249–263. https://
doi.org/10.1002/joc.3409.
Werth, S., White, D. & Bliss, D. W. 2017 GRACE detected rise of groundwater in the Sahelian Niger River Basin. Journal of Geophysical
Research: Solid Earth 122 (12), 10,459–10,477. https://doi.org/10.1002/2017JB014845.
Wilhite, D. & Pulwarty, R. S. 2017 Drought and Water Crises: Integrating Science, Management, and Policy. CRC Press.
WMO 2012 World Meteorological Organization. In: Standardized Precipitation Index User Guide (Svoboda, M., Hayes, M. & Wood, D., eds).
WMO-No. 1090; WMO, Geneva, Switzerland, p. 16. Available from: https://library.wmo.int/doc_num.php?explnum_id=7768
World Bank 2017 Federal Democratic Republic of Ethiopia, Rural Productive Safety Net Project (P163438). Available from: http://
documents1.worldbank.org/curated/en/830381505613638420/pdf/project-appraisal-document-pad-P163438-EU-edits-for-Board-
version-08252017.pdf (accessed 11 July 2020).
Xie, Y., Huang, S., Liu, S., Leng, G., Peng, J., Huang, Q. & Li, P. 2018 GRACE-based terrestrial water storage in northwest China: changes and
causes. Remote Sensing 10 (7), 1163. https://doi.org/10.3390/rs10071163.
Xu, L., Chen, N., Zhang, X. & Chen, Z. 2019 Spatiotemporal changes in China’s terrestrial water storage from GRACE satellites and its
possible drivers. Journal of Geophysical Research: Atmospheres 124 (22), 11976–11993.
Yadeta, D., Kebede, A. & Tessema, N. 2020 Climate change posed agricultural drought and potential of rainy season for effective agricultural
water management, Kesem sub-basin, Awash Basin, Ethiopia. Theoretical and Applied Climatology 140 (12), 653–666. https://doi.org/
10.1007/s00704-020-03113-7.
Yang, P., Xia, J., Zhan, C., Qiao, Y. & Wang, Y. 2017 Monitoring the spatio-temporal changes of terrestrial water storage using GRACE data in
the Tarim River basin between 2002 and 2015. Science of the Total Environment 595, 218–228. https://doi.org/10.1016/j.scitotenv.2017.
03.268.
Yao, N., Li, Y., Lei, T. & Peng, L. 2018 Drought evolution, severity and trends in mainland China over 1961–2013. Science of the Total
Environment 616–617, 73–89. https://doi.org/10.1016/j.scitotenv.2017.10.327.
Yirdaw, S. Z., Snelgrove, K. R. & Agboma, C. O. 2008 GRACE satellite observations of terrestrial moisture changes for drought
characterization in the Canadian Prairie. Journal of Hydrology 356 (1–2), 84–92. https://doi.org/10.1016/j.jhydrol.2008.04.004.
Yu, W., Li, Y., Cao, Y. & Schillerberg, T. 2019 Drought assessment using GRACE terrestrial water storage deficit in Mongolia from 2002 to
2017. Water 11 (6), 1301. https://doi.org/10.3390/w11061301.
Zeleke, T. T., Giorgi, F., Diro, G. T. & Zaitchik, B. F. 2017 Trend and periodicity of drought over Ethiopia. International Journal of
Climatology 37 (13), 4733–4748. https://doi.org/10.1002/joc.5122.
Zhang, X., Li, M., Ma, Z., Yang, Q., Lv, M. & Clark, R. 2019 Assessment of an evapotranspiration deficit drought index in relation to impacts
on ecosystems. Advances in Atmospheric Sciences 36 (11), 1273–1287. https://doi.org/10.1007/s00376-019-9061-6.
Zhong, L., Hua, L. & Yan, Z. 2020 Datasets of meteorological drought events and risks for the developing countries in Eurasia. Big Earth
Data 4 (2), 191–223. https://doi.org/10.1080/20964471.2019.1710383.

First received 25 September 2022; accepted in revised form 24 January 2023. Available online 2 February 2023

Downloaded from http://iwaponline.com/jwcc/article-pdf/doi/10.2166/wcc.2023.361/1170238/jwc2023361.pdf


by guest

You might also like