Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Natural Hazards (2020) 103:2989–3010

https://doi.org/10.1007/s11069-020-04116-3

ORIGINAL PAPER

Development of drought hazard index for vulnerability


assessment in Pakistan

Shahzada Adnan1,2 · Kalim Ullah1

Received: 14 March 2019 / Accepted: 11 June 2020 / Published online: 21 June 2020
© Springer Nature B.V. 2020

Abstract
Drought is a silent meteorological disaster that spreads over time, affecting water availabil-
ity for agriculture and livelihood in any region. The prediction of drought is a complex phe-
nomenon; however, the negative impacts of drought are mitigated by monitoring drought
events over a region. The present study provides spatial and temporal drought climatology
over Pakistan, using 60-years (1951–2010) observational gridded data (0.5° × 0.5°) of pre-
cipitation from Global Precipitation Climatological Center and soil moisture from Climate
Prediction Center. Using precipitation and soil moisture datasets, a novel drought hazard
index is developed to determine drought vulnerability across different districts of Pakistan.
Our findings identified 19 districts that are extremely vulnerable to drought, with northern
regions being vulnerable to mild drought, whereas central and southern districts are vul-
nerable to high drought events. By using standardized precipitation index and soil mois-
ture anomaly, six severe drought years were identified as 1952, 1969, 1971, 2000, 2001,
and 2002 in different parts of the country. Deficiency of monsoon rainfall is a major cause
of droughts in southern and rain-fed regions. This study is helpful for drought managers,
hydrologists, and contingency planners to prepare mitigation and adaptation plans toward
sustainable development in Pakistan.

Keywords Drought hazard index · Vulnerability assessment · Climatology of Pakistan

1 Introduction

Droughts are natural disasters affecting large areas and populations due to their unpre-
dictable nature. About 50% of the earth’s land is vulnerable to droughts (Kogan 1997).
According to the Emergency Events Database (EM-DAT), more than 200 million peo-
ple were affected and 11 million killed by droughts between 1900 and 2016 in the world
(Guha-Sapir 2017). The effects of droughts persist long after their termination, unlike
other hydrometeorological and geological disasters such as floods, cyclones, tornados, vol-
canic outbursts, and earthquakes (Vogt et al. 2011). Additionally, frequent droughts lead to

* Kalim Ullah
kalim_ullah@comsats.edu.pk
1
Department of Meteorology, COMSATS University Islamabad (CUI), Islamabad, Pakistan
2
Pakistan Meteorological Department, Islamabad, Pakistan

13
Vol.:(0123456789)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2990 Natural Hazards (2020) 103:2989–3010

desertification and land degradation, a situation that was observed between the 1960s and
1970s in the Sahel (Zeng 2003). Rising global mean annual temperatures resulting from
changing climate have further increased the frequency of droughts and dry conditions (Dai
et al. 2004; Sheffield et al. 2012).
Droughts can generally be divided into four major types based on the nature of water
deficiency. The precipitation deficiency over a long period leads to meteorological
droughts, and the shortage of soil moisture and water demands cause agricultural droughts,
whereas the scarcity of water in reservoirs and low water levels in dams and rivers over
a statistical average cause hydrological drought (Dracup et al. 1980). Collectively, these
three droughts create socioeconomic droughts, where the imbalance between demand and
supply of goods enforces the community to migrate (Wilhite and Glantz 1985). In recent
years, groundwater drought has been suggested as the fifth type of drought (Mishra and
Singh 2010).
Precipitation is one of the key triggering factors of droughts. Analysis of the precipita-
tion patterns in a region provides critical information about drought occurrence and prev-
alence. The standardized precipitation index (SPI) is known as a simple drought index,
widely used around the world to identify drought events (WMO 2012). This index uses
only precipitation to provide early warnings of drought and helps in drought severity
assessment. Based on different SPI timescales, different types of droughts such as meteoro-
logical, agricultural, and hydrological can be identified. SPI is successfully used by many
national hydrological and meteorological centers to determine the severity class of dryness
and wetness (Giddings et al. 2005). This study aims to determine the historical episode,
frequency, intensity, category, type, and return period of drought, identifying their criteria,
and develop an index to map drought vulnerability in 145 districts of Pakistan by using
long-term climatological datasets of rainfall and soil moisture. The study will also iden-
tify severe drought years and propagation of soil moisture anomaly, which ultimately helps
to identify the response of ground moisture during droughts years. The results will help
researchers, disaster managers, and policymakers to develop food security and drought
contingency plans of adaptation and mitigation for sustainable development.

2 Data and methodology

2.1 Study area

Pakistan lies in South Asia in the domain of (23°39 N–37°01 N and 60°49 E–77°40 E)
comprising of four provinces, namely Sindh, Punjab, Khyber Pakhtunkhwa (KP), and
Balochistan (Fig. 1). The Sindh and Punjab provinces receive most of the annual rain-
fall during monsoon season with very little rainfall in the winter. Balochistan, which
is the largest province, has a wide range of annual rainfall with winter and summer
precipitations dominant in districts receiving 71–231 mm and 112–402 mm, respec-
tively. KP and Federally Administered Tribal Areas (FATA) receive annual rainfall
between 250 and 1450 mm from south to north. The climate of Gilgit-Baltistan (GB)
province is very cold due to high elevation and mostly dominated by western distur-
bances, which causes rainfall in this region, whereas Azad Jammu and Kashmir (AJK)
receive the highest rainfall during the summer monsoon. The annual rainfall range in
GB is 200–500 mm and AJK is 850–1400 mm. Southern Pakistan experiences arid

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Natural Hazards (2020) 103:2989–3010 2991

Fig. 1  Topographical map showing the geographical location of meteorological stations

to extremely arid climate with high evapotranspiration rate that makes these regions
highly vulnerable to drought (Adnan et al. 2017a).

2.2 Precipitation and soil moisture datasets

Pakistan Meteorological Department (PMD) has 53 stations across the country as


shown in Fig. 1. R-ClimDex software was used to determine the quality check and
control of in situ precipitation data (Zhang et al. 2005). Spatial distributions of the
stations do not provide sufficient coverage for drought analysis at district-level hazard
mapping. To fill the gap between PMD stations, monthly gauged data (0.5° × 0.5°) of
precipitation from 1951 to 2010 are acquired from the Global Precipitation Climato-
logical Center (GPCC). These data have a very strong correlation with in situ station
rainfall data (Adnan et al. 2015) and used in many drought studies in previous years
due to their quality (e.g., Becker et al. 2013; Marengo et al. 2011; Kurnik et al. 2011;
Dai 2011). To determine the drought propagation, monthly soil moisture data of the
same resolution and duration as GPCC were retrieved from Climate Prediction Center
(CPC) (Fan and van den Dool 2008). The Drought Hazard Index (DHI) is developed by
analyzing the historical records of drought frequency, dependency on seasonal precipi-
tation, and soil moisture to prepare a district-level DHI map of Pakistan.

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2992 Natural Hazards (2020) 103:2989–3010

2.3 Drought hazard index (DHI)

Different drought indices are used worldwide to monitor drought. However, we aimed to
develop an equation that can be used to identify drought vulnerability over the study region.
For this purpose, we used rainfall, soil moisture, and the total number of drought years
from 1951 to 2010 to develop the equation. Keeping in view the climatology and depend-
ency on seasonal rainfall, a rating score was assigned for each of the classes for Sindh and
other regions (Table 1). Soil moisture is a key component to identify the drought. There-
fore, we incorporated the ratio of seasonal soil moisture to the annual soil moisture as a
component in the equation. Moreover, the ratio of the total number of drought years was
also used. The equation was developed by combining all these parameters which helped to
calculate the drought vulnerability in each district of Pakistan. Based on the above results,
the DHI was developed and five hazard classes ranging from extremely high to very low
have been introduced, namely; < 0.60 (very low), 0.60–0.75 (low), 0.75–1.0 (moderate),
1.00–1.50 (high), and > 1.50 (extremely high). This index was tested over 145 districts of
the country to develop a drought hazard map of Pakistan (Fig. 2). The following equation
was developed to calculate DHI:
( )
1 Td SMi−j
DHI = + SIndex + (1)
3 Ty SMAnnual

where Td is the total number of droughts; Ty is the total number of years; and SIndex is
the seasonal (winter/monsoon) dominant rainfall index (Table 2); for monsoon, ­SMi–j = soil
moisture (July to December), and for winter, S ­ Mi–j = soil moisture (January to June);
­SMannual = annual soil moisture.

2.4 Regional drought identification model (ReDIM)

Identifying the characteristics of drought over a local region is crucial for planning and
managing water resources and implementing preparedness and adaptation measures. The
ReDIM developed by Rossi and Cancelliere (2003) was used in this study to identify past
drought events, intensity, duration, and return period. ReDIM uses long-term precipitation
data (at least 30 years) to calculate standardized precipitation index (SPI) developed by
McKee et al. (1993) to determine the drought at various timescales. The probability at dif-
ferent timescales makes SPI a useful tool for monitoring the impacts of drought on agri-
culture and soil moisture over a short time and water reservoirs, streams, and rivers over
a long time. Based on different SPI timescales of 6, 9, and 12 months, meteorological,

Table 1  Criteria used for the Indicator Class limit and rating score
hazard assessment of drought
using percentage of normal
Percentage of sea- > 89 79–89 70–79 59–70
rainfall for different regions of
sonal rainfall for
Pakistan
Sindh province
Percentage of sea- > 60 51–60 41–50 30–40
sonal rainfall for
other province
Index value 4 3 2 1

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Natural Hazards (2020) 103:2989–3010 2993

Fig. 2  Drought hazard map showing the vulnerability index for each district of Pakistan

agricultural, and hydrological droughts were calculated, respectively (Adnan et al. 2015).
Based on regional analysis, ReDIM was used to calculate the areal extent and precipitation
severity in each region.

2.5 Statistical tests

Two statistical tests: Mann–Kendall τ-test (Mann 1945; Kendall 1975) and Sen’s slope
method (Sen 1968), are used to determine trend significance and magnitude, respectively,
at a 95% confidence level. The Mann–Kendall test (Eq. 2) is used to calculate the signifi-
cant linear trend for annual precipitation over each district.

∑ ∑
n−1 n
( )
S= sgn xj − xi (2)
i=1 j=i+1

where xi and xj are the data values of i and j (j > i), n is the number of data points, and
sgn(xj− xi) is the sign function. Variance is calculated using Eq. (3):
∑q � �� �
n(n − 1)(2n + 5) − p=1 tp − 1 2tp + 5
Var(S) = (3)
18
where n is the number of data points, q represents the number of tied groups, and tp indi-
cates the number of data points in the pth group. So, the Mann–Kendall significance (Z)
can be calculated by using Eq. (4):

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2994

13
Table 2  Drought monitoring indicator for Pakistan
Category of drought SPI Departure (%) Drought intensity
Sindh Balochistan Punjab KP FATA​ GB AJK

Dry period 3-months − 40 to- 80 − 25 to − 55 > − 20 − 15 to − 30 − 25 to − 45 − 20 to − 30 − 15 to − 25 Mild


> − 80 > − 55 > − 30 > − 45 Moderate
Meteorological 6 months − 40 to − 65 − 25 to − 35 − 20 to − 30 − 15 to − 25 − 20 to − 35 − 30 to − 40 − 15 to − 20 Mild
− 66 to − 80 − 36 to − 55 > − 30 > − 25 − 35 to − 50 > − 40 > − 20 Moderate
− 81 to − 95 > − 55 > − 50 Severe
> − 95 Extreme
Agriculture 9 months − 40 to − 65 − 25 to − 35 − 15 to − 25 − 10 to − 20 − 20 to − 35 − 20 to − 30 − 10 to − 20 Mild
− 66 to − 80 − 36 to − 55 > − 25 − 20 to − 30 − 35 to − 45 > − 30 > − 20 Moderate
− 81 to − 95 > − 55 > − 30 − 45 to − 55 Severe
> − 95 > − 55 Extreme
Hydrological 12 months − 40 to − 60 − 20 to − 30 − 15 to − 25 − 15 to − 25 − 20 to − 30 − 20 to − 30 − 10 to − 25 Mild
− 61 to − 75 − 31 to − 40 − 26 to − 35 − 25 to − 35 − 30 to − 40 − 30 to − 40 − 25 to − 40 Moderate
− 76 to − 85 − 41 to − 50 − 36 to − 50 − 35 to − 50 − 40 to − 50 − 40 to − 50 − 40 to − 50 Severe
> − 85 > − 50 > − 50 > − 50 > − 50 > − 50 > − 50 Extreme
Extreme hydrological 24 months − 25 to − 45 − 10 to − 20 − 10 to − 20 − 15 to − 25 − 15 to − 25 − 15 to − 25 − 15 to − 25 Mild
− 46 to − 60 − 21 to − 35 − 21 to − 35 − 25 to − 35 − 25 to − 35 − 25 to − 35 − 25 to − 35 Moderate
− 61 to − 75 − 36 to − 50 − 36 to − 50 − 35 to − 45 − 35 to − 45 − 35 to − 45 − 35 to − 45 Severe
> − 75 > − 50 > − 50 > − 45 > − 45 > − 45 > − 45 Extreme
Natural Hazards (2020) 103:2989–3010

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Natural Hazards (2020) 103:2989–3010 2995

⎧ √S−1 if S > 0
⎪ Var(S)
Z = ⎨0 if S = 0 (4)
⎪ √S+1 if S < 0
⎩ Var(S)

The positive and negative Z-values represent the increasing and decreasing trend,
respectively. The trend analysis is carried out at a 95% confidence level. The null hypoth-
esis means no significant trend is present (|Z| > 1.645). The linear trend of the slope is
determined by nonparametric statistical techniques proposed by Theil (1992) and Sen
(1968). This method calculates the slopes (Ti) of all data points, and the median of these N
values determines the Sen’s estimator (Ti) of slope (Eq. 5):
xj − x k
Ti = for i = 1, 2, 3, … N (5)
j−k

where xj and xk represent the data value at time j and k(j > k), respectively. However, the
trend magnitude (Q) of Sen’ Slope can be determined as by using Eq. (6):

⎧ T(N + 1) N is odd ⎫
⎪ � ∕2 � ⎪
Q=⎨ ⎬ (6)
⎪ 1∕2 T(N)∕ + T(N + 2)∕ N is even ⎪
⎩ 2 2 ⎭

The positive and negative values of Q correspond to increasing and decreasing trend,
respectively.

3 Results and discussions

3.1 Analysis of precipitation

A time-series analysis was conducted for each province to identify the correlations between
GPCC and meteorological station data from 1951 to 2010 (Fig. 3). The annual precipita-
tion analysis for Sindh province indicates that the precipitation of in situ data and GPCC
coincides well with each other (Fig. 3a). The time-series analysis depicts a slight overes-
timation in GPCC data (Fig. 3b) for Punjab and Balochistan provinces up to 1980. The
data coincide well with GPCC data in the later part of the twentieth century, probably due
to additional meteorological stations of PMD in the region (Fig. 3c). At the KP region, a
slight overestimation of precipitation was noted up to the 1960s but the agreement between
the two datasets improves in the later years (Fig. 3d). This may also be related to the estab-
lishment of more meteorological stations in KP. Similar patterns were also observed in the
GB and AJK regions where data agreement improves over time (Fig. 3e, f). The FATA sta-
tion rainfall data are the only region where GPCC data are underestimating throughout the
period (Fig. 3g); however, in situ data may not be reliable as only one station data are avail-
able for this region. Generally, station rainfall data coincide well with GPCC data which is
also evident in the rainfall analysis of Pakistan (Fig. 3h).
To determine the coefficient of determination (R2) for each of the regions, the regres-
sion line is drawn between GPCC and station annual rainfall data. The highest coef-
ficient of determination was observed for Sindh (R2 = 0.996), FATA (R2 = 0.915), AJK
(R2 = 0.892), Balochistan (R2 = 0.88), Punjab (R2 = 0.864), KP (R2 = 0.86), and GB

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2996 Natural Hazards (2020) 103:2989–3010

Fig. 3  Comparisons between annual GPCC and meteorological stations’ rainfall: a Sindh, b Punjab, c Balo-
chistan, d Khyber Pakhtunkhwa, e Gilgit-Baltistan, f Azad Jammu Kashmir, g FATA and h Pakistan during
1951–2010

(R2 = 0.831) (Fig. 4). The overall annual rainfall of GPCC and 53 in situ data shows a
strong relationship (R2 = 0.957). Furthermore, the high value of R2 signifies that almost
96% of the variation has been explained by predicting the outcome, indicating that the
GPCC data can be used if meteorological station data are not available for the country.
Two seasonal rainfall systems (monsoon and western disturbances) approach Pakistan
and contribute to most of the annual rainfall. The intraseasonal variability of rainfall is
one of the major causes of floods and droughts in Pakistan (Muslehuddin et al. 2005).
To determine the contribution of seasonal rainfall over Pakistan, the percent of nor-
mal rainfall is calculated for 145 districts of the country (Fig. 5). The analysis shows
that Sindh (78.2%) and Punjab (61.5%) provinces are highly dependent on summer

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Natural Hazards (2020) 103:2989–3010 2997

Fig. 4  Correlation between annual GPCC and meteorological stations’ rainfall: a Sindh, b Punjab, c Balo-
chistan, d Khyber Pakhtunkhwa, e Gilgit-Baltistan, f Azad Jammu Kashmir, g FATA, and h Pakistan during
1951–2010

monsoon, whereas the GB (41.6%) and Balochistan (43.7%) regions rely mostly on win-
ter rainfall for their annual rainfall. The KP and FATA receive rainfall in both summer
(46.4 and 34.8%) and winter (30.4 and 36.3%), respectively. Mann–Kendall τ test and
Sen’s slope analysis revealed significant positive (increasing) trends for 29 districts of
Pakistan, i.e., Punjab (13), FATA (6), KP (5), GB (4), Balochistan (1), while only one
district of AJK showed a negative trend (Fig. 6), which has been confirmed by previous
studies of Mondal et al. (2012) and Hanif et al. (2013). The results show that annual
rainfall has significantly increased in the majority of the northern and central districts

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2998 Natural Hazards (2020) 103:2989–3010

Fig. 5  Percentage of seasonal rainfall at districts of Pakistan during 1951–2010. The light-dark color bar
represents the percent of rainfall in winter and summer monsoon seasons, respectively

of Pakistan, with a maximum increase of 3.58 mm/year occurred at Mohmand Agency


(FATA) and a maximum decrease of − 1.87 mm/year occurred at Neelum (AJK).

3.2 Analysis of drought

To identify the climatic variation for a longer period, the percentage departure (PD) is cal-
culated for rainfall and soil moisture by using Eq. (7) as follows:
[( )]
Xi − X̄
PD(%) = × 100 (7)

where xi is the precipitation of ith month, X̄ is the normal precipitation. Then, the rela-
tionships between annual rainfall departure (ARD) and soil moisture departure (SMD)
with 12-month SPI are explored over different provinces of Pakistan (Fig. 7 and Table 3).
These results help to identify the temporal behavior of drought with respect to change in
precipitation and soil moisture. Strong correlations are observed between the ARD, SMD,
and 12-month SPI for all the provinces of Pakistan (Table 3). The highest correlation is
observed over Sindh (ARD-12-SPI), while lowest values were found over GB (SMD-12-
SPI). These results depict that ARD to 12-SPI and SMD are well coordinated with each
other. The results captured the variation in soil moisture and rainfall departure during the

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Natural Hazards (2020) 103:2989–3010 2999

Fig. 6  Spatial annual precipitation trends and magnitude of trend for Pakistan at 95% confidence level

historical drought years with high correlation and appeared as good indicators to monitor
drought over the region.
The analysis of precipitation and soil moisture variations shows that Sindh experiences
its worst drought during 1969, where the departures of rainfall and soil moisture were
− 84.14% and − 73.60%, respectively (Fig. 7a). Severe drought in Punjab was observed
during 1972, where the rainfall and soil moisture departures were − 29.63% and − 30.67%,
respectively (Fig. 7b). However, the maximum deficit in soil moisture (− 49.50%) was
observed in 2000 over Punjab province in which moderate drought condition was expe-
rienced. In Balochistan and FATA, the maximum deficit in soil moisture and rainfall
was observed in 2000 indicating extreme drought (Fig. 7c, g). The years 2000 and 2001
were the most intense drought years on record for FATA. KP and GB experienced severe
droughts in 2001 as indicated by the lowest SPI (− 1.86 and − 1.2, respectively) due to rain-
fall departures, which, in turn, reduced the soil moisture as shown in Fig. 7d, e. AJK expe-
rienced drought conditions in 2002 where the rainfall and soil moisture departures were
− 26% and − 31%, respectively (Fig. 7f). Overall, the country suffered from severe drought
from 2000 to 2002 (Fig. 7h). The results also indicate that rainfall deficiency leads to lower
SPI values and reduced soil moisture, which ultimately enhances the drought severity over
an area (Wu et al. 2001).
The spatial correlation is calculated between soil moisture departures (%) and SPI at
different timescales of 3, 6, 9, 12, and 24 months for each district of Pakistan (Fig. 8).
The districts of central and southern Pakistan showed a strong relationship between the

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
3000 Natural Hazards (2020) 103:2989–3010

Fig. 7  Time-series comparisons among 12-month SPI, percentage departure of soil moisture and rainfall in
Pakistan (1951–2010)

SMD and SPI, which becomes stronger as the timescale progresses (Kumar et al. 2009).
However, the correlation decreases at a long timescale of 24 months for the whole
region. The correlation in the northern region becomes weaker due to high variability
in precipitation and soil moisture which makes these areas less susceptible to drought.
The provinces of Sindh, Balochistan Punjab, KP, FATA, and AJK showed a strong cor-
relation between soil moisture departure (%) and SPI at 9 and 12 months for Balochistan
and GB. The highest correlation between SMD and SPI was noticed for Sindh (Larkana,
r9-SPI = 0.86), Balochistan (Nasirabad, r12-SPI = 0.85), Punjab (Mianwali r9-SPI = 0.81),
KP (Lakki Marwat, r9-SPI = 0.81), FATA (FR Tank, r9-SPI = 0.76), GB (Hunza Nagar,
r12-SPI = 0.53), and AJK (Mirpur r9-SPI = 0.61).

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Natural Hazards (2020) 103:2989–3010 3001

Table 3  Correlation analysis Province/state Correlation


between ARD, 12SPI, and SMD
for different regions of Pakistan ARD-12SPI ARD-SMD SMD-12SPI

Sindh 0.97 0.80 0.77


Punjab 0.64 0.62 0.68
Balochistan 0.76 0.70 0.81
KP 0.77 0.67 0.73
GB 0.88 0.77 0.36
AJK 0.60 0.50 0.56
FATA​ 0.85 0.77 0.78
Pakistan 0.68 0.68 0.72

The study also classified the rainfall departure (%) on different timescales to identify
the intensity and category of droughts in Pakistan (Table 2). Keeping in view the climato-
logical behavior of precipitation for each of the provinces, the different range of percent-
age departure (%) along with drought category and intensity was defined at different time-
scales. The advantage of using rainfall departure (%) is that it can easily be calculated, and
it does not require long-term data like other drought indices such as SPI, SPEI, and RDI.
By analyzing historical data of 60 years (1951–2010) of precipitation, it was discovered
that the drought intensity (mild to extreme) depends upon the selection of timescale dura-
tion and high-intensity droughts (severe and extreme) cannot be calculated on a smaller
timescale of less than 3 months. The drought frequency analysis provides spatial details
regarding the intensity and severity of drought over the country. Based on 3-month SPI,
mild and moderate droughts were found to be more frequent in northern and southern parts
of the country (Fig. 9). Total drought frequency was high in the southern region includ-
ing Sindh, southern Punjab, and some parts of Balochistan. The long-term historical data
analysis shows that the severity of drought cannot be determined at short timescale (1-, 2-,
3-month SPI). The highest frequency of mild drought was observed in most parts of Pun-
jab, northwestern and southeastern Balochistan, southern KP and FATA, and northern GB
and AJK. The highest frequency of moderate drought was observed in Sindh and southern
Punjab.
Most of the Pakistan comprises two wet periods (January–March and July–September)
led by two dry periods (April–June and October–December), whereas in some parts, only
one wet period (comprises 3 months) prevails during a year. The 6-month SPI showed that
high frequency of mild drought is prevalent over most of the central and southern Paki-
stan, whereas moderate droughts are more frequent in the west to southwest parts including
Sindh. Similarly, severe and extreme drought frequencies were high in central and south-
ern Pakistan. The total drought frequency in the southern part of the country including
Balochistan, Sindh, and southern Punjab was very high, which makes these regions more
vulnerable to meteorological drought (Fig. 10). The results of 9-month SPI showed that
the mild drought frequency is very high over central and northern parts, whereas moderate
drought and severe drought are frequent in the southern part of the country (Fig. 11). The
frequency of extreme drought was very low. The total drought frequency in the agriculture
sector was very high over different areas of the country. These regions depend on a single
weather system, and deficiency in seasonal rainfall may cause drought due to which agri-
cultural activity remains at risk.

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
3002 Natural Hazards (2020) 103:2989–3010

Fig. 8  Spatial distribution of correlation between SMD and SPI (3, 6, 9, 12, and 24 months) for Pakistan
districts

Based on 12-month SPI, the mild drought was more frequent in most of the northern
parts, whereas moderate, severe, and extreme droughts were more frequent in the south-
ern part (Fig. 12). The drought frequency was generally high in most parts of the coun-
try. Due to the low amount of seasonal rainfall, the southwestern part remains highly
vulnerable to hydrological drought. Based on these results, high-frequency drought over
different regions of Pakistan including Sindh, Balochistan, central Punjab, and southern
KP was noted. The extreme hydrological drought was calculated over Pakistan based on

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Natural Hazards (2020) 103:2989–3010 3003

Fig. 9  Drought frequency based on 3-month SPI (dry period) of Pakistan over 1951–2010

24-month SPI (Fig. 13). The frequency of mild drought was prevalent in northern parts,
whereas the intense (moderate, severe, and extreme) droughts were common in the
south. The results show that the total drought frequency was very high over most of the
agricultural areas of central and southern parts which strongly affects water reservoirs
in the country. The longest episode of drought was experienced during 2000–2002 over
entire South Asia, including Pakistan. The analysis of SPI on different timescales (i.e.,
3, 6, 9, 12, and 24 months) indicates that the frequency of severe drought was more in
southern districts of Pakistan. The climate of southern Pakistan is arid to extremely arid
(Adnan et al. 2017b), and crop water demands remain very high due to high tempera-
ture and evapotranspiration. The low rainfall and high evaporative demand of the atmos-
phere deplete the soil moisture and increase the probability of intense drought. As the

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
3004 Natural Hazards (2020) 103:2989–3010

Fig. 10  Drought frequency based on 6-month SPI (meteorological drought) of Pakistan over 1951–2010

crop water requirement is very high, proper irrigation is required for agricultural activ-
ity in this region. The study has identified the most severe drought years as 1952, 1969,
1971, 2000, 2001, and 2002 over Pakistan. The soil moisture anomaly was also calcu-
lated (Fig. 14), and the results show that most parts of the country were under drought
during the severe drought years. However, the southern region was most vulnerable to
drought. The soil moisture deficit ranges − 50 to − 75% in most of the affected districts
in central Pakistan, and − 100% during the rest of the drought years, especially in south-
ern Pakistan, which was well below normal. The spatial analysis identified 2000–2002
as the longest drought period. This consecutive 3-year drought covered two-third of the

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Natural Hazards (2020) 103:2989–3010 3005

Fig. 11  Drought frequency based on 9-month SPI (agriculture drought) of Pakistan over 1951–2010

country with intensity ranging from moderate to extreme. It has also been noted that soil
moisture is a good indicator of drought as indicated by Adnan et al. (2015).
Using the climatological data (1951–2010), each district has been classified into
a drought category based on the return period of drought intensity and its percentage
affected area (Table 4). The results show that the extreme drought of CAT-IV was expe-
rienced in most regions including southern Sindh, southwestern Balochistan, southern
Punjab, KP, and FATA where more than 80% of the area was affected. Most of the CAT-I
and CAT-IV extreme droughts were experienced in northern areas like Skardu (GB) and
Kotli (AJK). These results were used to prepare the drought hazard map that can be used
for developing drought contingency plans. The map shows that the country is vulnerable
to drought, but the southern parts are more vulnerable to drought as compared to the

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
3006 Natural Hazards (2020) 103:2989–3010

Fig. 12  Drought frequency based on 12-month SPI (hydrological drought) of Pakistan over 1951–2010

northern parts. Out of 145 districts assessed, 19 districts were extremely vulnerable, 24
were high, 56 were moderate, 33 were low, and 12 were very less vulnerable (Fig. 1).
The provincial results show that 15 districts in Balochistan, three in Punjab, and one in
Sindh were highly vulnerable to drought. Most of the highly vulnerable districts belong
to Balochistan, Sindh, and southern Punjab. The southwestern and southeastern parts of
the country were more susceptible to drought as these regions receive a major portion
of annual rainfall during winter and summer, respectively. Hence, the deficiency of sea-
sonal rainfall might have caused high-intensity droughts in these districts. The districts
of northern Pakistan were less vulnerable to drought due to the two seasonal rainfall
systems (i.e., summer monsoon and western disturbance).

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Natural Hazards (2020) 103:2989–3010 3007

Fig. 13  Drought frequency based on 24-month SPI (extreme hydrological drought) of Pakistan over 1951–
2010

4 Conclusions

This study provides information regarding the drought vulnerability distribution in Paki-
stan during 1951–2010. The GPCC data showed a better performance with station data of
precipitation over the whole provinces (maximum in Sindh, R2 = 0.996) as well as in the
country (R2 = 0.957). The Sindh and GB regions highly depend upon monsoon (78%) and
winter (42%) rainfall, respectively. The deficiency in seasonal rainfall is one of the major
reasons for drought occurrence in the country. The rainfall has decreased (− 3.34 mm/year)
over some areas of AJK, though not statistically significant, but significantly increased
over Punjab, KP, and FATA (maximum of 3.58 mm/year) at 95% confidence level. Based

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
3008 Natural Hazards (2020) 103:2989–3010

Fig. 14  Soil moisture departures (%) for the most severe historical drought years in Pakistan

Table 4  Drought categories, percentage of area affected, and intensity return period for districts of Pakistan
Drought category Affected area (%) Mild Moderate Severe drought Extreme drought
drought drought (year) (year)
(year) (year)

CAT-1 < 40 3–4 6–10 Once in 15 Once in 20


CAT-II 40–60 5–6 11–15 16–20 21–25
CAT-III 60–80 7–8 16–20 21–30 25–35
CAT-IV > 80 >8 >20 >30 >35
CAT-I: the percentage of area drought affected is less than 40%
CAT-II: the percent of affected area is between 40% and 60%
CAT-III: the percent of area affected is between 60% and 80%
CAT-IV: the percent of area affected is greater than 80%

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Natural Hazards (2020) 103:2989–3010 3009

on different SPI timescales of 3, 6, 9, 12, and 24 months, the drought type, intensity, and
frequency were calculated, which showed that mild drought was more frequent in north-
ern Pakistan, while intense and severe droughts were prevalent over central and southern
Pakistan. The SPI results showed that high-intensity drought cannot be determined at the
smaller timescale of less than 3 months, which makes it much more realistic. Using the cli-
matological data of SPI and soil moisture, the most intense drought years which cover two-
third part of the country were identified as 1952, 1969, 1971, 2000, 2001, and 2002. The
DHI equation helps to identify the drought pones regions and may be used to evaluate the
hazard assessment over a small area. Based on DHI, drought hazard map over 145 districts
has been developed which depicts 19 districts are extremely vulnerable, 24 are high, 56 are
moderate, 33 are low, and 12 are very less vulnerable. Moreover, a total of 19 districts of
Pakistan, i.e., one in Sindh, 15 in Balochistan, and three in Punjab, are extremely vulner-
able to drought. Thus, this study provides a gateway to disaster managers, food security
planners, agriculturists, agronomists, researchers, and policymakers to develop a contin-
gency plan of adaptation and mitigation of drought challenges in the country.

Acknowledgements This study was supported by the Higher Education Commission of Pakistan (Grant
Nos: 8035/Balochistan/NRPU/R&D/HEC/2017 and 2185 (Startup Research Grant)). The comments of M.
Umar, A. H. Khan, and the anonymous reviewers were helped improve the paper. Moreover, the authors
thank the Pakistan Meteorological Department for providing the data.

References
Adnan S, Ullah K, Gao S (2015) Characterization of drought and its assessment over Sindh-Pakistan during
1951-2010. J Meteor Res 29:837–857
Adnan S, Ullah K, Khan AH, Gao S (2017a) Meteorological impacts on evapotranspiration in different cli-
matic zones of Pakistan. J Arid Land 9:938–952
Adnan S, Ullah K, Gao S, Khosa AH, Wang Z (2017b) Shifting of agro-climatic zones, their drought vul-
nerability, and precipitation and temperature trends in Pakistan. Int J Climatol 37(S1):529–543
Becker A, Finger P, Meyer-Christoffer A, Rudolf B, Schamm K, Schneider U, Ziese M (2013) A description
of the global land-surface precipitation data products of the Global Precipitation Climatology Centre
with sample applications including centennial (trend) analysis from 1901–present. Earth Syst Sci Data
5(1):71–99
Dai A (2011) Drought under global warming: a review. Wires Clim Change 2(1):45–65
Dai A, Trenberth KE, Qian T (2004) A global dataset of Palmer Drought Severity Index for 1870–2002:
relationship with soil moisture and effects of surface warming. J Hydrometeorol 5(6):1117–1130
Dracup JA, Lee KS, Paulson EG Jr (1980) On the definition of droughts. Water Resour Res 16(2):297–302
Fan Y, Van den Dool H (2008) A global monthly land surface air temperature analysis for 1948–present. J
Geoph Res 113:1103. https​://doi.org/10.1029/2007J​D0084​70
Giddings L, Soto M, Rutherford BM, Maarouf A (2005) Standardized precipitation index zones for Mexico.
Atmósfera 18(1):33–56
Guha-Sapir D (2017) EM-DAT: the emergency events database. (http://emdat​.be/emdat​_db/)
Hanif M, Khan AH, Adnan S (2013) Latitudinal precipitation characteristics and trends in Pakistan. J
Hydrol 492:266–272
Kendall MG (1975) Rank correlation methods, 4th edn. Charles Griffin, London
Kogan FN (1997) Global drought watch from space. Bull Am Meteorol Soc 78(4):621–636
Kumar MN, Murthy CS, Sesha Sai MVR, Roy PS (2009) On the use of Standardized Precipitation Index
(SPI) for drought intensity assessment. Meteorol Appl 16(3):381–389
Kurnik B, Barbosa P, Vogt J (2011) Testing two different precipitation datasets to compute the standardized
precipitation index over the Horn of Africa. Int JRemote Sen 32(21):5947–5964
Mann HB (1945) Nonparametric tests against trend. Economet J Economet Soc 13:245–259
Marengo JA, Tomasella J, Alves LM, Soares WR, Rodriguez DA (2011) The drought of 2010 in the con-
text of historical droughts in the Amazon region. Geophy Res Lett 38(12):L12703. https​://doi.
org/10.1029/2011G​L0474​36

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
3010 Natural Hazards (2020) 103:2989–3010

McKee TB, Doeskin NJ, Kleist J (1993) The relationship of drought frequency and duration to time scales.
In: 8th Conference on applied climatology. Amer Meteor Soc, Anaheim CA, pp 179–184
Mishra AK, Singh VP (2010) A review of drought concepts. J Hydrol 391(1):202–216
Mondal A, Kundu S, Mukhopadhyay A (2012) Rainfall trend analysis by Mann-Kendall test: a case study of
north-eastern part of Cuttack district, Orissa. Int J Geol Earth and Environ Sci 2(1):70–78
Muslehuddin M, Mir H, Faisal N (2005) Sindh summer (June-September) monsoon rainfall prediction. Pak
J Meteorol 2(4):91–108
Rossi G, Cancelliere A (2003) At-site and regional drought identification by ReDIM model. In: Rossi G et al
(eds) Tools for drought mitigation in mediterranean regions. Kluwer Academic Publishing, Dordrecht,
pp 37–54
Sen PK (1968) Estimates of the regression coefficient based on Kendall’s tau. J Am Stat Associ
63(324):1379–1389
Sheffield J, Wood EF, Roderick ML (2012) Little change in global drought over the past 60 years. Nature
491(7424):435–438
Theil H (1992) A rank-invariant method of linear and polynomial regression analysis. In: Raj B, Koerts J
(eds) Henri Theil’s contributions to economics and econometrics. Springer, Dordrecht, pp 345–381
Vogt JV, Safriel U, Von Maltitz G, Sokona Y, Zougmore R, Bastin G, Hill J (2011) Monitoring and assess-
ment of land degradation and desertification: towards new conceptual and integrated approaches. Land
Degrad Dev 22(2):150–165
Wilhite DA, Glantz MH (1985) Understanding: the drought phenomenon: the role of definitions. Water Int
10(3):111–120
World Meteorological Organization (WMO) (2012) Standardized Precipitation Index User Guide (Svoboda
M, Hayes M, Wood D), (WMO-No. 1090), Geneva
Wu H, Hayes MJ, Weiss A, Hu Q (2001) An evaluation of the Standardized Precipitation Index, the China-Z
Index and the statistical Z-Score. Int J Climatol 21(6):745–758
Zeng N (2003) Drought in the Sahel. Science 302(5647):999–1000
Zhang X, Hegerl G, Zwiers FW, Kenyon J (2005) Avoiding inhomogeneity in percentile-based indices of
temperature extremes. J Clim 18(11):1641–1651

Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center
GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers
and authorised users (“Users”), for small-scale personal, non-commercial use provided that all
copyright, trade and service marks and other proprietary notices are maintained. By accessing,
sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of
use (“Terms”). For these purposes, Springer Nature considers academic use (by researchers and
students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and
conditions, a relevant site licence or a personal subscription. These Terms will prevail over any
conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription (to
the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of
the Creative Commons license used will apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may
also use these personal data internally within ResearchGate and Springer Nature and as agreed share
it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not otherwise
disclose your personal data outside the ResearchGate or the Springer Nature group of companies
unless we have your permission as detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial
use, it is important to note that Users may not:

1. use such content for the purpose of providing other users with access on a regular or large scale
basis or as a means to circumvent access control;
2. use such content where to do so would be considered a criminal or statutory offence in any
jurisdiction, or gives rise to civil liability, or is otherwise unlawful;
3. falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association
unless explicitly agreed to by Springer Nature in writing;
4. use bots or other automated methods to access the content or redirect messages
5. override any security feature or exclusionary protocol; or
6. share the content in order to create substitute for Springer Nature products or services or a
systematic database of Springer Nature journal content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a
product or service that creates revenue, royalties, rent or income from our content or its inclusion as
part of a paid for service or for other commercial gain. Springer Nature journal content cannot be
used for inter-library loans and librarians may not upload Springer Nature journal content on a large
scale into their, or any other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not
obligated to publish any information or content on this website and may remove it or features or
functionality at our sole discretion, at any time with or without notice. Springer Nature may revoke
this licence to you at any time and remove access to any copies of the Springer Nature journal content
which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or
guarantees to Users, either express or implied with respect to the Springer nature journal content and
all parties disclaim and waive any implied warranties or warranties imposed by law, including
merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published
by Springer Nature that may be licensed from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a
regular basis or in any other manner not expressly permitted by these Terms, please contact Springer
Nature at

onlineservice@springernature.com

You might also like