You are on page 1of 11

Journal of Hydrology 623 (2023) 129877

Contents lists available at ScienceDirect

Journal of Hydrology
journal homepage: www.elsevier.com/locate/jhydrol

Quantitative evaluation of drought risk related to vegetation productivity


in China
Wenyan Ge a, b, Xiuxia Li c, Mengxia Xie a, b, Bowen Yu d, Juying Jiao a, b, Jianqiao Han a, b,
Fei Wang a, b, *
a
Institute of Soil and Water Conservation, Northwest A&F University, Yangling 712100, Shaanxi Province, China
b
Institute of Soil and Water Conservation, Chinese Academy of Sciences & Ministry of Water Resources, Yangling 712100, Shaanxi Province, China
c
School of Management, Tianjin University of Commerce, Tianjin 300134, China
d
China Academy of Electronics and Information Technology, Beijing 100041, China

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

Handling Editor: Emmanouil Anagnostou Quantitative assessment of vegetation drought risk is crucial for addressing drought problems and vegetation
management under global warming. In this study, we employed a framework based on long-term (2001–2019)
Keywords: vegetation Gross Primary Productivity (GPP) and the Standardized Precipitation Evapotranspiration Index (SPEI)
Vegetation to quantify drought risk, considering vegetation vulnerability, exposure, and hazard to drought. The results
Drought risk assessment
revealed that vegetation drought risk was influenced by drought hazard (55.15%), exposure (24.46%), and
Hazard
vulnerability (20.39%). Approximately 48% of study pixels experienced moderate to high drought risk, primarily
Vulnerability
Exposure concentrated in northeastern to southwestern China. In terms of vegetation types, forest suffered the highest
drought risk, followed by cropland, while grassland had the mildest drought risk. Notably, croplands in northern
China experienced relatively high drought risk, posing a threat to food security. Therefore, it is imperative to
consider appropriate crop management practices such as irrigation and crop rotation to mitigate the impact of
droughts. Temperature and precipitation are key climatic factors affecting vegetation drought risk in moderate-
to-high risk areas. Warming inhibited drought risk by promoting vegetation growth, but also consumed water to
promote drought risk. Meanwhile, increased rainfall in parts of northern China may be caused by extreme
weather events that intensify vegetation vulnerability, thus aggravating drought risk. Our research provided
scientific guidance to understand and cope with drought risk induced by climate change on vegetation.

1. Introduction with climate change, further aggravates water scarcity, amplifying the
occurrence and severity of droughts globally (Abhishek and Kinouchi,
Global climate change worsens extreme hydrological events, such as 2022; He et al., 2021).
floods and droughts, with devastating consequences for agriculture, Drought risk results from the intricate interplay of drought hazard,
ecosystems, economies, and societies (Abhishek and Kinouchi, 2022; vulnerability, and exposure, leading to potential adverse consequences
Song et al., 2019; van Dijk et al., 2013; Abhishek and Sayama, 2021). For or losses (Carrão et al., 2016; Edenhofer, 2015). Drought hazard, the
instance, unprecedented flooding in the Chao Phraya River Basin in likelihood of drought occurrences (Carrão et al., 2016), could be un­
2011 killed more than 800 people and caused economic losses of more derstood as the frequency, duration, intensity, and spatial range of
than $45 billion (Abhishek and Sayama, 2021), while persistent drought (Lavell et al., 2012; Li et al., 2020). The Probability Density
droughts in India exaggerate water scarcity (Abhishek and Kinouchi, Function (PDF) coupled with relevant drought indices has proven
2022; Xiong et al., 2022). Drought, the most lethal and costly disaster effective in assessing drought hazard (Islam et al., 2017). Drought
among natural disasters, is becoming more frequent and intense due to exposure pertains to the number of entities, such as populations and
climate change (Carrão et al., 2016; Van Dijk et al., 2013). Monitoring livelihoods, susceptible to droughts (Field et al., 2012). Drought
and managing drought is challenging due to its prolonged duration, vulnerability relates to the capacity to withstand and recover from
gradual effects, and diffuse spatial distribution. Urbanization, coupled drought impacts (Zhao et al., 2020).

* Corresponding author at: Institute of Soil and Water Conservation, Northwest A&F University, Yangling 712100, Shaanxi Province, China.
E-mail address: wafe@ms.iswc.ac.cn (F. Wang).

https://doi.org/10.1016/j.jhydrol.2023.129877

Available online 28 June 2023


0022-1694/© 2023 Elsevier B.V. All rights reserved.
W. Ge et al. Journal of Hydrology 623 (2023) 129877

Vegetation is a crucial component of the terrestrial ecosystem, un­ guiding effective mitigation strategies.
dergoing significant shifts due to global climate change (Duo et al., In this study, building on the existing scheme proposed by Li et al.
2016; Ge et al., 2022; Piao et al., 2010). Droughts can severely hinder (2020), we developed a framework of drought risk to assess the impacts
plant growth, causing damage or even death (Wu et al., 2018). As global of drought on vegetation by quantifying three independent de­
warming continues and groundwater is overexploited (such as Penin­ terminants: drought hazard, vulnerability and exposure (Carrão et al.,
sular India, (Xiong et al., 2022)), drought caused by climate change has 2016; Zhang et al., 2017b; Zhao et al., 2020) in different climate zones in
intensified and far-reaching impacts on vegetation (Xu et al., 2019), China, based on GPP and the Standardized Precipitation Evapotranspi­
posing risks to environmental balance and sustainable development ration Index (SPEI). In addition, we investigated the influence of cli­
(Hanewinkel et al., 2013; Wan et al., 2018). matic factors on GPP and GPP-related drought risk in China. The
Gross Primary Productivity (GPP) quantifies the amount of carbon primary purposes of this study were: (1) to evaluate vegetation drought
fixed by terrestrial ecosystems through photosynthesis, serving as a risk by quantifying drought hazard, exposure and vulnerability in China;
direct indicator of vegetation growth and health (Xiao et al., 2019; (2) to analyze its characteristics in different climate-related vegetation
Padfield et al., 2017). Under the background of climate change, persis­ zones; (3) to identify the main climatic factors influencing vegetation
tent and severe drought events caused huge destruction to the composite drought risk.
and structure of ecosystems, weakening the carbon sink function of the
ecosystem, leading to reduction in GPP (Wei et al., 2023; Zhang et al., 2. Study area and datasets
2023; Frank et al., 2015). Furthermore, droughts are expected to have
greater impacts on GPP in future climate models (Xu et al., 2019). 2.1. Study area
Evaluating the influence of drought risk on GPP offers valuable insights
into overall ecosystem health and functionality under dry conditions, China, with its extensive territory and diverse climate, exhibits a
particularly in the context of climate change. complex climatic landscape. Based on the current Köppen-Geiger
Being the most populous and rapidly urbanizing country in the climate classification (Beck et al., 2018), China can be categorized into
world, China faces a serious water crisis and water-related disasters (e. five primary climate groups: tropical (A), dry (B), temperate (C), con­
g., drought) due to the severe freshwater deficit and uneven distribution tinental (D), and polar (E) (Fig. S1a). These primary groups can be
of water resources (He et al., 2021; Wang et al., 2020; Zhou et al., 2023). further subdivided into twelve climate-related vegetation subgroups,
Implementing afforestation in China since 1999 has substantially taking into account the distribution of various vegetation types (Fig. 1;
improved vegetation cover (Ge et al., 2021; Zhou et al., 2021). However, (Liu et al., 2002)), including Af (forest in tropic), Bc (cropland in dry), Bf
the significant water consumption associated with these large-scale (forest in dry), Bg (grass in dry), Cc (cropland in temperate), Cf (forest in
projects has also been identified as a major factor exacerbating temperate), Cg (grass in temperate), Dc (cropland in continental), Df
droughts, particularly in northern China (Feng et al., 2016; Zhou et al., (forest in continental), Dg (grass in continental), Ef (forest in polar), Eg
2023). Meanwhile, increased temperature caused by climate change also (grass in polar). Other subgroups (i.e., cropland and grass in tropical,
aggravates droughts in China by disrupting precipitation patterns and and cropland in polar) each take up a small percentage of area (<0.3%),
increasing evapotranspiration (Gosling and Arnell, 2016; Zhang et al., and were thereby excluded from this study. Additionally, this study
2020; Zheng et al., 2022). Severe and recurrent drought events in China excluded vast desert regions in northwestern China, such as the Takli­
have caused significant ecological damage to terrestrial ecosystems makan and Badain Jaran deserts, due to the absence of vegetation in
(Fang et al., 2019; Jiang et al., 2021; Song et al., 2019). Quantitative these areas (Zhang et al., 2017a).
assessment of drought impact on vegetation in China is crucial for un­
derstanding ecosystem vulnerability and response to drought, and

Fig. 1. The Köppen-Geiger climate and vegetation type classification map of China (referenced from Beck et al., 2018), and the proportions of vegetation in climate
zones depicted by the pie chart.

2
W. Ge et al. Journal of Hydrology 623 (2023) 129877

2.2. Datasets Higher DVi value suggests more vulnerable vegetation to drought
and vice versa.
Here, we obtained the long-term (2001–2019) Terra GPP Gap-Filled Here, we used the drought resistance feature of vegetation docu­
8-Day L4 Global 500 m SIN Grid (MOD17A2HGF v006) dataset and mented by Zhang et al. (2017a) to represent the drought sensitivity of
Vegetation Continuous Fields (VCF) Yearly L3 250 m SIN Grid products vegetation (DS):
obtained from the National Aeronautics and Space Administration ( )
Ri,j = corr GPPi , SPEI i,j 4 ≤ i ≤ 10, 1 ≤ j ≤ 24 (4)
(NASA). The MODIS Gap-Filed GPP data and VCF products were
downloaded The GPP data were improved by filling the cloud- ( )
DSi = max Ri,j (5)
contaminated LAI/FPAR gaps before calculation (DiMiceli et al., 2015;
Running and Zhao, 2015). Besides, the weekly Vegetation Health Index
where i refers the ith month, ranging from April to October; j represents
(VHI) data during 2001–2019 at 4 km spatial resolution were derived
the drought timescale, ranging from 1 to 24 months; Ri, j is the Pearson
from the National Oceanic and Atmospheric Administration (NOAA).
correlation coefficients of GPP and SPEI in the ith month with the
The monthly VHI dataset was derived using the maximum value com­
timescale of j month; and DSi represents the drought sensitivity of
posite (MVC) method.
vegetation productivity in the ith month.
We acquired monthly precipitation and mean temperature data at a
DA is the adaptability of vegetation to drought (Li et al., 2020):
1 km spatial resolution from Peng et al. (2019), which has been vali­
dated using observations at the weather station across China and widely n ⃒ ⃒
1∑ ⃒(GPPi − GPPmean )/GPPmean ⃒
applied in carbon budget accounting and ecosystem service evaluation DA = ⃒ ⃒, SPEI ≤ − 1.0 (6)
n i ⃒ SPEI i − SPEI mean ⃒
(Ge et al., 2022; Li et al., 2023; Yu et al., 2022). The monthly downward
surface shortwave radiation and potential evapotranspiration with 4 km where GPPi and SPEIi are timeseries of GPP and SPEI-12 in pixel i under
spatial resolution was obtained from the Climatology Lab (Abatzoglou drought conditions (SPEI ≤ − 1.0), respectively. GPPmean and SPEImean
et al., 2018). To ensure data consistency, all datasets in this study were are the mean values of GPP and SPEI during the study period, respec­
resampled using the nearest neighbor interpolation method to a 4-km tively. Higher DA value suggests less adaptable vegetation to drought.
spatial resolution and one-month temporal resolution. Drought exposure refers to the number of physical entities located in
The multiscalar drought index SPEI has been proved to be more an area where drought events are likely to occur (Field et al., 2012; Zhao
effective for drought types and impacts identification than other drought et al., 2020). For vegetation, the drought exposure can be quantified by
indices such as the Standardized Precipitation Index (SPI) and Palmer the density of vegetation (VD) and the health condition of vegetation
Drought Severity Index (PDSI), especially their sensitivity to droughts (VHI):
(Wei et al., 2023; Zhao et al., 2018; Zhang et al., 2023; Zhang et al.,
2017a). Therefore, we utilized SPEI to evaluate drought risk on GPP in 1∑ n
DEi = VDi,y × (100 − VHI i ) (7)
this study. To calculate SPEI at timescales ranging from 1 to 24 months, n y=1
we utilized monthly grid data of temperature and potential evapo­
transpiration and employed the SPEI calculator provided by Vicente- where VDi,y is the density of vegetation in pixel i of yth year, VHIi is the
Serrano et al. vegetation health index of the study period in pixel i, and DEi is the
exposure of vegetation to drought in pixel i. A higher DE value dem­
2.3. Methodology onstrates a higher exposure.
Here, we used the vegetation structure quantitative product MODIS
2.3.1. Drought risk assessment VCF to quantify the density of vegetation (Hansen et al., 2003). Spe­
Drought risk (DR) is the resulting representation of three indepen­ cifically, the monthly variations of vegetation were directly represented
dent determinants (Edenhofer, 2015): by the yearly trees VCF data because of the perennial characteristics of
woody and its stable amount in the short term. Conversely, the density of
1 n∑
=10
herbaceous vegetation (i.e., grass and crop) is unstable within a year
DR = × DH i × DV i × DEi (1)
7 i=4 because of their annual characteristics. The monthly vegetation density
(VD) therefore were calculated as follows:
where DHi, DVi and DEi represent drought hazard, drought vulnerability
VDx,y = VCF tree,y + VCF non− × RatioGPP (8)
and drought exposure in the ith month, respectively, with the growing
tree,y

season, i.e., from April to October (Piao et al., 2010). Here, we focused GPPx,y
on the drought risk of vegetation caused by meteorological droughts. RatioGPP = , 0≤ RatioGPP ≤ 1 (9)
GPPy,mean
Drought hazard (DH) is determined by drought frequency and
intensity: where VDx,y is the density of vegetation in xth month of yth year, VCFtree,y
∫ − 1.0 and VCFnon-tree, y are the yth year of the tree and non-tree VCF, respec­
DH = SPEI × f (SPEI) (2) tively. GPPx,y is the GPP in xth month of yth year, and GPP y, mean is the
mean value of GPP in yth year.
− 3.5

∫ − 1.0 To ensure data consistency, the monthly DH, DS, DA, DE and DR were
where − 3.5 SPEI is the drought frequency, and f(SPEI) is the drought
standardized to a dimensionless quantity within the range of 0–1 for the
intensity. The SPEI value less than − 1.0 in a specific period indicates
assessment:
drought condition. Hence, the drought frequency here is obtained by
integrating SPEI less than − 1.0 during the study period. Additionally, t′ =
t − tmin
(10)
SPEI-12 (i.e., SPEI with 12-month timescale) were used to calculate the tmax − tmin
drought frequency and intensity to evaluate drought hazard caused by
meteorological droughts (Islam et al., 2017; Li et al., 2020). where tmax and tmin are the maximum and minimum values of the same
The vulnerability of vegetation to drought (DV) could be character­ variable in all pixels from April to October, respectively.
ized by vegetation drought sensitivity (DS) and adaptability (DA) (Gao
et al., 2018; Li et al., 2020): 2.3.2. Analyses
Partial correlation examines the correlation between two variables
DV i = DSi × DAi (3) while accounting for the influence of one or more additional variables

3
W. Ge et al. Journal of Hydrology 623 (2023) 129877

(Adeyeri et al., 2022). It allows us to assess whether the observed rela­ 3. Results
tionship between the two variables is direct or influenced by the addi­
tional variables. We utilized partial correlation analysis to investigate 3.1. Drought risk components assessment
the key risk components (hazard, exposure, and vulnerability) as well as
climatic factors (temperature, precipitation, and solar radiation) in 3.1.1. Drought hazard
relation to drought risk: High drought hazard predominantly affected regions in south­
western China, the southeast coast, the Loess Plateau, and northeastern
rij|k= √̅̅̅̅̅̅ (11)
China (Fig. 2a, Fig. S3). There was also a notable rainfall deviation from
rij − rik
2
̅√̅̅̅̅̅̅2̅
1− r 1− r
northeast to southwest China, consistent with the patterns of drought
ik jk

where rij|k is the partial correlation coefficient between drought risk and hazard (Fig. S2). Drought hazard in northeastern China steadily
compositej (e.g., hazard) or climatej (e.g., temperature) by controlling increased from summer to autumn, while the Yungui Plateau experi­
compositek or climatek, rij is the correlation coefficient between drought enced moderate to high drought hazard in growing season. In contrast,
risk and compositej or climatej, rik is the correlation coefficient between the Tibetan Plateau and Hainan Province exhibited relatively low levels
drought risk and NDVI and compositek or climatek, and rjk is the corre­ of drought hazard throughout the study period.
lation coefficient between compositej (climatej) and compositek Drought hazard varied significantly across different climate zones
(climatek). (Fig. 2b and c). The continental zone suffered the highest drought haz­
The Conditional Permutation Importance (CPI) was employed to ard, followed by temperate, dry and tropical zones, while the polar zone
evaluate the primary risk component associated with drought risk. The suffered the mildest drought hazard (Fig. S3). Additionally, grass in the
CPI can be described as follows (Adeyeri et al., 2022). To calculate the continental zone faced the highest drought hazard (Fig. S3). Overall,
prediction error of regression trees in random forest, specifically the grass had the lowest drought hazard, while forests were most susceptible
to drought (Fig. S3). Throughout the growing season, drought hazard in
error R(t) and R(k) to tree t, based on predictors p and ntrees of trees using
(t)
the continental zone gradually weakened from May, reaching its lowest
the out of band sample β(t) , the following formula can be utilized before point in July, but then increased in subsequent months (Fig. S3). The
permuting the out-of-bag values Xk: drought hazard patterns in the dry and temperate zones showed similar
( ) variations.
∑ ŷi (t) − yi 2
R(t) = ⃒ (t) ⃒ (12)
⃒β ⃒
(t)
i∈β 3.1.2. Drought exposure
Overall, vegetation drought exposure showed similar patterns to the
( )2
(t) mean values of GPP throughout the growing season (Fig. 3). Southern
∑ ŷi (k) − yi
R(t) = ⃒ (t) ⃒ (13) China had high drought exposure for vegetation during the growing
(k) ⃒β ⃒
i∈β(t) season (Fig. 3, Fig. S4). In northern China, plants were more prone to
drought exposure in summer (Fig. S4). Additionally, vegetation in the
( )
ŷi (t) (t)
(14) Tibetan and Mongolian Plateaus experienced low levels of drought
(k) = f xi(k)
exposure.
(
xi(k) = xi1 , ⋯, xik− 1 , xp(i)k , xik+1 , ⋯, xip
)
(15) Vegetation in temperate and tropical zones was most vulnerable to
drought, while alpine vegetation exhibited the lowest exposure
⃒ ⃒
where ⃒β(t) ⃒ refers to the out-of-bag cardinal number sample for tree t; (Fig. 3b). The temporal variations of drought exposure during the
growing season were similar across climate zones, with peak values in
ŷi represents the function f (t) (xi ), which denotes the random forest
(t)
summer and decreasing values in autumn (Fig. S4). Forest, particularly
prediction of the out-of-bag observation i before permutation; xp(i)k is in the continental zone, had the highest drought exposure among
the ith Xk observation after permutation. vegetation types, while grassland, especially alpine grassland, had the
For the classification trees, the prediction error R(t) and R(k) can be lowest exposure (Fig. 3c, Fig. S4).
(t)

calculated as follows:
( )2 3.1.3. Drought vulnerability
∑ ŷi (t) ∕= yi Higher adaptability values indicate lower adaptability of vegetation
R(t) = ⃒ (t) ⃒ (16)
i∈β(t)
⃒β ⃒ to drought (Li et al., 2020). Vegetation in northern China and the Ti­
betan Plateau exhibited poorer adaptability compared to the southern
(
(t)
)2 region (Fig. 4a). Grass exhibited the lowest adaptability, while forest
∑ ŷi (k) ∕ = yi showed the highest adaptability to drought. Temporally, vegetation
R(t) = ⃒ (t) ⃒ (17)
(k)
i∈β(t)
⃒β ⃒ generally exhibited weaker adaptability to drought in summer
compared to spring and autumn (Fig. S5). Among the climate zones,
Then, the raw variable importance score PI(k) for each variable is vegetation in the polar zone had the least adaptability to drought, fol­
computed by taking the mean importance across all trees: lowed by the dry, continental, temperate, and tropical zones.
∑ntree (t) Vegetation in southern China exhibited the lowest sensitivity to
t=1 PI (k) drought (Fig. 4d). Forest demonstrated the lowest sensitivity to drought,
PI (k) = (18)
ntree while cropland and grass exhibited similar sensitivity throughout the
growing season (Fig. S6). Regarding climate zones, vegetation in the dry
where PI(k) = R(k) = R(t) .
(t) (t)
zone was most sensitive to drought (Fig. 4e and f), while alpine vege­
Additionally, we employed Partial Least Squares regression method tation exhibited the highest resistance to drought. Temporally, vegeta­
(PLSR) to analyze the influence of climatic factors on drought risk with tion in the dry and continental zones exhibited increased sensitivity to
two primary outputs, the variable importance projection (VIP) and the drought during summer compared to other seasons (Fig. S6).
standardized model coefficients (MC). VIP quantifies the importance of Drought vulnerability of vegetation was jointly determined by
each climatic factor in explaining the variation of drought risk, whose sensitivity and adaptability (Fig. 4, Fig. S7). Vegetation displayed higher
threshold was often set to one (Guo et al., 2019). Meanwhile, MC in­ susceptibility to drought in spring and early summer compared to other
dicates the effects of climatic factors on drought risk. seasons. Besides, vegetation in northern regions exhibited greater
vulnerability to drought during the growing season. Consistent with

4
W. Ge et al. Journal of Hydrology 623 (2023) 129877

Fig. 2. Spatial variations of drought hazard caused by meteorological droughts over the growing season across China from 2001 to 2019. (a) The spatial variation of
drought hazard; (b) and (c) violin plots of impacts of drought hazard on vegetation on climate zones and climate-related vegetation, respectively.

Fig. 3. Spatial variation of GPP and drought exposure over the growing season in China (2001–2019): (a) mean GPP values; (b) and (c) violin plots of mean GPP
values based on climate zones and climate-related vegetation, respectively; (d) spatial variations of drought exposure; (e) and (f) violin plots of drought exposure
impacts on climate zones and climate-related vegetation, respectively.

5
W. Ge et al. Journal of Hydrology 623 (2023) 129877

Fig. 4. Drought vulnerability assessment over the growing season across China from 2001 to 2019. (a) (d) and (g) are the spatial variations of drought adaptability,
sensitivity and vulnerability, respectively; (b), (c), (e), (f), (h) and (i) are violin plots of impacts of drought adaptability, sensitivity and vulnerability on climate zones
and climate-related vegetation, respectively.

6
W. Ge et al. Journal of Hydrology 623 (2023) 129877

drought adaptability and sensitivity, grass demonstrated the highest 3.3.2. Impacts of climate variability on drought risk
drought vulnerability, while forest showed the lowest (Fig. S7). Addi­ Approximately 39.56% of the study area exhibited temperature-
tionally, drought vulnerability varied across different climate zones, dominated drought risk, primarily in the Loess Plateau, north, and
with the dry zone exhibiting the highest vulnerability, followed by the southwest China. Additionally, around 33.18% and 27.26% of vegeta­
continental, polar, temperate, and tropical zones (Fig. 4h, Fig. S7). tion pixels were dominated by precipitation and solar radiation,
respectively. The strongest relationships between precipitation risk and
3.2. Drought risk assessment solar radiation risk were predominantly observed in northeast China
(Fig. 8).
Regions with moderate to high vegetation drought risk (DS > 0.09, The critical impacts of climatic factors on drought risk were illus­
the average drought risk value of China) were primarily located from trated by the variable importance of PLSR projection (VIP) (Fig. S9).
northeastern China to southwestern China (48%; Fig. 5, Fig. S8). Increased temperature significantly aggravated drought risk in the
Conversely, vegetation on the Tibetan Plateau faced relatively low North China Plain and southeast coast (18.4%), while significant
drought risk. Vegetation in temperate zone suffered highest drought inhibiting effects were found in northeast China (25%) (Fig. S9a).
risk, followed by continental, dry, tropical and polar zones (Fig. S8d). Increased precipitation generally promoted drought risk in north and
Furthermore, cropland showed the highest drought risk, while grassland central China (23.9%), with significant inhibiting effects observed in
exhibited the lowest. Temporally, drought risk increased from spring to southwest China and the Tibetan Plateau (Fig. S9b). Solar radiation had
summer and then gradually declined in the continental and dry zones. In promoting effects on drought risk mainly in northern China (21.3%),
contrast, drought risk in the tropical and temperate zones exhibited while scattered inhibiting effects were observed nationwide (26.1%)
insignificant variations throughout the growing season. (Fig. S9c).

3.3. Driving forces analysis 4. Discussions

3.3.1. Impacts of drought risk components on drought risk 4.1. Impacts of drought risk in different climate-related vegetation zones
Autocorrelation and multicollinearity of drought risk and its com­
ponents were assessed using correlation coefficient heatmap and var­ Drought is a destructive and widespread natural disaster that occurs
iogram inflation factor (VIF) in Fig. 6. Despite the strong correlation in nearly every corner of the world, even in humid areas and tropical
between exposure and adaptability, the VIF values for all four risk rainforests. Extensive work has been carried out to assess drought risk in
components were below 5, indicating the absence of significant multi­ the current climate change context, particularly focusing on its impact
collinearity. Thus, these components are suitable for explaining drought on humans and agriculture (Dai et al., 2020; Fang et al., 2019; Hoque
risk. et al., 2021; Zhao et al., 2020). However, limited attention has been
The maximum partial correlation coefficients (Fig. 7a-c) were uti­ given to evaluating drought risk specifically for different vegetation
lized to identify the primary risk components of drought risk. Addi­ types. Hence, quantitative assessment of drought risk on vegetation in
tionally, CPI analysis (Fig. 7d) was employed to determine the dominant this study is of great significance for vegetation management.
risk component. Drought hazard played a dominant role in drought risk Previous studies have quantified agricultural drought exposure by
in central China, accounting for 55.15% of the study area. Exposure considering factors such as population density or irrigation conditions
dominated drought risk in 24.46% of the area, mainly in northeast China (Hoque et al., 2021; Meza et al., 2020). However, in studies focused on
and the Tibetan Plateau. Drought vulnerability determined drought risk vegetation, it is essential to define drought exposure as the inventory of
in southwest China, comprising 20.39% of the area. The CPI results in elements (i.e., vegetation) under drought conditions, following the IPCC
Fig. 7d confirmed the dominant role of drought hazard in vegetation definition (Field et al., 2012; Zhao et al., 2020). Human disturbances
drought risk, consistent with the partial correlation coefficients. like irrigation have been quantified by vegetation drought sensitivity in
this study (equations (4)–(5)) (Zhang et al., 2017a). Consequently, to
avoid mistakenly conflating exposure and vulnerability, we quantify

Fig. 5. Assessment of drought risk over the growing season across China from 2001 to 2019. (a) is the spatial variations of drought risk, (b) and (c) are violin plots of
impacts of drought risk on climate zones and climate-related vegetation, respectively. The dashed horizontal lines in plots b and c are the average drought risk values
of China (i.e., 0.08).

7
W. Ge et al. Journal of Hydrology 623 (2023) 129877

Fig. 6. The heatmap of (a) correlation coefficients and (b) the VIF test among the four components of drought risk.

Fig. 7. (a) Spatial variation of the dominant drought risk components based on maximum partial correlation coefficients, (b) area fractions dominated by hazard,
exposure and vulnerability, respectively, (c) patterns of pixels dominated by risk components in each climate-related vegetation type, and (d) conditional permu­
tation importance (CPI) of the drought risk components.

Fig. 8. (a) Spatial variation of the dominant climate factors on drought risk based on maximum partial correlation coefficients, (b) area fractions dominated by
climate factors, and (c) patterns of pixels dominated by climatic factors in each climate-related vegetation types.

8
W. Ge et al. Journal of Hydrology 623 (2023) 129877

drought exposure of vegetation using the density of vegetation and its Furthermore, adverse weather events accompanying extreme precipi­
health condition instead of irrigation status (Field et al., 2012). It is tation, such as hails and strong winds, pose additional threats to vege­
important to note that exposure alone is insufficient for determining tation growth (Schlie et al., 2019; Botzen et al., 2010). Consequently,
risk, as vegetation can be exposed to drought without necessarily being increased extreme precipitation events intensifies vegetation vulnera­
vulnerable to it (Field et al., 2012). For example, forest was more bility and amplifies vegetation drought risk.
exposed to drought than grass and cropland, but least vulnerable to it. The impacts of temperature on vegetation drought risk also exhibited
The drought adaptability refers to the loss of vegetation productivity a dual nature. On one hand, increased temperature can have positive
during drought events, while drought sensitivity measures the response effects on vegetation growth, potentially enhancing resistance to climate
speed of vegetation to drought. Forests, characterized by their strong change and reducing vulnerability. However, higher temperatures also
resistance to droughts, generally exhibited the lowest adaptability and lead to increased evapotranspiration, exacerbating drought conditions.
sensitivity to drought (Zhang et al., 2017a), resulting in lower vulner­ As a result, warming has both promoting and inhibiting effects on
ability. However, forest still faced higher drought risk than grass due to vegetation drought risk in China.
its higher drought hazard and exposure (especially in northeastern
China). 4.3. Reliability, limitation and uncertainty of the assessment
The shallow roots of herbaceous plants limit them to access water
from deep soil layers, resulting in high sensitivity to water deficit (Zhang We evaluated the vegetation-related drought risk in China by
et al., 2017a). Although irrigation systems in pastures can reduce the considering drought hazard, vulnerability, and exposure (Edenhofer,
vulnerability of grass to drought (Yu et al., 2011), the area covered by 2015). Unlike previous studies (Li et al., 2020), we used the density of
irrigation for grass is much smaller compared to natural grasslands, vegetation and its health condition instead of vegetation types to
resulting in limited effectiveness in mitigating drought effects. Never­ represent drought exposure, aiming to improve the accuracy and
theless, alpine vegetation exhibited relative low sensitivity to drought, repeatability of the assessment. The spatial variation of vegetation
even though they are also dominated by meadow and grass, due to the drought vulnerability observed in our study is consistent with the
joint influences of precipitation and the melting of snowpack and frozen findings of previous quantitative studies (Fang et al., 2019; Gao et al.,
soil on water availability (Zhang et al., 2017a). On the other hand, 2018; Li et al., 2020), indicating a certain level of reliability. Further­
alpine ecosystems have been identified as vulnerable and sensitive areas more, our study, with higher spatial resolution datasets, provides more
to climate change (Crossman et al., 2012), leading to considerable losses detailed insights into the impacts of drought risk on climate-related
in droughts (i.e., the highest drought adaptability) for alpine vegetation vegetation, allowing for comparisons and discussions.
especially grass. However, uncertainties remain. The coarse spatial resolution of
Cropland, similar to grass, exhibits short-term seasonal periodicity drought index limits the accuracy of risk evaluation. Many other factors
and shallow root systems. However, its vulnerability to drought is unaccounted in this study, such as bushfires, soil salinization or insect
influenced not only by the nature of the vegetation itself but also by pests, can also contribute to poor vegetation growth and increased
human interventions such as irrigation (Ge et al., 2021). Consequently, drought risk (Cao et al., 2022). Additionally, the assessment of drought
cropland has lower drought vulnerability compared to grass but higher risk to vegetation was limited to the growing season, neglecting the
vulnerability compared to forest. In northern China (the Northeast Plain, winter period when vegetation is absent in northern China, potentially
North China Plain, and Loess Plateau), cropland is exposed to significant underestimating winter drought-induced risk on vegetation. Further­
drought risk due to high drought hazard, posing threats to food security more, quantifying the contrasting effects of climatic factors on drought
in the country (Li et al., 2014). Besides, the renowned groundwater risk and addressing the uncertainties associated with their complex in­
funnel in the North China Plain exacerbates water shortages, posing teractions requires further investigation. The effects of groundwater and
constraints on the sustainable development of food production and irrigation on vegetation growth, particularly in crops, also warrant
ecological well-being (Zhou et al., 2023). To address these challenges, further exploration. Therefore, further research is needed to quantify the
appropriate agricultural practices such as efficient irrigation, crop comprehensive driving factors influencing drought risk and assess the
rotation, and fertilization should be implemented in the North China risk of winter drought on vegetation in southern China.
Plain. Similarly, the Loess Plateau also faces a critical challenge of
increasing crop production with limited water resources due to water 5. Conclusion
scarcity and population growth (Zhang et al., 2014). Therefore, irriga­
tion management supported by check-dams/reservoirs and the alternate We developed a framework to assess vegetation drought risk and
crop system should be considered on the plateau to ensure food security. examined its spatial patterns in different climate-related vegetation
zones in China using GPP and multiscalar SPEI. Our findings revealed
4.2. Analysis of driving factors that drought risk in vegetation was influenced by drought hazard
(55.15%), exposure (24.46%), and vulnerability (20.39%). Approxi­
Temperature and precipitation played crucial roles in shaping the mately 48% of the study area exhibited moderate to high drought risk,
drought risk patterns in regions with moderate to high drought risk. primarily in dry, continental, and some temperate zones. Forest had the
While insufficient rainfall is a direct cause of meteorological drought, it highest drought risk, followed by cropland and grass. The pronounced
was observed that increased precipitation significantly contributed to drought risk in cropland in northern China poses a threat to food secu­
drought risk in the dry and continental zones. This phenomenon can be rity, emphasizing the need for appropriate agricultural practices such as
attributed to the occurrence of extreme precipitation events character­ irrigation, crop rotation, and fertilization to ensure sustainable crop
ized by long intervals. These infrequent yet intense rainfall events production.
exacerbate within-season drought, leading to heightened drought haz­ While increased temperature promotes vegetation growth across
ards and thus increased vegetation drought risk (Knapp et al., 2008). China, it also contributes to heightened drought risk by exacerbating
Meanwhile, the increase in extreme precipitation events indicates water depletion, particularly in northern regions. Additionally, the
declining climate stability, which weakens vegetation resistance to increased rainfall observed in certain parts of northern China could be
climate change and further intensifies its vulnerability to drought (Li attributed to extreme weather events, intensifying vegetation drought
et al., 2020). Additionally, excessive waterlogging induced by deluges vulnerability and exacerbating drought risk. Our study quantified
poses risks to vegetation growth by damaging root, reducing soil oxygen vegetation drought risk across various climate zones and shed light on
availability, and delaying crop planting/harvesting (Li et al., 2019; the influence of climatic factors on drought risk. These findings offer
Jabloun et al., 2015; Urban et al., 2015; Parent et al., 2008). valuable scientific insights for vegetation management in the face of

9
W. Ge et al. Journal of Hydrology 623 (2023) 129877

climate change. Dai, M., Huang, S., Huang, Q., Leng, G., Guo, Y., Wang, L., Fang, W., Li, P., Zheng, X.,
2020. Assessing agricultural drought risk and its dynamic evolution characteristics.
Agric. Water Manag. 231, 106003.
CRediT authorship contribution statement Dimiceli, C., Carroll, M., Sohlberg, R., Kim, D., Kelly, M., & Townshend, J. (2015).
MOD44B MODIS/terra vegetation continuous fields yearly L3 global 250m SIN grid
V006. NASA EOSDIS Land Processes Distributed Active Archive Center, 10.
Wenyan Ge: Conceptualization, Methodology, Software, Formal Duo, A., Zhao, W., Qu, X., Jing, R., Xiong, K., 2016. Spatio-temporal variation of
analysis, Writing – original draft, Writing – review & editing. Xiuxia Li: vegetation coverage and its response to climate change in North China plain in the
Formal analysis, Methodology, Software, Validation. Mengxia Xie: last 33 years. Int. J. Appl. Earth Obs. Geoinf. 53, 103–117.
Edenhofer, O., 2015. Climate Change 2014: Mitigation of Climate Change. Cambridge
Writing – original draft, Methodology, Software, Validation. Bowen Yu: University Press.
Software, Validation, Investigation. Juying Jiao: Writing – review & Fang, W., Huang, S., Huang, Q., Huang, G., Wang, H., Leng, G., Wang, L., Guo, Y., 2019.
editing, Investigation. Jianqiao Han: Investigation, Validation. Fei Probabilistic assessment of remote sensing-based terrestrial vegetation vulnerability
to drought stress of the Loess Plateau in China. Remote Sens. Environ. 232, 111290.
Wang: Investigation, Supervision, Writing – review & editing.
Feng, X., Fu, B., Piao, S., Wang, S., Ciais, P., Zeng, Z., Lü, Y., Zeng, Y., Li, Y., Jiang, X.,
Wu, B., 2016. Revegetation in China’s Loess Plateau is approaching sustainable
water resource limits. Nat. Clim. Chang. 6 (11), 1019–1022.
Declaration of Competing Interest Field, C.B., Barros, V., Stocker, T.F., Dahe, Q., 2012. Managing the Risks of Extreme
Events and Disasters to Advance Climate Change Adaptation: Special Report of the
Intergovernmental Panel on Climate Change. Cambridge University Press.
The authors declare that they have no known competing financial
Frank, D., Reichstein, M., Bahn, M., Thonicke, K., Frank, D., Mahecha, M.D., Smith, P.,
interests or personal relationships that could have appeared to influence Velde, M., Vicca, S., Babst, F., Beer, C., Buchmann, N., Canadell, J.G., Ciais, P.,
the work reported in this paper. Cramer, W., Ibrom, A., Miglietta, F., Poulter, B., Rammig, A., Seneviratne, S.I.,
Walz, A., Wattenbach, M., Zavala, M.A., Zscheischler, J., 2015. Effects of climate
extremes on the terrestrial carbon cycle: concepts, processes and potential future
Data availability impacts. Glob. Chang. Biol. 21 (8), 2861–2880.
Gao, J., Jiao, K., Wu, S., 2018. Quantitative assessment of ecosystem vulnerability to
The authors do not have permission to share data. climate change: methodology and application in China. Environ. Res. Lett. 13 (9),
094016.
Ge, W., Deng, L., Wang, F., Han, J., 2021. Quantifying the contributions of human
activities and climate change to vegetation net primary productivity dynamics in
Acknowledgement China from 2001 to 2016. Sci. Total Environ. 773, 145648.
Ge, J., Hou, M., Liang, T., Feng, Q., Meng, X., Liu, J., Bao, X., Gao, H., 2022.
Spatiotemporal dynamics of grassland aboveground biomass and its driving factors
The authors are grateful to National Aeronautics and Space Admin­
in North China over the past 20 years. Sci. Total Environ. 826, 154226.
istration (NASA), National Earth System Science Data Center (http:// Gosling, S.N., Arnell, N.W., 2016. A global assessment of the impact of climate change on
loess.geodata.cn/# (accessed 3 March 2023)), and the Climatology Lab water scarcity. Clim. Change 134 (3), 371–385.
(http://www.climatologylab.org/ (accessed 3 March 2023)) for Guo, L., Wang, J., Li, M., Liu, L., Xu, J., Cheng, J., Gang, C., Yu, Q., Chen, J., Peng, C.,
Luedeling, E., 2019. Distribution margins as natural laboratories to infer species’
providing dataset. flowering responses to climate warming and implications for frost risk. Agric. For.
Meteorol. 268, 299–307.
Hanewinkel, M., Cullmann, D.A., Schelhaas, M.-J., Nabuurs, G.-J., Zimmermann, N.E.,
Funding 2013. Climate change may cause severe loss in the economic value of European
forest land. Nat. Clim. Chang. 3 (3), 203–207.
Hansen, M.C., DeFries, R.S., Townshend, J.R.G., Carroll, M., Dimiceli, C., Sohlberg, R.A.,
This research was jointly funded by the National Natural Science 2003. Global percent tree cover at a spatial resolution of 500 meters: first results of
Foundation of China (42007063, U2243213), China Postdoctoral Sci­ the MODIS vegetation continuous fields algorithm. Earth Interact. 7 (10), 1–15.
ence Foundation (2020M673499), and Young Scholars of the West He, C., Liu, Z., Wu, J., Pan, X., Fang, Z., Li, J., Bryan, B.A., 2021. Future global urban
water scarcity and potential solutions. Nat. Commun. 12, 4667.
(XAB2019B10). Hoque, M.-A., Pradhan, B., Ahmed, N., Sohel, M.S.I., 2021. Agricultural drought risk
assessment of Northern New South Wales, Australia using geospatial techniques. Sci.
Total Environ. 756, 143600.
Appendix A. Supplementary data
Islam, A.R.M.T., Shen, S., Hu, Z., Rahman, M.A., 2017. Drought hazard evaluation in
boro paddy cultivated areas of western Bangladesh at current and future climate
Supplementary data to this article can be found online at https://doi. change conditions. Adv. Meteorol. 2017, 1–12.
Jabloun, M., Schelde, K., Tao, F., Olesen, J.E., 2015. Effect of temperature and
org/10.1016/j.jhydrol.2023.129877.
precipitation on nitrate leaching from organic cereal cropping systems in Denmark.
Eur. J. Agron. 62, 55–64.
References Jiang, W., Wang, L., Zhang, M., Yao, R., Chen, X., Gui, X., Sun, J., Cao, Q., 2021. Analysis
of drought events and their impacts on vegetation productivity based on the
integrated surface drought index in the Hanjiang River Basin, China. Atmos. Res. 254,
Abatzoglou, J.T., Dobrowski, S.Z., Parks, S.A., Hegewisch, K.C., 2018. TerraClimate, a
105536.
high-resolution global dataset of monthly climate and climatic water balance from
Knapp, A.K., Beier, C., Briske, D.D., Classen, A.T., Luo, Y., Reichstein, M., Smith, M.D.,
1958–2015. Sci. Data 5, 1–12.
Smith, S.D., Bell, J.E., Fay, P.A., Heisler, J.L., Leavitt, S.W., Sherry, R., Smith, B.,
Abhishek, Kinouchi, T., 2022. Multidecadal land water and groundwater drought
Weng, E., 2008. Consequences of more extreme precipitation regimes for terrestrial
evaluation in Peninsular India. Remote Sensing 14 (6), 1486.
ecosystems. Bioscience 58 (9), 811–821.
Abhishek, Kinouchi, T., & Sayama, T. (2021). A comprehensive assessment of water
Lavell, A., Oppenheimer, M., Diop, C., Hess, J., Lempert, R., Li, J., Muir-Wood, R.,
storage dynamics and hydroclimatic extremes in the Chao Phraya River Basin during
Myeong, S., Moser, S., Takeuchi, K., Cardona, O.-D., Hallegatte, S., Lemos, M.,
2002–2020. Journal of Hydrology, 603, 126868.
Little, C., Lotsch, A., Weber, E., 2012. In: Managing the Risks of Extreme Events and
Adeyeri, O.E., Zhou, W., Wang, X., Zhang, R., Laux, P., Ishola, K.A., Usman, M., 2022.
Disasters to Advance Climate Change Adaptation: Special Report of the
The trend and spatial spread of multisectoral climate extremes in CMIP6 models. Sci.
Intergovernmental Panel on Climate Change. Cambridge University Press,
Rep. 12 (1), 21000.
pp. 25–64.
Beck, H.E., Zimmermann, N.E., McVicar, T.R., Vergopolan, N., Berg, A., Wood, E.F.,
Li, Y., Zhang, W., Ma, L., Wu, L., Shen, J., Davies, W.J., Oenema, O., Zhang, F., Dou, Z.,
2018. Present and future Köppen-Geiger climate classification maps at 1-km
2014. An analysis of China’s grain production: looking back and looking forward.
resolution. Sci. Data 5, 1–12.
Food Energy Secur. 3, 19–32.
Botzen, W.J.W., Bouwer, L.M., Van den Bergh, J.C.J.M., 2010. Climate change and
Li, Y., Guan, K., Schnitkey, G.D., DeLucia, E., Peng, B., 2019. Excessive rainfall leads to
hailstorm damage: empirical evidence and implications for agriculture and
maize yield loss of a comparable magnitude to extreme drought in the United States.
insurance. Resour. Energy Econ. 32 (3), 341–362.
Glob. Chang. Biol. 25 (7), 2325–2337.
Cao, S., Zhang, L., He, Y., Zhang, Y., Chen, Y., Yao, S., Yang, W., Sun, Q., 2022. Effects
Li, Y., Liu, W., Feng, Q., Zhu, M., Yang, L., Zhang, J., Yin, X., 2023. The role of land use
and contributions of meteorological drought on agricultural drought under different
change in affecting ecosystem services and the ecological security pattern of the Hexi
climatic zones and vegetation types in Northwest China. Sci. Total Environ. 821,
Regions, Northwest China. Sci. Total Environ. 855, 158940.
153270.
Li, K., Tong, Z., Liu, X., Zhang, J., Tong, S., 2020. Quantitative assessment and driving
Carrão, H., Naumann, G., Barbosa, P., 2016. Mapping global patterns of drought risk: an
force analysis of vegetation drought risk to climate change: Methodology and
empirical framework based on sub-national estimates of hazard, exposure and
application in Northeast China. Agric. For. Meteorol. 282-283, 107865.
vulnerability. Glob. Environ. Chang. 39, 108–124.
Crossman, N.D., Bryan, B.A., Summers, D.M., 2012. Identifying priority areas for
reducing species vulnerability to climate change. Divers. Distrib. 18, 60–72.

10
W. Ge et al. Journal of Hydrology 623 (2023) 129877

Liu, J., Liu, M., Deng, X., Zhuang, D., Zhang, Z., Luo, D., 2002. The land use and land Wu, X., Liu, H., Li, X., Ciais, P., Babst, F., Guo, W., Zhang, C., Magliulo, V., Pavelka, M.,
cover change database and its relative studies in China. J. Geog. Sci. 12 (3), Liu, S., Huang, Y., Wang, P., Shi, C., Ma, Y., 2018. Differentiating drought legacy
275–282. effects on vegetation growth over the temperate Northern Hemisphere. Glob. Chang.
Meza, I., Siebert, S., Döll, P., Kusche, J., Herbert, C., Eyshi Rezaei, E., Nouri, H., Biol. 24 (1), 504–516.
Gerdener, H., Popat, E., Frischen, J., Naumann, G., Vogt, J.V., Walz, Y., Sebesvari, Z., Xiao, J., Chevallier, F., Gomez, C., Guanter, L., Hicke, J.A., Huete, A.R., Ichii, K., Ni, W.,
Hagenlocher, M., 2020. Global-scale drought risk assessment for agricultural Pang, Y., Rahman, A.F., Sun, G., Yuan, W., Zhang, L.i., Zhang, X., 2019. Remote
systems. Nat. Hazards Earth Syst. Sci. 20 (2), 695–712. sensing of the terrestrial carbon cycle: a review of advances over 50 years. Remote
Padfield, D., Lowe, C., Buckling, A., Ffrench-Constant, R., Jennings, S., Shelley, F., Sens. Environ. 233, 111383.
Ólafsson, J.S., Yvon-Durocher, G., Jeyasingh, P., 2017. Metabolic compensation Xiong, J., Guo, S., Kinouchi, T., 2022. Leveraging machine learning methods to quantify
constrains the temperature dependence of gross primary production. Ecol. Lett. 20 50 years of dwindling groundwater in India. Sci. Total Environ. 835, 155474.
(10), 1250–1260. Xu, C., McDowell, N.G., Fisher, R.A., Wei, L., Sevanto, S., Christoffersen, B.O., Weng, E.,
Parent, C., Capelli, N., Berger, A., Crèvecoeur, M., Dat, J.F., 2008. An overview of plant Middleton, R.S., 2019. Increasing impacts of extreme droughts on vegetation
responses to soil waterlogging. Plant Stress 2 (1), 20–27. productivity under climate change. Nat. Clim. Chang. 9 (12), 948–953.
Peng, S., Ding, Y., Liu, W., Li, Z., 2019. 1 km monthly temperature and precipitation Yu, Z., Ciais, P., Piao, S., Houghton, R.A., Lu, C., Tian, H., Agathokleous, E., Kattel, G.R.,
dataset for China from 1901 to 2017. Earth Syst. Sci. Data 11, 1931–1946. Sitch, S., Goll, D., Yue, X., Walker, A., Friedlingstein, P., Jain, A.K., Liu, S., Zhou, G.,
Piao, S., Ciais, P., Huang, Y., Shen, Z., Peng, S., Li, J., Zhou, L., Liu, H., Ma, Y., Ding, Y., 2022. Forest expansion dominates China’s land carbon sink since 1980. Nat.
Friedlingstein, P., Liu, C., Tan, K., Yu, Y., Zhang, T., Fang, J., 2010. The impacts of Commun. 13 (1), 5374.
climate change on water resources and agriculture in China. Nature 467 (7311), Yu, Y., Liu, J., Wang, H., Liu, M., 2011. Assess the potential of solar irrigation systems for
43–51. sustaining pasture lands in arid regions–A case study in Northwestern China. Appl.
Running, S.W., Zhao, M. (2015). Daily GPP and annual NPP (MOD17A2/A3) products Energy 88 (9), 3176–3182.
NASA Earth Observing System MODIS land algorithm. MOD17 User’s Guide, 2015, 1- Zhang, L., Chen, F., Lei, Y., 2020. Climate change and shifts in cropping systems together
28. exacerbate China’s water scarcity. Environ. Res. Lett. 15 (10), 104060.
Schlie, E.E.J., Wuebbles, D., Stevens, S., Trapp, R., Jewett, B., 2019. A radar-based study Zhang, Q., Kong, D., Singh, V.P., Shi, P., 2017a. Response of vegetation to different time-
of severe hail outbreaks over the contiguous United States for 2000–2011. Int. J. scales drought across China: spatiotemporal patterns, causes and implications.
Climatol. 39 (1), 278–291. Global Planet. Change 152, 1–11.
Song, L., Li, Y., Ren, Y., Wu, X., Guo, B., Tang, X., Shi, W., Ma, M., Han, X., Zhao, L., Zhang, S., Sadras, V., Chen, X., Zhang, F., 2014. Water use efficiency of dryland maize in
2019. Divergent vegetation responses to extreme spring and summer droughts in the Loess Plateau of China in response to crop management. Field Crop Res 163,
Southwestern China. Agric. For. Meteorol. 279, 107703. 55–63.
Urban, D.W., Roberts, M.J., Schlenker, W., Lobell, D.B., 2015. The effects of extremely Zhang, Q., Zhang, J., Wang, C., 2017b. Risk assessment of drought disaster in typical area
wet planting conditions on maize and soybean yields. Clim. Change 130 (2), of corn cultivation in China. Theor. Appl. Climatol. 128 (3-4), 533–540.
247–260. Zhang, T., Zhou, J., Yu, P., Li, J., Kang, Y., Zhang, B., 2023. Response of ecosystem gross
van Dijk, A.I.J.M., Beck, H.E., Crosbie, R.S., de Jeu, R.A.M., Liu, Y.Y., Podger, G.M., primary productivity to drought in northern China based on multi-source remote
Timbal, B., Viney, N.R., 2013. The Millennium Drought in southeast Australia sensing data. J. Hydrol. 616, 128808.
(2001–2009): Natural and human causes and implications for water resources, Zhao, A., Zhang, A., Cao, S., Liu, X., Liu, J., Cheng, D., 2018. Responses of vegetation
ecosystems, economy, and society. Water Resour. Res. 49 (2), 1040–1057. productivity to multi-scale drought in Loess Plateau, China. Catena 163, 165–171.
Vicente-Serrano S.M., López-Moreno J.I., Beguería S., ’A multi-scalardrought index Zhao, J., Zhang, Q., Zhu, X., Shen, Z., Yu, H., 2020. Drought risk assessment in China:
sensitive to global warming: The standardized precipitation evapotranspiration evaluation framework and influencing factors. Geogr. Sustain. 1 (3), 220–228.
index - SPEI’ (in prep.). Zheng, X., Huang, S., Peng, J., Leng, G., Huang, Q., Fang, W., Guo, Y., 2022. Flash
Wan, J.-Z., Wang, C.-J., Qu, H., Liu, R., Zhang, Z.-X., 2018. Vulnerability of forest droughts identification based on an improved framework and their contrasting
vegetation to anthropogenic climate change in China. Sci. Total Environ. 621, impacts on vegetation over the Loess Plateau China. Water Resour. Res. 58
1633–1641. e2021WR031464.
Wang, X., Xiao, X., Zou, Z., Dong, J., Qin, Y., Doughty, R.B., Menarguez, M.A., Chen, B., Zhou, Y., Fu, D., Lu, C., Xu, X., Tang, Q., 2021. Positive effects of ecological restoration
Wang, J., Ye, H., Ma, J., Zhong, Q., Zhao, B., Li, B., 2020. Gainers and losers of policies on the vegetation dynamics in a typical ecologically vulnerable area of
surface and terrestrial water resources in China during 1989–2016. Nat. Commun. China. Ecol. Eng. 159, 106087.
11 (1), 3471. Zhou, Y., Dong, J., Cui, Y., Zhao, M., Wang, X., Tang, Q., Zhang, Y., Zhou, S.,
Wei, X., He, W., Zhou, Y., Cheng, N., Xiao, J., Bi, W., Liu, Y., Sun, S., Ju, W., 2023. Metternicht, G., Zou, Z., Zhang, G., Xiao, X., 2023. Ecological restoration exacerbates
Increased sensitivity of global vegetation productivity to drought over the recent the agriculture-induced water crisis in North China Region. Agric. For. Meteorol.
three decades. J. Geophys. Res. Atmos. e2022JD037504. 331, 109341.

11

You might also like