Tracing Time-Varying Characteristics of Meteorological Drought through Nonstationary Joint Deficit Index

R. Vinnarasi aDepartment of Civil Engineering, Indian Institute of Technology Delhi, New Delhi, India
bDepartment of Civil Engineering, Indian Institute of Technology Roorkee, Roorkee, Uttarakhand, India

Search for other papers by R. Vinnarasi in
Current site
Google Scholar
PubMed
Close
https://orcid.org/0000-0001-9834-6779
,
C. T. Dhanya aDepartment of Civil Engineering, Indian Institute of Technology Delhi, New Delhi, India

Search for other papers by C. T. Dhanya in
Current site
Google Scholar
PubMed
Close
, and
Hemant Kumar aDepartment of Civil Engineering, Indian Institute of Technology Delhi, New Delhi, India

Search for other papers by Hemant Kumar in
Current site
Google Scholar
PubMed
Close
Free access

Abstract

Standardized precipitation index (SPI) is one of the frequently used meteorological drought indices. However, the time-varying characteristics observed in the historical precipitation data questions the reliability of SPI and motivated the development of nonstationary SPI. To overcome some of the limitations in the existing nonstationary drought indices, a new framework for drought index is proposed, incorporating the temporal dynamics in the precipitation. The proposed drought index is developed by coupling the joint deficit index with the extended time sliding window–based nonstationary modeling (TSW-NSM). The proposed nonstationary joint deficit index (NJDI) detects the signature of nonstationarity in the distribution parameter and models both long-term (i.e., trend) and short-term (i.e., step-change) temporal dynamics of distribution parameters. The efficacy of NJDI is demonstrated by employing it to identify the meteorological drought-prone areas over India. The changes observed in the distribution parameter of rainfall series reveal an increasing number of dry days in recent decades all over India, except the northeast. Comparison of NJDI and stationary joint deficit index (JDI) reveals that JDI overestimates drought when frequent severe dry events are clustered and underestimates when these events are scattered, which indicates that the traditional index is biased toward the lowest magnitude of precipitation while classifying the drought. Moreover, NJDI could closely capture historical droughts and their spatial variations, thereby reflecting the temporal dynamics of rainfall series and the changes in the pattern of dry events over India. NJDI proves to be a potentially reliable index for drought monitoring in a nonstationary climate.

Significance Statement

Drought is one of the most severe natural disasters and is expected to intensify under a warming climate. There has been progress on developing newer methodologies to characterize drought severity under the changing climate. However, some limitations remain in capturing the temporal changes of the precipitation, especially the nonstationarity in variance (rapid increases in extreme precipitation versus the average precipitation). We propose an extension to the time sliding window approach to capture the nonstationarity in variance and mean while incorporating short-term (step-change) and long-term (trend) temporal dynamics. We apply the methodology to a subcontinent-sized heterogeneous country of India with gridded rainfall dataset (0.25° × 0.25°, 1901–2013) from the India Meteorological Department. We find that the majority of grids exhibit negative and positive trends in shape and scale parameters, respectively, which ultimately leads to an increase in the number of drier events, whereas a contradictory pattern is exhibited in the northwest Indian region (decrease in the number of dry days). The proposed index gives a potential framework for drought monitoring under the changing climate and can be extended to develop a multivariate nonstationary joint deficit index by incorporating other hydrological variables (e.g., soil moisture, diurnal temperature range, streamflow).

© 2023 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: R. Vinnarasi, vinnarasi@ce.iitr.ac.in

Abstract

Standardized precipitation index (SPI) is one of the frequently used meteorological drought indices. However, the time-varying characteristics observed in the historical precipitation data questions the reliability of SPI and motivated the development of nonstationary SPI. To overcome some of the limitations in the existing nonstationary drought indices, a new framework for drought index is proposed, incorporating the temporal dynamics in the precipitation. The proposed drought index is developed by coupling the joint deficit index with the extended time sliding window–based nonstationary modeling (TSW-NSM). The proposed nonstationary joint deficit index (NJDI) detects the signature of nonstationarity in the distribution parameter and models both long-term (i.e., trend) and short-term (i.e., step-change) temporal dynamics of distribution parameters. The efficacy of NJDI is demonstrated by employing it to identify the meteorological drought-prone areas over India. The changes observed in the distribution parameter of rainfall series reveal an increasing number of dry days in recent decades all over India, except the northeast. Comparison of NJDI and stationary joint deficit index (JDI) reveals that JDI overestimates drought when frequent severe dry events are clustered and underestimates when these events are scattered, which indicates that the traditional index is biased toward the lowest magnitude of precipitation while classifying the drought. Moreover, NJDI could closely capture historical droughts and their spatial variations, thereby reflecting the temporal dynamics of rainfall series and the changes in the pattern of dry events over India. NJDI proves to be a potentially reliable index for drought monitoring in a nonstationary climate.

Significance Statement

Drought is one of the most severe natural disasters and is expected to intensify under a warming climate. There has been progress on developing newer methodologies to characterize drought severity under the changing climate. However, some limitations remain in capturing the temporal changes of the precipitation, especially the nonstationarity in variance (rapid increases in extreme precipitation versus the average precipitation). We propose an extension to the time sliding window approach to capture the nonstationarity in variance and mean while incorporating short-term (step-change) and long-term (trend) temporal dynamics. We apply the methodology to a subcontinent-sized heterogeneous country of India with gridded rainfall dataset (0.25° × 0.25°, 1901–2013) from the India Meteorological Department. We find that the majority of grids exhibit negative and positive trends in shape and scale parameters, respectively, which ultimately leads to an increase in the number of drier events, whereas a contradictory pattern is exhibited in the northwest Indian region (decrease in the number of dry days). The proposed index gives a potential framework for drought monitoring under the changing climate and can be extended to develop a multivariate nonstationary joint deficit index by incorporating other hydrological variables (e.g., soil moisture, diurnal temperature range, streamflow).

© 2023 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: R. Vinnarasi, vinnarasi@ce.iitr.ac.in

1. Introduction

Drought is one of the most severe natural disasters, whose adverse effects span the entire ecosystem, agriculture, and economy. It is also called a creeping disaster as its immediate impact may not be visible like other climate extremes, although it culminates in colossal damage. The impending threat of intensification of natural disasters because of the warming climate, the rapid growth of population, and expansion of agriculture, energy, and industrial sectors, further aggravate the distress caused by drought (Mishra and Singh 2010; Wang et al. 2015). Furthermore, drought is one of the least understood natural disasters due to the complex mechanisms involved in its occurrence (Kao and Govindaraju 2010). Hitherto, there is no precise scientific definition for drought. Hence, so often, it is characterized as a moisture deficit observed in any component of the hydrological cycle (Farahmand and Aghakouchak 2015; Kao and Govindaraju 2010), and accordingly, it is classified into four categories: meteorological (precipitation deficit), agricultural (root-zone soil moisture deficit), hydrological (streamflow deficit), and socioeconomic (deficit of supply of socioeconomic goods with respect to demand caused by a weather-related deficit of water supply). These drought definitions are strongly correlated to each other.

Several indices have been developed to assess and monitor different types of droughts in terms of their severity, duration, and magnitude (Mishra and Singh 2010; Singh et al. 2021). One of the extensively utilized drought indices is the standardized precipitation index (SPI), which is commonly used to characterize the meteorological drought because of its mathematical simplicity and easy spatial comparison due to its probabilistic nature (McKee et al. 1993). SPI is hence recommended as a standard drought index by the World Meteorological Organization (WMO) (Rashid and Beecham 2019). An extension of SPI, to analyze the hydrological drought by using runoff instead of precipitation, is the standardized runoff index (SRI) (Shukla and Wood 2008). Later, SPI is modified by incorporating the effects of both precipitation and temperature to obtain the standardized precipitation evapotranspiration index (SPEI) (Vicente-Serrano et al. 2010). Although the concept of standardized index (SI) is extensively used for drought assessment, results at multiple time scales lead to ambiguity. The copula-based approach introduced by Kao and Govindaraju (2010) provides the capacity to combine drought index values at multiple scales (such as monthly and seasonal) into one index for a comprehensive picture of drought. It is further enhanced to combine multiple droughts like meteorological, agricultural, and hydrological (including the groundwater component) droughts, known as the multivariate standardized drought index (MSDI) (Hao and Aghakouchak 2013) and integrated drought index (Shah and Mishra 2020).

All the indices derived from SI are popular because of its probabilistic nature. However, these indices assume that the statistical parameters are stationary, i.e., the mean/variance will not change over time, whereas in reality, the climatic variables (like precipitation and temperature) undergo changes due to either natural variability or anthropogenic factors (Sarhadi et al. 2016; Cheng et al. 2014; Shrestha et al. 2016; Vinnarasi et al. 2017; Vinnarasi and Dhanya 2019). Specifically precipitation, which is one of the crucial indicators to identify the onset of drought, has undergone significant changes. Recent studies observed intensification of extreme precipitation and found that wet regions are becoming dry and vice versa (Vinnarasi and Dhanya 2016; Singh et al. 2014). The stationary assumption fails to effectively capture these variations, which leads to either underestimation or overestimation of the drought parameters. This is illustrated in Fig. S1 (see online supplementary information) by comparing the stationary probability density function (PDF) with the PDFs of different changing parameters. For instance, a decreasing trend in both parameters (α: scale parameter; β: shape parameter) of the gamma distribution decreases the mean of the precipitation, increasing the probability of drier events. However, this dynamic behavior is not captured by conventional SI and leads to the underestimation of drought severity. Thus, for a realistic drought assessment, hazard preparedness, and sustainable management, a nonstationary-based drought index is essential. Generally, nonstationarity in a time series is modeled using time-varying probabilistic parameters (i.e., location, scale, and shape). Though the nonstationary model is extensively used in extreme event modeling (Mondal and Mujumdar 2015; Zhang and Zwiers 2013; El Adlouni et al. 2007; Smith 2002; Katz et al. 2002; Yilmaz and Perera 2014; Sugahara et al. 2009; Agilan and Umamahesh 2017; Cheng and AghaKouchak 2014; Luke et al. 2017; Sarhadi and Soulis 2017; Cheng et al. 2014), its application in drought assessment is less explored.

Türkeş and Tatlı (2009) were the first to propose a modified SPI to compute the dynamic behavior in precipitation using local time mean. Dubrovsky et al. (2009) reviews the existing studies and the SPI using different parameter values of gamma distribution at different times to assess climate change. Later, a standardized nonstationary precipitation index is developed using the nonstationary gamma distribution by employing time (Russo et al. 2013; Wang et al. 2015). A second approach has been to include climate indices such as El Niño–Southern Oscillation (ENSO) as covariates in defining scale and shape parameters of gamma distribution (Li et al. 2015; Zou et al. 2018; Bazrafshan and Hejabi 2018; Song et al. 2020; Sun et al. 2020; Rashid and Beecham 2019; Wang et al. 2020). These models rely on including the most relevant climate covariates for the given study basin which are identified through a statistic such as Akaike information criterion (AIC) and, as such are difficult to identify clearly (with AICs of stationary and nonstationary approaches differing by less than 1; Yu et al. 2019; Zou et al. 2018).

Most of these studies modeled nonstationarity merely by trends in the precipitation series, which reflects the change in mean. However, numerous studies reported that extreme precipitation is increasing at a faster rate than the average precipitation, which implies that the variance may have changed more than the mean of the precipitation (Goswami et al. 2006; Vittal et al. 2013). Therefore, addressing nonstationarity merely based on the trend may not bring reality to the nonstationary modeling; and, while at times, it just may merely reproduce the stationary model. A novel robust approach of time sliding window–based nonstationary model (TSW-NSM) was developed to detect the signature of nonstationarity in the distribution parameters by Vinnarasi and Dhanya (2019), and applied it to develop intensity-duration model. Besides, nonstationarity is usually modeled only as long-term change (i.e., trend) in the distribution parameters, though there might be clear evidence of short-term change (i.e., step change/sudden change) in the precipitation series (Vittal et al. 2013). The detection and incorporation of short-term change (such as a changepoint) are essential in extreme event modeling (Bazrafshan and Hejabi 2018) since it represents the transition of the probability distribution from one phase to another. The detection of short-term change in the distribution parameters, and its incorporation in nonstationary modeling has been attempted for hydrological droughts (Liu et al. 2019; Zou et al. 2018) in the last few years with promising results. The application of changepoint methodology and time sliding window approach for meteorological droughts with a large record length has not yet been attempted in authors’ knowledge. Hence, this study proposes an extended TSW-NSM approach to detect both short-term and long-term changes in the distribution parameters rather than the mere trend in the precipitation time series. Besides, drought is a multivariate phenomenon and occurs at an indefinite time scale; hence, it is difficult to model it due to the dynamic behavior of hydrological variables. Therefore, multiple SIs of various time scales and their interdependence in the context of nonstationarity need to be examined comprehensively to assess the overall status of drought (Mishra and Singh 2011; Kao and Govindaraju 2010). Owing to the limitations of traditional drought indices, a realistic index needs to be developed to extract the drought characteristics, in a changing climate, by incorporating both long-term and short-term temporal dynamics. Hence, this study proposes a robust index, termed as the nonstationary joint deficit index (NJDI), by coupling the extended TSW-NSM with the joint deficit index, to account the seasonality and to effectively model both short-term and long-term temporal dynamics. The efficacy of the proposed NJDI is further demonstrated by deploying it to identify the meteorological drought episodes over drought-prone regions of Indian subcontinent.

2. Study area and data description

The multivariate copula-based time-varying NJDI is estimated over the Indian subcontinent, which is a densely populated and agricultural-dependent country, where the management of water resources is a major challenge for water professionals. Besides, India comprises heterogeneous climatic zones, i.e., arid/semiarid in the northwest, warm–temperate in central India, subtropical humid in the northeast and foothills of Himalayas, semiarid and equatorial grassland in southern India and warm–humid coastal regions. Moreover, in India, the occurrences of floods and droughts have intensified due to the recent changes in the climate. In this study, NJDI is derived using a high-resolution (0.25° × 0.25°) daily gridded rainfall (Pai et al. 2014), obtained from the India Meteorological Department (IMD) for the period of 1901–2013. The spatial domain of data lies between 6.5° and 38.5°N and 66.5° and 100°E. The rainfall records from over 6995 rain gauge stations are converted to gridded data using the inverse distance weighted scheme (Shepard 1968). The resulting dataset has 4964 grid points. The hilly regions of upper northern and northeastern India (536 grid points) were not considered for the analyses due to the paucity of rain gauges. After a thorough quality check, the grids having continuous zero precipitation values for more than 30 years are considered as missing (498 grid points) and are exempted from the analyses. Thus, the analysis is performed over remaining 3930 grid points.

3. Methodology

Here joint deficit index (JDI), an extended version of SI (McKee et al. 1993), introduced by Kao and Govindaraju (2010) is adopted to develop NJDI. The framework of NJDI comprises (i) extraction of rainfall depth for different time scales using a modified SI approach and identifying its marginal distribution, (ii) detection of nonstationarity and estimation of changing parameters, and (iii) computing nonstationary joint deficit index. The framework is shown as a flowchart in Fig. 1.

Fig. 1.
Fig. 1.

Flowchart of the proposed framework.

Citation: Journal of Climate 36, 12; 10.1175/JCLI-D-22-0437.1

a. Modified standardized index and its marginal distribution

First, the rainfall depth for a given time scale is extracted using Eq. (1):
Xsm(y)=Xs[12(y1)+m],
Xs(t)=i=ts+1tX(i),
where X(i) represents the monthly precipitation amount measured at time i, Xs(t) represents the cumulative depth of rainfall for given time scale (s) in month t, s is time-scale size in months (1–12 months), y is year index, and m is month index: January (1), February (2), …, December (12).
Each monthly aggregated rainfall depth (XSJan,XSFeb,,XSDec) is fitted in a two-parameter gamma distribution [Eq. (2)], as suggested by McKee et al. (1993) and has been extensively used in many studies (Pai et al. 2011; Li et al. 2015; Wang et al. 2015):
XsmΓ(α,β),Fgam(x|α,β)=1βαΓ(α)0xtα1et/βdt,
where Xsm represents the time series of cumulative rainfall for month m (January, February, …, December) at time-scale s, Γ(α, β) denotes a two-parameter gamma probability distribution function with shape (α) and scale (β) parameters, and Fgam(x|α, β) represents the cumulative density function (CDF) of this gamma distribution. The goodness of fit of the distribution is checked using the Kolmogorov–Smirnov (KS) test at 5% significance level, and the series that fails the test is fitted using empirical distribution. Finally, the modified SI (SImod) for a given time window is computed, as shown in Eq. (3):
SISmod=φ1[FXsm(xsm|αsm,βsm)],
where FXsm(xsm|αsm,βsm) denotes the value of cumulative distribution function for the Xsm time series using the fitted two-parameter gamma distribution function for month m and scale-size s, and φ−1(⋅) denotes the standard inverse normal function as given by McKee et al. (1993). The primary advantage of SImod is that it incorporates the seasonal variations and overcomes the issue of autocorrelation (Kao and Govindaraju 2010). Moreover, SImod can compute the drought characteristics for a variety of time scales, which helps both short-term and long-term water resources management (Mishra and Singh 2011).

b. Detection of nonstationarity

The basic definition of nonstationarity implies change in any statistical parameter of the probabilistic distribution over time (Smith 2002). Therefore, it is essential to check for any signature of nonstationarity in the statistical parameters before deciding whether to adopt a stationary/nonstationary model. Changes in the time series alone cannot reflect the actual change in the statistical parameters. Here, the time sliding window (TSW) approach proposed by Vinnarasi and Dhanya (2019) is extended to detect both the short-term changes and the long-term changes in the statistical parameters. Short-term change, also termed as changepoint, occurs when the regime moves from one state to another in a short amount of time, that can be visualized in terms of a step change/sudden jump in a time series. The long-term change (trend) occurs in a relatively long time, and it is reflected as a concave/convex trend in the time series. The steps involved in detecting the nonstationarity in a time series, for a specific rainfall depth and time scale, are outlined below.

  1. Using a predefined time sliding window (q), the aggregated rainfall depth for a given time scale [XSm(t)=x1,x2,,xn] is reconstructed into many realizations (R1, R2, …, Rnq) as follows:
    R1={x1,x2,,xq}Rnq={xnq,xnq+1,,xn}.
  2. Each realization (R1, R2, …, Rnq) is fitted with a two-parameter gamma distribution (Γ) whose parameters are estimated using maximum likelihood approach:
    RiΓ(αRi,βRi)i=1,2,,nq.
  3. Parameter time series (PTS) are formed by using these estimates from each realization. For example,
    α=αR1,αR2,,αRnqβ=βR1,βR2,,βRnq.
  4. Then, the parametric global method is used to detect the changepoint (Killick et al. 2012) in the above mentioned α and β time series (PTS).

  5. If a changepoint is detected in one or both PTS (say in shape, α, PTS), then the series is divided into two segments: α1 denotes the shape PTS segment before the changepoint and α2 denotes the shape PTS segment after the changepoint.

  6. The nonparametric moving block bootstrap Mann–Kendall test is used to detect any significant trend in each segment (when changepoint present) or for the whole PTS (when changepoint is absent). The linear fit estimates for a very short time series will have high uncertainty, and hence, we select the changepoint such that the split time series are at least 35 years long to have reasonable confidence in the estimates.

  7. If the parameter shows a significant trend, then it is considered to be nonstationary, and the time-varying parameters are estimated by imposing a linear trend in the following manner (the procedure is shown for shape parameter α; a similar procedure is followed for scale parameter β as well).

    1. In case of changepoint (at t = t0), with a significant trend in both segments,
      α={α1=a11+a12tt=1,2,,t0α2=a21+a22tt=t0+1,,T,
      where a11 and a12 denote the coefficients of linear relationship for the first segment (t = 1, 2, …, t0). Similarly, a21 and a22 denote the coefficients of linear relationship for the second segment [t = (t0 + 1): T].
    2. In case of changepoint (at t = t0), with a significant trend in only one of the segments (say, the first segment),
      α={α1=a11+a12tt=1,2,,t0α2=a21t=t0+1,,T,
      where a11 and a12 denote the coefficients of linear relationship for the first segment (t = 1, 2, …, t0) and a21 denotes the time-invariant estimate for the second segment (t = t0 + 1, …, T). Please note that if the second segment has trend, it will have linear formulation and first segment will have a constant value.
    3. In case of no changepoint and presence of significant trend in the entire time series,
      α=a01+a02tt=1,2,,T,
      where a01 and a02 denote the coefficients of linear relationship for the entire time series (t = 1, 2, …, T).
    4. In case of no changepoint with insignificant trend, the parameter is estimated as a constant value (a00).

Vinnarasi and Dhanya (2019) used Bayesian inference (see supplementary information for the methodology details) to estimate time-varying parameters in Eqs. (7)(9) (i.e., a01, a02, …, a22). The estimates from Bayesian inference are more robust compared to ordinary least squares estimates in providing uncertainty in the estimates. However, Bayesian inference is computationally intensive and demands large storage space, as the total number of instances would be 144 (cases) × 3930 (grids). Hence, we have used linear fit estimation for all the grids in this study after verifying that the Bayesian and linear fit estimates are reasonably close for selected grids.

c. NJDI

To examine the overall status of a drought, time-varying marginal distributions of aggregated rainfall depths (usm) at different time scales of each month are combined through empirical copula to compute the joint cumulative probability. The procedure involved in computing NJDI is described below.

  1. Compute the marginal distribution for each month, and for each time scale using time-varying distribution parameters. For instance, the marginal distribution of January month for the 12-month time scale of precipitation is
    u12,tJan=Fgam(R12,tJan|α12,tJan,β12,tJan).
  2. The marginal cumulative distribution function of different months for each time scale is reconstructed with respect to the same ending time:
    u12,tistheunionofu12,tJan,u12,tFeb,,u12,tDec.
  3. Then, the joint cumulative distribution of all the reconstructed marginal distributions is computed using empirical copula (Kao and Govindaraju 2010).

    1. For single variable,
      P[U1,tu1,t,U2,tu2,t,,Ud,tu12,t]=qt.
    2. For multiple variables, steps (i) to (iii) need to be repeated for the consecutive variables, and then the joint cumulative distribution is computed using
      P[U1,tu1,t,,U12,tu12,t,U13,tυ1,t,,Ud,tυ12,t]=qt.

  4. It may happen that different combinations of (u1,t, …, u12,t) may result in same value of qt. We then use the Kendall distribution function (Kc) (Nelsen et al. 2003; Kao and Govindaraju 2010) to which will enable us to compute a quantifiable index:
    Kc(qt)=P[CU1,,Ud(U1,,Ud)qt].
  5. Finally, nonstationary standardized index (NSI; refers to drought information at a single time scale, say 6 months) and NJDI (refers to drought information at multiple time scales, say, 1, 2, …, 12 months) are computed by converting the joint cumulative distribution and cumulative distribution of each month, respectively, to the standard normal deviate with zero mean and unit variance. Then, the joint cumulative distribution of all the reconstructed marginal distributions is computed using empirical copula (Kao and Govindaraju 2010):

    1. NSImod for 12-month time scale (s = 12) is computed as
      SIsmod=SI12mod=F1(U12,t).
    2. NJDI is computed as
      NJDI=F1[Kc(qt)],
      where positive NJDI (0.5 < Kc < 1) indicates overall wet condition and negative NJDI denotes overall dry condition (0 < Kc < 0.5) (Kao and Govindaraju 2010). The drought classifications of NSI and NJDI are the same as traditional SI (McKee et al. 1993), as shown in Table 1.

Table 1.

Drought classification based on the SI values.

Table 1.

4. Results and discussion

a. Extraction of marginals

First, the aggregated rainfall series is extracted for all the months (January, February, …, December) and for 12 time scales (1–12 months), totaling 144 series for each grid, following the procedure given in section 3a, to account for the seasonal variability in India. Since gamma distribution is limited by the presence of zero values, the series having large proportions of zero values are not analyzed which are primarily located in the extremely arid regions located in the western Indian states. The marginal distribution is computed by fitting a two-parameter gamma distribution to each series, and the goodness of fit is estimated using the KS test. There are some aggregated precipitation series (5.6%) which fail the KS test and we use empirical distribution to estimate their probability distribution. The modified SI is extracted for all the aggregated rainfall series. For illustration, the modified SI for a representative grid located at 26°N, 93°E is compared with the conventional SI, as shown in Fig. S2 (see supplementary information).

b. Detection and estimation of nonstationary parameters

Once the marginal distribution of each series is found, the signature of nonstationarity is detected using the TSW approach, as mentioned in section 3b. Here, a 30-yr time sliding window is chosen to slice each series into 83 segments. The size of TSW (30 years) is chosen such that it is large enough for the goodness of fit for selected probability distribution (two-parameter gamma at each segment) (Vinnarasi and Dhanya 2019). Around 0.69% of segments are not considered because they failed to pass the KS test. The PTS showing significant trend is modeled as a linear function of time. We chose linear regression model as it adequately captured the nonstationarity present in the parameter time series without introducing too much complexity in the model and to avoid uncertainty due to the addition of parameters (Luke et al. 2017).

Initially, Bayesian inference utilizing DEMC Metropolis–Hastings algorithm (Cheng et al. 2014; Vrugt et al. 2009; see supplementary information for details about methodology) with five Markov chains is employed to estimate the nonstationary/stationary parameters. The resultant estimates from the Bayesian inference and linear fit estimates are provided in Table 2 for two representative grids. It can be observed that the median of the posterior distribution computed using Bayesian inference and the linear fit estimation are reasonably close (relative difference ≤ 10%). Therefore, linear fit estimation is applied to all the other grids, since Bayesian inference is computationally intensive and demands large storage space, as the total number of instances would be 144 (cases) × 3930 (grids).

Table 2.

Nonstationary parameter statistics for two chosen grids using Bayesian inference and linear fit estimate.

Table 2.

Any change in the distribution parameter is expected to alter the distribution of the events as shown in Fig. S1; therefore, the changes in shape (α) and scale (β) parameters of the gamma distribution affect the mean (αβ) and variance of the distribution (αβ2). To illustrate this, the parameter series of shape (α) and scale (β) parameters, time sliding window mean, time sliding window variance, and the changepoints for a grid located at 21.5°N, 79.25°E, for the month of September, are shown in Figs. 2a and 2b, respectively. It reveals a decreasing trend in shape parameter and increasing trend in scale parameter before changepoint, which is reflected as no shift in the mean but an increase in variance (i.e., the spread of distribution). However, after the changepoint, both shape and scale parameters follow a decreasing trend that is reflected as a decreasing trend in both the mean and variance. This suggests a decreasing trend in expected annual precipitation with a shrinking uncertainty band (variance), which may result in an increase in dry events. The pattern of changes in mean and variance is also revealed in the scatterplot, in which the events are spread before the changepoint and converge afterward (Figs. 2c–h). Though the changepoint of shape parameter is different from that of the scale parameter (Fig. 2a), the model generated using distinct changepoints could not capture the trend and the spread of rainfall events (Fig. 2g). Hence different alternatives of changepoints were modeled and compared for the following cases: (i) changepoint of mean adopted for both scale and shape parameters (Fig. 2d), (ii) changepoint of variance adopted for both scale and shape parameters (Fig. 2f), (iii) changepoint of shape PTS is adopted for both scale and shape parameters (Fig. 2e), and (iv) changepoint of scale PTS is adopted for both scale and shape parameters (Fig. 2c). Finally, it was observed that the changepoint in mean covers a broader range and captures the divergence and convergence in the spread of data (Fig. 2d). Since the changepoint in scale and shape parameters are close to each other (Fig. 2a), the envelope for changepoints based on variance (Fig. 2f) and scale parameter (Fig. 2e) follows the trend of the mean (Fig. 2d). We have used the changepoint in the TSW mean (Fig. 2d) for deriving NJDI in this study as it also encodes information about short-term changes in TSW variance and shape parameters. The performance of considering the changepoint in the TSW mean is good in the case of gamma distribution because the mean of the gamma distribution comprises an equal contribution of scale and shape parameters. This may be different for different distributions.

Fig. 2.
Fig. 2.

Illustration of the parameter series and the changepoints for a grid located at 21.5°N, 79.25°E for the month of September. (a) Variation of parameters α (blue line) and β (red line) of gamma distribution computed using time sliding window series and their changepoints (solid circles); (b) mean (blue line) and variance (red line) of the gamma distribution and their changepoints (solid circles). Nonstationary modeling of precipitation for the grid: (c) changepoint of α is considered for both α and β; (d) changepoint in the mean is considered as a changepoint for α and β; (e) changepoint of β is considered for both α and β; (f) changepoint in variance is considered for α and β; (g) the respective changepoints of α and β are considered; and (h) without changepoint. The light gray region, dark gray region, and black line represent the area between the 0.05 and 0.95 quantiles, the area between the 0.25 and 0.75 quantiles, and the 0.5 quantile, respectively, while the red dots indicate the observed values.

Citation: Journal of Climate 36, 12; 10.1175/JCLI-D-22-0437.1

Further, the analysis of spatial variations of the nonstationary parameters shows that only less than 10% grids show no long-term or short-term trend in the time series and more than 55% of grids have undergone significant short-term changes. Thus, more than half the grids (55%) show presence of changepoint (short-term change). The majority of grids show a negative long-term trend in the shape parameter and a positive long-term trend in the scale parameter. In the grids showing short-term changes, the majority of the grids reveal a positive trend in shape parameter before changepoint and negative trend after changepoint, while the scale parameter shows a contrary behavior. The trend in the mean depends on the relative rate of change of the scale and shape parameter, and not merely on the sign of change. As an illustration, the spatial variations of the distribution parameters for summer monsoon (June–September) at 1-month time scale are shown in Figs. S3–S6 of the supplementary information, which clearly reveal that majority of grids have undergone changes; especially the shape and scale parameters predominantly show negative and positive trends, respectively.

Further, to emphasize the effect of incorporating time-varying parameters on the distribution of the events, the marginal distribution for two grids located at 10.5°N, 78°E (southern India) and 26.5°N, 72.75°E (northwest) are compared in Fig. 3. Here the parameter series show negative/positive and positive/negative trends for shape/scale parameters, respectively, without any changepoint (Figs. 3a,b). The distributions are computed for the following cases: (i) constant parameter–stationary case (black dotted line), (ii) initial period (1925) of the time-varying parameter estimated using linear trend (black line), and (iii) latter period (1985) of the time-varying parameter estimated using linear trend (gray line). The PDF plots for the above cases are shown in Fig. 3c, which reveals an evident change in the mean and variance of the distribution, in response to the choice of the parameter. It is worthwhile to note that the probability of dry events has increased in the grid which shows negative and positive trends in shape and scale parameters, respectively (right panel), whereas an opposite behavior in the grid which shows positive and negative trends in shape and scale parameter, respectively (left panel), where the probability of the dry events has decreased. This behavior is overlooked in the case of a stationary approach. It can be inferred from Figs. S3 and S4 that dry events are increasing in southern India and the northwest region, especially in recent decades whereas a decrease in the dry events is expected in the northwest region.

Fig. 3.
Fig. 3.

Nonstationarity in the rainfall series of September (1-month time scale) for locations (left) 10.5°N, 78°E and (right) 26.5°N, 72.75°E. (a) Parameter series of α and its trend, (b) parameter series of β and its trend, and (c) PDF of rainfall depth (inset of the right panel is the zoomed portion of PDF of rainfall depth from 0 to 2 mm). Black solid represents the PDF based on 1925 parameter values, gray represents the PDF based on 1985 parameter values, and black dotted represents the PDF based on stationary parameters.

Citation: Journal of Climate 36, 12; 10.1175/JCLI-D-22-0437.1

c. Nonstationary joint deficit index

The joint cumulative probability is computed using empirical copula by combining the marginal cumulative probability of all the time scales. The stationary drought indices, computed using a constant distribution parameter, are compared with the nonstationary drought indices. We initially compare the stationary (SIsmod) and nonstationary (NSIsmod) indices for a single time scale, say, 3 months. For instance, the classification of historical droughts by both NSI3mod and SI3mod, for the month of September, for a grid located at 10.5°N, 78°E (southern India), is depicted in Fig. 4. The rainfall depth of 90–100 mm is classified as D2 (moderate drought) by the traditional approach, whereas in the case of the proposed nonstationary index, the classification for the same amount of rainfall depth is time dependent, i.e., varies from year to year, taking into account the nonstationarity in the precipitation series. The major differences between JDI and NJDI arise due to the use of a time-varying gamma distribution to model the rainfall time series. Similar observation was made by Shiau (2020) for SPI and NSPI that although both time-invariant and time-varying gamma distributions are region-specific the latter is a relative temporal reference that adapts to the trending rainfall series. A lower negative value of NJDI which uses time-varying gamma distribution represents only a more severe degree of dryness but not a greater rainfall deficiency compared to JDI. For example, SI3mod identifies the rainfall depth of 93.1 mm in the year 1968 and 98.2 mm in the year 1994 as D2 (moderate drought), while NSI3mod identifies them as D3 (severe drought) and D1 (mild drought), respectively. To have a better insight, the 30-yr moving average of the rainfall depth (magenta line) is also overlaid on the plot. For instance, the 30-yr moving-average rainfall for 1968 (1953–82) is 241.1 mm and for 1994 (1979–2008) it is 168.3 mm, respectively, and the rainfall deficit in 1968 (148 mm) is more than that of 1994 (70.1 mm). Therefore, the proposed index classifies 1968 and 1994 as D3 and D1, respectively, considering the nonstationarity in the historical rainfall records, while the traditional approach failed to capture this. Likewise, the deficits in rainfall depth in 1934 and 1952 are 171.9 and 186.9 mm, respectively, which are classified as D4 and D3 by the traditional method, though the deficit is higher in the latter year. Whereas the proposed method classifies both as D4. This further emphasizes the incapability of stationarity-based drought index in reflecting the inherent temporal dynamics in the rainfall series. Similar differences at 6-, 9-, and 12-month time scales are observed between SIsmod and NSIsmod, as shown in Figs. S7–S9 of supplementary information.

Fig. 4.
Fig. 4.

Historical drought classification of 3-month time-scale rainfall depth of September for the period 1902–2013 using the modified (a) NSI and (b) SI at grid location 10.5°N, 78°E.

Citation: Journal of Climate 36, 12; 10.1175/JCLI-D-22-0437.1

Further, it can be noted that SI3mod identifies more droughts with higher severity in the latter part of the series and ignores dry events in the initial part of the series, which may be due to the increased number of drier days (lesser rainfall depth than the initial period) in the latter period. This indicates that the drought index based on stationary assumption is biased by the drier events and fails to capture the severity of droughts appropriately, since it uses absolute thresholds to classify the droughts. This is further illustrated by comparing NSI3mod and SI3mod for another grid located at 26.5°N, 72.75°E, where SI3mod identifies fewer severe droughts in the latter period and many severe events in the initial period, as depicted in Fig. 5. This can again be attributed to the concentration of drier events in the initial part, which implies that the traditional method has overestimated the severity of drought in the earlier part and underestimated it in the latter part, whereas the proposed method classifies the drought with better accuracy and harmonious distribution. A similar pattern is observed at 6-, 9-, and 12-month time scales also, as shown in Figs. S10–S12 of supplementary information. This clearly suggests that the proposed nonstationarity-based drought index is capable of resolving the drawback of over/underestimation by the stationary methods by incorporating the dynamic behavior of the rainfall series into modeling.

Fig. 5.
Fig. 5.

Historical drought classification of 3-month time-scale rainfall depth of September for the period 1902–2013 using the modified (a) NSI and (b) SI at grid location 26.5°N, 72.75°E.

Citation: Journal of Climate 36, 12; 10.1175/JCLI-D-22-0437.1

We now compare the stationary (JDI) and nonstationary (NJDI) indices which provide information about drought at considering time scales of 1, 2, …, 12 months. The inferences from comparison of JDI and NJDI reflect the overall drought status of a location (grid). Six historical drought years are chosen, viz. 1918, 1920, 1941, 1965, 1987, and 2002 to compare the spatial distribution of NJDI and JDI (Mishra et al. 2018, 2019; Mishra 2020; Pai et al. 2011). Figures 6a–c show the NJDI and JDI for the drought years 1918, 1920, and 1941, which occurred pre-1950. It can be seen overall that JDI underpredicted the drought in all the three years, except in the northwest region. The severe drought (D4) in the year 1918 due to the El Niño is identified in 6% of grids by NJDI and less than 6% by JDI, which too, are concentrated only in the northwest. Likewise, JDI underestimates the drought in the southern, central, and northeast India for the years 1918 and 1920. This may be due to the fact that drier days are concentrated in the northwest (arid region) than other regions pre-1950, as reported by Vinnarasi and Dhanya (2016). However, a contrary pattern is revealed post-1950, which is depicted in Figs. 6d–f, where JDI overpredicts the drought in southern and northeastern India, which is again due to an increase in the aggregation of drier days in the recent decades in these regions (Vinnarasi and Dhanya 2016). The possible cause for the recent increase in the drier days are due to the increase in diurnal temperature range which in turn influenced by both global (multidecadal oscillation, global warming) and local (urbanization, dewpoint temperature, etc.) factors (Ali and Mishra 2017; Mishra et al. 2022; Vinnarasi et al. 2017; Zhang et al. 2019).

Fig. 6.
Fig. 6.

Spatial comparison of drought classification between NJDI and JDI for (a) September 1918, (b) September 1920, (c) September 1941, (d) September 1965, (e) September 1987, and (f) September 2002.

Citation: Journal of Climate 36, 12; 10.1175/JCLI-D-22-0437.1

As illustrated above for the month of September, similar observations are made for other months also. Figure 7 compares the classification of drought by JDI and NJDI for the years 1918 (early twentieth century) and 2002 (early twenty-first century) for the Indian summer monsoon months of June to October. The severity of 1918 drought is considered to be higher than the 2002 drought as the 1918 drought affected the country in an unforeseen manner (Shewale and Kumar 2005). Shewale and Kumar (2005) also noted that the Indian summer monsoon (June to September) rainfall was almost 25% lower than the long-term average annual rainfall of 880 mm in 1918 resulting in almost 70% of the area under moderate and severe droughts. The JDI and NJDI classify 54% and 60% of area under moderate to extreme drought in 1918 at end of Indian summer monsoon (October) while the 2002 drought covers 27% and 23% of country area as per JDI and NJDI, respectively (Fig. 7 and Fig. S13 in supplementary information). Thus, the drought severity-area estimates from JDI and NJDI are close to previously reported in the literature. The monthly resolution of JDI and NJDI allows use to compare the evolution of the drought as well. The 2002 drought recovered after the month of July (Fig. 7 and Fig. S13) i.e., area under moderate (D2) to extreme (D4) drought decreased from 49% in July to 23% in October (based on NJDI; see Fig. 7 and Fig. S13). The 1918 drought on the other hand maintained the severity from July (42% area under moderate to extreme drought) until the end of Indian summer monsoon in October (59%). NJDI is in closer proximity in estimating the severity of 1918 drought as described in the literature (Shewale and Kumar 2005) than JDI. In the case of 1918, JDI tends to overpredict the drought severity in the northwest and underpredict it in other regions for all the months compared to NJDI. The severity is represented by the classification of drought as D4 category especially in the month of July (Fig. 7c). Whereas, in the case of 2002, JDI tends to underpredict the drought severity in the northwest and overpredict it in other regions with respect to NJDI. These differences are clearly seen in the month of July, which can also be ascribed to the reasons mentioned earlier. Moreover, both JDI and NJDI identified October 1918 as the predominant drought, which also coincides with the major historical drought that occurred in 1918 (Mishra et al. 2019; Mishra 2020).

Fig. 7.
Fig. 7.

Spatial comparison of drought classification between NJDI and JDI for (top two rows) 1918 and (bottom two rows) 2002 in (a) June, (b) July, (c) August, (d) September, and (e) October.

Citation: Journal of Climate 36, 12; 10.1175/JCLI-D-22-0437.1

Overall, it can be inferred that stationarity-based indices (SIsmod or JDI) either underpredict or overpredict the severity of the drought compared to historical observations, and the predictions are biased toward aggregated drier events, failing to capture the temporal dynamics in the precipitation. The proposed nonstationary index (NJDI) can capture the time-varying nature of the meteorological drought and can be effectively employed for drought assessment under a changing climate. This study computes the dynamic meteorological drought characteristics using monthly accumulated rainfall data. However, the monthly variation in the meteorological variables cannot capture the intramonthly distribution which produces a skewed distribution of wet days. This may lead to misinterpretation of drought and pose a threat to agriculture. The proposed nonstationary drought index framework can be easily expanded for standardized net precipitation drought index proposed by Singh et al. (2021), which accounts for intramonthly distributions. The proposed methodology can also be employed for studying intramonthly droughts if submonthly precipitation information is available.

5. Conclusions

This study proposes a new nonstationary drought index to monitor the drought under a dynamic climate. The proposed index, termed as the nonstationary joint deficit index (NJDI), combines different time scales of modified standardized precipitation index (McKee et al. 1993; Kao and Govindaraju 2010) using empirical copula to compute the overall deficit status. Unlike traditional drought indices, the proposed index uses a nonstationary probability distribution function, which incorporates the temporal dynamics of a climatic variable into modeling. The extended time sliding window (TSW) approach is first utilized to detect the signature of nonstationarity in each distribution parameter for both long-term and short-term changes. If a significant trend is observed, then that parameter is modeled using a linear function of time; otherwise, it is kept as constant. Further, the proposed index is used to identify the meteorological drought-prone areas over India by analyzing 0.25° × 0.25° gridded daily precipitation data. The cumulative monthly rainfall depth is aggregated at different time scales (i.e., 1–12 months), and nonstationarity is detected in each series by arranging it monthwise. The following observations are derived from this study:

  1. Long-term and short-term changes are evident in both scale and shape parameters, especially in the recent decades (post-1970s). The majority of grids exhibit negative and positive trends in shape and scale parameters, respectively, which ultimately leads to an increase in the number of drier events. A contradictory pattern of shape and scale parameters is exhibited in the northwest Indian region, which resulted in a decrease in the number of dry days.

  2. The drought classification based on the proposed index could effectively capture the inherent temporal dynamics in the precipitation series, which would not have been reflected in the stationarity-based drought index. The comparison of stationary- and nonstationary-based drought index reveals that the stationary-based index underestimated the drought where frequent wet events have occurred, while overestimating it where frequent dry events have occurred since the stationarity-based drought indices are biased by the subsequent occurrence of drier events and use an absolute threshold to classify droughts.

  3. The overestimation and underestimation of major drought years by the stationary index (JDI) matches both spatially and temporally with the changing pattern of the number of dry days reported by Vinnarasi and Dhanya (2016), whereas the proposed nonstationary index (NJDI) could appropriately capture the major historical droughts with precision.

  4. The inefficacy of the stationary index (JDI) in capturing the historical drought variation can be primarily attributed to the stationarity assumption, since the presence of nonstationarity can be clearly noticed from the difference between the drought assessment of the proposed NJDI and JDI as well as presence of short- and long-term trends in the accumulated rainfall time series. The proposed NJDI is more realistic in assessing the drought characteristics such as severity and duration under a nonstationary climate.

Hence, the proposed index (NJDI) gives a potential framework for drought monitoring under a changing climate. This study is restricted with only one variable (i.e., precipitation); however, as discussed earlier, drought is a multivariate phenomenon, and therefore a multivariate nonstationary joint deficit index needs to be developed by modeling the dynamic behavior of various hydrological variables that influence drought.

Acknowledgments.

The work was supported by the funds received from Science and Engineering Research Board (SERB), DST, India, to the author C. T. Dhanya, under the SERB Women Excellence Award Scheme (Grant SB/WEA-04/2017).

Data availability statement.

This study uses 0.25° × 0.25° daily gridded rainfall dataset from Pai et al. (2014) which is openly available through the India Meteorological Department (IMD) website at https://www.imdpune.gov.in/Clim_Pred_LRF_New/Grided_Data_Download.html.

REFERENCES

  • Agilan, V., and N. V. Umamahesh, 2017: Modelling nonlinear trend for developing non-stationary rainfall intensity–duration–frequency curve. Int. J. Climatol., 37, 12651281, https://doi.org/10.1002/joc.4774.

    • Search Google Scholar
    • Export Citation
  • Ali, H., and V. Mishra, 2017: Contrasting response of rainfall extremes to increase in surface air and dewpoint temperatures at urban locations in India. Sci. Rep., 7, 1228, https://doi.org/10.1038/s41598-017-01306-1.

    • Search Google Scholar
    • Export Citation
  • Bazrafshan, J., and S. Hejabi, 2018: A non-stationary reconnaissance drought index (NRDI) for drought monitoring in a changing climate. Water Resour. Manage., 32, 26112624, https://doi.org/10.1007/s11269-018-1947-z.

    • Search Google Scholar
    • Export Citation
  • Cheng, L., and A. AghaKouchak, 2014: Nonstationary precipitation intensity-duration-frequency curves for infrastructure design in a changing climate. Sci. Rep., 4, 7093, https://doi.org/10.1038/srep07093.

    • Search Google Scholar
    • Export Citation
  • Cheng, L., A. AghaKouchak, E. Gilleland, and R. W. Katz, 2014: Non-stationary extreme value analysis in a changing climate. Climatic Change, 127, 353369, https://doi.org/10.1007/s10584-014-1254-5.

    • Search Google Scholar
    • Export Citation
  • Dubrovsky, M., M. D. Svoboda, M. Trnka, M. J. Hayes, D. A. Wilhite, Z. Zalud, and P. Hlavinka, 2009: Application of relative drought indices in assessing climate-change impacts on drought conditions in Czechia. Theor. Appl. Climatol., 96, 155171, https://doi.org/10.1007/s00704-008-0020-x.

    • Search Google Scholar
    • Export Citation
  • El Adlouni, S., T. B. M. J. Ouarda, X. Zhang, R. Roy, and B. Bobée, 2007: Generalized maximum likelihood estimators for the nonstationary generalized extreme value model. Water Resour. Res., 43, W03410, https://doi.org/10.1029/2005WR004545.

    • Search Google Scholar
    • Export Citation
  • Farahmand, A., and A. Aghakouchak, 2015: A generalized framework for deriving nonparametric standardized drought indicators. Adv. Water Resour., 76, 140145, https://doi.org/10.1016/j.advwatres.2014.11.012.

    • Search Google Scholar
    • Export Citation
  • Goswami, B. N., V. Venugopal, D. Sengupta, M. S. Madhusoodanan, and P. K. Xavier, 2006: Increasing trend of extreme rain events over India in a warming environment. Science, 314, 14421445, https://doi.org/10.1126/science.1132027.

    • Search Google Scholar
    • Export Citation
  • Hao, Z., and A. Aghakouchak, 2013: Multivariate standardized drought index: A parametric multi-index model. Adv. Water Resour., 57, 1218, https://doi.org/10.1016/j.advwatres.2013.03.009.

    • Search Google Scholar
    • Export Citation
  • Kao, S.-C., and R. S. Govindaraju, 2010: A copula-based joint deficit index for droughts. J. Hydrol., 380, 121134, https://doi.org/10.1016/j.jhydrol.2009.10.029.

    • Search Google Scholar
    • Export Citation
  • Katz, R. W., M. B. Parlange, and P. Naveau, 2002: Statistics of extremes in hydrology. Adv. Water Resour., 25, 12871304, https://doi.org/10.1016/S0309-1708(02)00056-8.

    • Search Google Scholar
    • Export Citation
  • Killick, R., P. Fearnhead, and I. A. Eckley, 2012: Optimal detection of changepoints with a linear computational cost. J. Amer. Stat. Assoc., 107, 15901598, https://doi.org/10.1080/01621459.2012.737745.

    • Search Google Scholar
    • Export Citation
  • Li, J. Z., Y. X. Wang, S. F. Li, and R. Hu, 2015: A nonstationary standardized precipitation index incorporating climate indices as covariates. J. Geophys. Res. Atmos., 120, 12 08212 095, https://doi.org/10.1002/2015JD023920.

    • Search Google Scholar
    • Export Citation
  • Liu, S., S. Huang, Y. Xie, Q. Huang, H. Wang, and G. Leng, 2019: Assessing the non-stationarity of low flows and their scale-dependent relationships with climate and human forcing. Sci. Total Environ., 687, 244256, https://doi.org/10.1016/j.scitotenv.2019.06.025.

    • Search Google Scholar
    • Export Citation
  • Luke, A., J. A. Vrugt, A. AghaKouchak, R. Matthew, and B. F. Sanders, 2017: Predicting nonstationary flood frequencies: Evidence supports an updated stationarity thesis in the United States. Water Resour. Res., 53, 54695494, https://doi.org/10.1002/2016WR019676.

    • Search Google Scholar
    • Export Citation
  • McKee, T. B., N. J. Doesken, and J. Kleist, 1993: The relation of drought frequency and duration to time scales. Eighth Conf. on Applied Climatology, Anaheim, CA, Amer. Meteor. Soc., 179–184.

  • Mishra, A. K., and V. P. Singh, 2010: A review of drought concepts. J. Hydrol., 391, 202216, https://doi.org/10.1016/j.jhydrol.2010.07.012.

    • Search Google Scholar
    • Export Citation
  • Mishra, A. K., and V. P. Singh, 2011: Drought modeling—A review. J. Hydrol., 403, 157175, https://doi.org/10.1016/j.jhydrol.2011.03.049.

    • Search Google Scholar
    • Export Citation
  • Mishra, V., 2020: Long-term (1870–2018) drought reconstruction in context of surface water security in India. J. Hydrol., 580, 124228, https://doi.org/10.1016/j.jhydrol.2019.124228.

    • Search Google Scholar
    • Export Citation
  • Mishra, V., R. Shah, S. Azhar, H. Shah, P. Modi, and R. Kumar, 2018: Reconstruction of droughts in India using multiple land-surface models (1951–2015). Hydrol. Earth Syst. Sci., 22, 22692284, https://doi.org/10.5194/hess-22-2269-2018.

    • Search Google Scholar
    • Export Citation
  • Mishra, V., A. D. Tiwari, S. Aadhar, R. Shah, M. Xiao, D. S. Pai, and D. Lettenmaier, 2019: Drought and famine in India, 1870–2016. Geophys. Res. Lett., 46, 20752083, https://doi.org/10.1029/2018GL081477.

    • Search Google Scholar
    • Export Citation
  • Mishra, V., A. D. Tiwari, and R. Kumar, 2022: Warming climate and ENSO variability enhance the risk of sequential extremes in India. One Earth, 5, 12501259, https://doi.org/10.1016/j.oneear.2022.10.013.

    • Search Google Scholar
    • Export Citation
  • Mondal, A., and P. P. Mujumdar, 2015: On the detection of human influence in extreme precipitation over India. J. Hydrol., 529, 11611172, https://doi.org/10.1016/j.jhydrol.2015.09.030.

    • Search Google Scholar
    • Export Citation
  • Nelsen, R. B., J. J. Quesada-Molina, J. A. Rodríguez-Lallena, and M. Úbeda-Flores, 2003: Kendall distribution functions. Stat. Probab. Lett., 65, 263268, https://doi.org/10.1016/j.spl.2003.08.002.

    • Search Google Scholar
    • Export Citation
  • Pai, D. S., L. Sridhar, P. Guhathakurta, and H. R. Hatwar, 2011: District-wide drought climatology of the southwest monsoon season over India based on standardized precipitation index (SPI). Nat. Hazards, 59, 17971813, https://doi.org/10.1007/s11069-011-9867-8.

    • Search Google Scholar
    • Export Citation
  • Pai, D. S., M. Rajeevan, O. P. Sreejith, B. Mukhopadhyay, and N. S. Satbha, 2014: Development of a new high spatial resolution (0.25° × 0.25°) long period (1901-2010) daily gridded rainfall data set over India and its comparison with existing data sets over the region. Mausam, 65, 118, https://doi.org/10.54302/mausam.v65i1.851.

    • Search Google Scholar
    • Export Citation
  • Rashid, M. M., and S. Beecham, 2019: Development of a non-stationary standardized precipitation index and its application to a South Australian climate. Sci. Total Environ., 657, 882892, https://doi.org/10.1016/j.scitotenv.2018.12.052.

    • Search Google Scholar
    • Export Citation
  • Russo, S., A. Dosio, A. Sterl, P. Barbosa, and J. Vogt, 2013: Projection of occurrence of extreme dry-wet years and seasons in Europe with stationary and nonstationary standardized precipitation indices. J. Geophys. Res. Atmos., 118, 76287639, https://doi.org/10.1002/jgrd.50571.

    • Search Google Scholar
    • Export Citation
  • Sarhadi, A., and E. D. Soulis, 2017: Time-varying extreme rainfall intensity-duration-frequency curves in a changing climate. Geophys. Res. Lett., 44, 24542463, https://doi.org/10.1002/2016GL072201.

    • Search Google Scholar
    • Export Citation
  • Sarhadi, A., D. H. Burn, M. Concepción Ausín, and M. P. Wiper, 2016: Time-varying nonstationary multivariate risk analysis using a dynamic Bayesian copula. Water Resour. Res., 52, 23272349, https://doi.org/10.1002/2015WR018525.

    • Search Google Scholar
    • Export Citation
  • Shah, D., and V. Mishra, 2020: Integrated drought index (IDI) for drought monitoring and assessment in India. Water Resour. Res., 56, e2019WR026284, https://doi.org/10.1029/2019WR026284.

    • Search Google Scholar
    • Export Citation
  • Shepard, D., 1968: A two-dimensional interpolation function for irregularly-spaced data. Proc. 1968 23rd ACM National Conf., New York, NY, ACM, 517–524, https://doi.org/10.1145/800186.810616.

  • Shewale, M. P., and S. Kumar, 2005: Climatological Features of Drought Incidences in India. Meteor. Monogr., No. 21, Indian Meteorological Department, 24 pp., https://www.imdpune.gov.in/Reports/drought.pdf.

  • Shiau, J.-T., 2020: Effects of gamma-distribution variations on SPI-based stationary and nonstationary drought analyses. Water Resour. Manage., 34, 20812095, https://doi.org/10.1007/s11269-020-02548-x.

    • Search Google Scholar
    • Export Citation
  • Shrestha, A. B., S. R. Bajracharya, A. R. Sharma, C. Duo, and A. Kulkarni, 2016: Observed trends and changes in daily temperature and precipitation extremes over the Koshi River basin 1975–2010. Int. J. Climatol., 37, 10661083, https://doi.org/10.1002/joc.4761.

    • Search Google Scholar
    • Export Citation
  • Shukla, S., and A. W. Wood, 2008: Use of a standardized runoff index for characterizing hydrologic drought. Geophys. Res. Lett., 35, L02405, https://doi.org/10.1029/2007GL032487.

    • Search Google Scholar
    • Export Citation
  • Singh, D., M. Tsiang, B. Rajaratnam, and N. S. Diffenbaugh, 2014: Observed changes in extreme wet and dry spells during the South Asian summer monsoon season. Nat. Climate Change, 4, 456461, https://doi.org/10.1038/nclimate2208.

    • Search Google Scholar
    • Export Citation
  • Singh, G. R., C. T. Dhanya, and A. Chakravorty, 2021: A robust drought index accounting changing precipitation characteristics. Water Resour. Res., 57, e2020WR029496, https://doi.org/10.1029/2020WR029496.

    • Search Google Scholar
    • Export Citation
  • Smith, E. P., 2002: An Introduction to Statistical Modeling of Extreme Values. Springer, 397 pp.

  • Song, Z., J. Xia, D. She, L. Zhang, C. Hu, and L. Zhao, 2020: The development of a nonstationary standardized precipitation index using climate covariates: A case study in the middle and lower reaches of Yangtze River basin, China. J. Hydrol., 588, 125115, https://doi.org/10.1016/j.jhydrol.2020.125115.

    • Search Google Scholar
    • Export Citation
  • Sugahara, S., R. P. da Rocha, and R. Silveira, 2009: Non-stationary frequency analysis of extreme daily rainfall in Sao Paulo, Brazil. Int. J. Climatol., 29, 13391349, https://doi.org/10.1002/joc.1760.

    • Search Google Scholar
    • Export Citation
  • Sun, X., Z. Li, and Q. Tian, 2020: Assessment of hydrological drought based on nonstationary runoff data. Hydrol. Res., 51, 894910, https://doi.org/10.2166/nh.2020.029.

    • Search Google Scholar
    • Export Citation
  • Türkeş, M., and H. Tatlı, 2009: Use of the standardized precipitation index (SPI) and a modified SPI for shaping the drought probabilities over Turkey. Int. J. Climatol., 29, 22702282, https://doi.org/10.1002/joc.1862.

    • Search Google Scholar
    • Export Citation
  • Vicente-Serrano, S. M., S. Beguería, and J. I. López-Moreno, 2010: A multiscalar drought index sensitive to global warming: The standardized precipitation evapotranspiration index. J. Climate, 23, 16961718, https://doi.org/10.1175/2009JCLI2909.1.

    • Search Google Scholar
    • Export Citation
  • Vinnarasi, R., and C. T. Dhanya, 2016: Changing characteristics of extreme wet and dry spells of Indian monsoon rainfall. J. Geophys. Res. Atmos., 121, 21462160, https://doi.org/10.1002/2015JD024310.

    • Search Google Scholar
    • Export Citation
  • Vinnarasi, R., and C. T. Dhanya, 2019: Bringing realism into a dynamic copula-based non-stationary intensity-duration model. Adv. Water Resour., 130, 325338, https://doi.org/10.1016/j.advwatres.2019.06.009.

    • Search Google Scholar
    • Export Citation
  • Vinnarasi, R., C. T. Dhanya, A. Chakravorty, and A. AghaKouchak, 2017: Unravelling diurnal asymmetry of surface temperature in different climate zones. Sci. Rep., 7, 7350, https://doi.org/10.1038/s41598-017-07627-5.

    • Search Google Scholar
    • Export Citation
  • Vittal, H., S. Karmakar, and S. Ghosh, 2013: Diametric changes in trends and patterns of extreme rainfall over India from pre-1950 to post-1950. Geophys. Res. Lett., 40, 32533258, https://doi.org/10.1002/grl.50631.

    • Search Google Scholar
    • Export Citation
  • Vrugt, J. A., C. J. F. Ter Braak, C. G. H. Diks, B. A. Robinson, J. M. Hyman, and D. Higdon, 2009: Accelerating Markov chain Monte Carlo simulation by differential evolution with self-adaptive randomized subspace sampling. Int. J. Nonlinear Sci. Numer. Simul., 10, 273290, https://doi.org/10.1515/IJNSNS.2009.10.3.273.

    • Search Google Scholar
    • Export Citation
  • Wang, Y., J. Li, P. Feng, and R. Hu, 2015: A time-dependent drought index for non-stationary precipitation series. Water Resour. Manage., 29, 56315647, https://doi.org/10.1007/s11269-015-1138-0.

    • Search Google Scholar
    • Export Citation
  • Wang, Y., L. Duan, T. Liu, J. Li, and P. Feng, 2020: A non-stationary standardized streamflow index for hydrological drought using climate and human-induced indices as covariates. Sci. Total Environ., 699, 134278, https://doi.org/10.1016/j.scitotenv.2019.134278.

    • Search Google Scholar
    • Export Citation
  • Yilmaz, A. G., and B. J. C. Perera, 2014: Extreme rainfall nonstationarity investigation and intensity–frequency–duration relationship. J. Hydrol. Eng., 19, 11601172, https://doi.org/10.1061/(ASCE)HE.1943-5584.0000878.

    • Search Google Scholar
    • Export Citation
  • Yu, J., T.-W. Kim, and D.-H. Park, 2019: Future hydrological drought risk assessment based on nonstationary joint drought management index. Water, 11, 532, https://doi.org/10.3390/w11030532.

    • Search Google Scholar
    • Export Citation
  • Zhang, W., G. Villarini, and M. Wehner, 2019: Contrasting the responses of extreme precipitation to changes in surface air and dew point temperatures. Climatic Change, 154, 257271, https://doi.org/10.1007/s10584-019-02415-8.

    • Search Google Scholar
    • Export Citation
  • Zhang, X., and F. W. Zwiers, 2013: Statistical indices for the diagnosing and detecting changes in extremes. Extremes in a Changing Climate, A. AghaKouchak et al., Eds., Water Science and Technology Library, Vol. 65, Springer, 1–14.

  • Zou, L., J. Xia, L. Ning, D. She, and C. Zhan, 2018: Identification of hydrological drought in eastern China using a time-dependent drought index. Water, 10, 315, https://doi.org/10.3390/w10030315.

    • Search Google Scholar
    • Export Citation

Supplementary Materials

Save
  • Agilan, V., and N. V. Umamahesh, 2017: Modelling nonlinear trend for developing non-stationary rainfall intensity–duration–frequency curve. Int. J. Climatol., 37, 12651281, https://doi.org/10.1002/joc.4774.

    • Search Google Scholar
    • Export Citation
  • Ali, H., and V. Mishra, 2017: Contrasting response of rainfall extremes to increase in surface air and dewpoint temperatures at urban locations in India. Sci. Rep., 7, 1228, https://doi.org/10.1038/s41598-017-01306-1.

    • Search Google Scholar
    • Export Citation
  • Bazrafshan, J., and S. Hejabi, 2018: A non-stationary reconnaissance drought index (NRDI) for drought monitoring in a changing climate. Water Resour. Manage., 32, 26112624, https://doi.org/10.1007/s11269-018-1947-z.

    • Search Google Scholar
    • Export Citation
  • Cheng, L., and A. AghaKouchak, 2014: Nonstationary precipitation intensity-duration-frequency curves for infrastructure design in a changing climate. Sci. Rep., 4, 7093, https://doi.org/10.1038/srep07093.

    • Search Google Scholar
    • Export Citation
  • Cheng, L., A. AghaKouchak, E. Gilleland, and R. W. Katz, 2014: Non-stationary extreme value analysis in a changing climate. Climatic Change, 127, 353369, https://doi.org/10.1007/s10584-014-1254-5.

    • Search Google Scholar
    • Export Citation
  • Dubrovsky, M., M. D. Svoboda, M. Trnka, M. J. Hayes, D. A. Wilhite, Z. Zalud, and P. Hlavinka, 2009: Application of relative drought indices in assessing climate-change impacts on drought conditions in Czechia. Theor. Appl. Climatol., 96, 155171, https://doi.org/10.1007/s00704-008-0020-x.

    • Search Google Scholar
    • Export Citation
  • El Adlouni, S., T. B. M. J. Ouarda, X. Zhang, R. Roy, and B. Bobée, 2007: Generalized maximum likelihood estimators for the nonstationary generalized extreme value model. Water Resour. Res., 43, W03410, https://doi.org/10.1029/2005WR004545.

    • Search Google Scholar
    • Export Citation
  • Farahmand, A., and A. Aghakouchak, 2015: A generalized framework for deriving nonparametric standardized drought indicators. Adv. Water Resour., 76, 140145, https://doi.org/10.1016/j.advwatres.2014.11.012.

    • Search Google Scholar
    • Export Citation
  • Goswami, B. N., V. Venugopal, D. Sengupta, M. S. Madhusoodanan, and P. K. Xavier, 2006: Increasing trend of extreme rain events over India in a warming environment. Science, 314, 14421445, https://doi.org/10.1126/science.1132027.

    • Search Google Scholar
    • Export Citation
  • Hao, Z., and A. Aghakouchak, 2013: Multivariate standardized drought index: A parametric multi-index model. Adv. Water Resour., 57, 1218, https://doi.org/10.1016/j.advwatres.2013.03.009.

    • Search Google Scholar
    • Export Citation
  • Kao, S.-C., and R. S. Govindaraju, 2010: A copula-based joint deficit index for droughts. J. Hydrol., 380, 121134, https://doi.org/10.1016/j.jhydrol.2009.10.029.

    • Search Google Scholar
    • Export Citation
  • Katz, R. W., M. B. Parlange, and P. Naveau, 2002: Statistics of extremes in hydrology. Adv. Water Resour., 25, 12871304, https://doi.org/10.1016/S0309-1708(02)00056-8.

    • Search Google Scholar
    • Export Citation
  • Killick, R., P. Fearnhead, and I. A. Eckley, 2012: Optimal detection of changepoints with a linear computational cost. J. Amer. Stat. Assoc., 107, 15901598, https://doi.org/10.1080/01621459.2012.737745.

    • Search Google Scholar
    • Export Citation
  • Li, J. Z., Y. X. Wang, S. F. Li, and R. Hu, 2015: A nonstationary standardized precipitation index incorporating climate indices as covariates. J. Geophys. Res. Atmos., 120, 12 08212 095, https://doi.org/10.1002/2015JD023920.

    • Search Google Scholar
    • Export Citation
  • Liu, S., S. Huang, Y. Xie, Q. Huang, H. Wang, and G. Leng, 2019: Assessing the non-stationarity of low flows and their scale-dependent relationships with climate and human forcing. Sci. Total Environ., 687, 244256, https://doi.org/10.1016/j.scitotenv.2019.06.025.

    • Search Google Scholar
    • Export Citation
  • Luke, A., J. A. Vrugt, A. AghaKouchak, R. Matthew, and B. F. Sanders, 2017: Predicting nonstationary flood frequencies: Evidence supports an updated stationarity thesis in the United States. Water Resour. Res., 53, 54695494, https://doi.org/10.1002/2016WR019676.

    • Search Google Scholar
    • Export Citation
  • McKee, T. B., N. J. Doesken, and J. Kleist, 1993: The relation of drought frequency and duration to time scales. Eighth Conf. on Applied Climatology, Anaheim, CA, Amer. Meteor. Soc., 179–184.

  • Mishra, A. K., and V. P. Singh, 2010: A review of drought concepts. J. Hydrol., 391, 202216, https://doi.org/10.1016/j.jhydrol.2010.07.012.

    • Search Google Scholar
    • Export Citation
  • Mishra, A. K., and V. P. Singh, 2011: Drought modeling—A review. J. Hydrol., 403, 157175, https://doi.org/10.1016/j.jhydrol.2011.03.049.

    • Search Google Scholar
    • Export Citation
  • Mishra, V., 2020: Long-term (1870–2018) drought reconstruction in context of surface water security in India. J. Hydrol., 580, 124228, https://doi.org/10.1016/j.jhydrol.2019.124228.

    • Search Google Scholar
    • Export Citation
  • Mishra, V., R. Shah, S. Azhar, H. Shah, P. Modi, and R. Kumar, 2018: Reconstruction of droughts in India using multiple land-surface models (1951–2015). Hydrol. Earth Syst. Sci., 22, 22692284, https://doi.org/10.5194/hess-22-2269-2018.

    • Search Google Scholar
    • Export Citation
  • Mishra, V., A. D. Tiwari, S. Aadhar, R. Shah, M. Xiao, D. S. Pai, and D. Lettenmaier, 2019: Drought and famine in India, 1870–2016. Geophys. Res. Lett., 46, 20752083, https://doi.org/10.1029/2018GL081477.

    • Search Google Scholar
    • Export Citation
  • Mishra, V., A. D. Tiwari, and R. Kumar, 2022: Warming climate and ENSO variability enhance the risk of sequential extremes in India. One Earth, 5, 12501259, https://doi.org/10.1016/j.oneear.2022.10.013.

    • Search Google Scholar
    • Export Citation
  • Mondal, A., and P. P. Mujumdar, 2015: On the detection of human influence in extreme precipitation over India. J. Hydrol., 529, 11611172, https://doi.org/10.1016/j.jhydrol.2015.09.030.

    • Search Google Scholar
    • Export Citation
  • Nelsen, R. B., J. J. Quesada-Molina, J. A. Rodríguez-Lallena, and M. Úbeda-Flores, 2003: Kendall distribution functions. Stat. Probab. Lett., 65, 263268, https://doi.org/10.1016/j.spl.2003.08.002.

    • Search Google Scholar
    • Export Citation
  • Pai, D. S., L. Sridhar, P. Guhathakurta, and H. R. Hatwar, 2011: District-wide drought climatology of the southwest monsoon season over India based on standardized precipitation index (SPI). Nat. Hazards, 59, 17971813, https://doi.org/10.1007/s11069-011-9867-8.

    • Search Google Scholar
    • Export Citation
  • Pai, D. S., M. Rajeevan, O. P. Sreejith, B. Mukhopadhyay, and N. S. Satbha, 2014: Development of a new high spatial resolution (0.25° × 0.25°) long period (1901-2010) daily gridded rainfall data set over India and its comparison with existing data sets over the region. Mausam, 65, 118, https://doi.org/10.54302/mausam.v65i1.851.

    • Search Google Scholar
    • Export Citation
  • Rashid, M. M., and S. Beecham, 2019: Development of a non-stationary standardized precipitation index and its application to a South Australian climate. Sci. Total Environ., 657, 882892, https://doi.org/10.1016/j.scitotenv.2018.12.052.

    • Search Google Scholar
    • Export Citation
  • Russo, S., A. Dosio, A. Sterl, P. Barbosa, and J. Vogt, 2013: Projection of occurrence of extreme dry-wet years and seasons in Europe with stationary and nonstationary standardized precipitation indices. J. Geophys. Res. Atmos., 118, 76287639, https://doi.org/10.1002/jgrd.50571.

    • Search Google Scholar
    • Export Citation
  • Sarhadi, A., and E. D. Soulis, 2017: Time-varying extreme rainfall intensity-duration-frequency curves in a changing climate. Geophys. Res. Lett., 44, 24542463, https://doi.org/10.1002/2016GL072201.

    • Search Google Scholar
    • Export Citation
  • Sarhadi, A., D. H. Burn, M. Concepción Ausín, and M. P. Wiper, 2016: Time-varying nonstationary multivariate risk analysis using a dynamic Bayesian copula. Water Resour. Res., 52, 23272349, https://doi.org/10.1002/2015WR018525.

    • Search Google Scholar
    • Export Citation
  • Shah, D., and V. Mishra, 2020: Integrated drought index (IDI) for drought monitoring and assessment in India. Water Resour. Res., 56, e2019WR026284, https://doi.org/10.1029/2019WR026284.

    • Search Google Scholar
    • Export Citation
  • Shepard, D., 1968: A two-dimensional interpolation function for irregularly-spaced data. Proc. 1968 23rd ACM National Conf., New York, NY, ACM, 517–524, https://doi.org/10.1145/800186.810616.

  • Shewale, M. P., and S. Kumar, 2005: Climatological Features of Drought Incidences in India. Meteor. Monogr., No. 21, Indian Meteorological Department, 24 pp., https://www.imdpune.gov.in/Reports/drought.pdf.

  • Shiau, J.-T., 2020: Effects of gamma-distribution variations on SPI-based stationary and nonstationary drought analyses. Water Resour. Manage., 34, 20812095, https://doi.org/10.1007/s11269-020-02548-x.

    • Search Google Scholar
    • Export Citation
  • Shrestha, A. B., S. R. Bajracharya, A. R. Sharma, C. Duo, and A. Kulkarni, 2016: Observed trends and changes in daily temperature and precipitation extremes over the Koshi River basin 1975–2010. Int. J. Climatol., 37, 10661083, https://doi.org/10.1002/joc.4761.

    • Search Google Scholar
    • Export Citation
  • Shukla, S., and A. W. Wood, 2008: Use of a standardized runoff index for characterizing hydrologic drought. Geophys. Res. Lett., 35, L02405, https://doi.org/10.1029/2007GL032487.

    • Search Google Scholar
    • Export Citation
  • Singh, D., M. Tsiang, B. Rajaratnam, and N. S. Diffenbaugh, 2014: Observed changes in extreme wet and dry spells during the South Asian summer monsoon season. Nat. Climate Change, 4, 456461, https://doi.org/10.1038/nclimate2208.

    • Search Google Scholar
    • Export Citation
  • Singh, G. R., C. T. Dhanya, and A. Chakravorty, 2021: A robust drought index accounting changing precipitation characteristics. Water Resour. Res., 57, e2020WR029496, https://doi.org/10.1029/2020WR029496.

    • Search Google Scholar
    • Export Citation
  • Smith, E. P., 2002: An Introduction to Statistical Modeling of Extreme Values. Springer, 397 pp.

  • Song, Z., J. Xia, D. She, L. Zhang, C. Hu, and L. Zhao, 2020: The development of a nonstationary standardized precipitation index using climate covariates: A case study in the middle and lower reaches of Yangtze River basin, China. J. Hydrol., 588, 125115, https://doi.org/10.1016/j.jhydrol.2020.125115.

    • Search Google Scholar
    • Export Citation
  • Sugahara, S., R. P. da Rocha, and R. Silveira, 2009: Non-stationary frequency analysis of extreme daily rainfall in Sao Paulo, Brazil. Int. J. Climatol., 29, 13391349, https://doi.org/10.1002/joc.1760.

    • Search Google Scholar
    • Export Citation
  • Sun, X., Z. Li, and Q. Tian, 2020: Assessment of hydrological drought based on nonstationary runoff data. Hydrol. Res., 51, 894910, https://doi.org/10.2166/nh.2020.029.

    • Search Google Scholar
    • Export Citation
  • Türkeş, M., and H. Tatlı, 2009: Use of the standardized precipitation index (SPI) and a modified SPI for shaping the drought probabilities over Turkey. Int. J. Climatol., 29, 22702282, https://doi.org/10.1002/joc.1862.

    • Search Google Scholar
    • Export Citation
  • Vicente-Serrano, S. M., S. Beguería, and J. I. López-Moreno, 2010: A multiscalar drought index sensitive to global warming: The standardized precipitation evapotranspiration index. J. Climate, 23, 16961718, https://doi.org/10.1175/2009JCLI2909.1.

    • Search Google Scholar
    • Export Citation
  • Vinnarasi, R., and C. T. Dhanya, 2016: Changing characteristics of extreme wet and dry spells of Indian monsoon rainfall. J. Geophys. Res. Atmos., 121, 21462160, https://doi.org/10.1002/2015JD024310.

    • Search Google Scholar
    • Export Citation
  • Vinnarasi, R., and C. T. Dhanya, 2019: Bringing realism into a dynamic copula-based non-stationary intensity-duration model. Adv. Water Resour., 130, 325338, https://doi.org/10.1016/j.advwatres.2019.06.009.

    • Search Google Scholar
    • Export Citation
  • Vinnarasi, R., C. T. Dhanya, A. Chakravorty, and A. AghaKouchak, 2017: Unravelling diurnal asymmetry of surface temperature in different climate zones. Sci. Rep., 7, 7350, https://doi.org/10.1038/s41598-017-07627-5.

    • Search Google Scholar
    • Export Citation
  • Vittal, H., S. Karmakar, and S. Ghosh, 2013: Diametric changes in trends and patterns of extreme rainfall over India from pre-1950 to post-1950. Geophys. Res. Lett., 40, 32533258, https://doi.org/10.1002/grl.50631.

    • Search Google Scholar
    • Export Citation
  • Vrugt, J. A., C. J. F. Ter Braak, C. G. H. Diks, B. A. Robinson, J. M. Hyman, and D. Higdon, 2009: Accelerating Markov chain Monte Carlo simulation by differential evolution with self-adaptive randomized subspace sampling. Int. J. Nonlinear Sci. Numer. Simul., 10, 273290, https://doi.org/10.1515/IJNSNS.2009.10.3.273.

    • Search Google Scholar
    • Export Citation
  • Wang, Y., J. Li, P. Feng, and R. Hu, 2015: A time-dependent drought index for non-stationary precipitation series. Water Resour. Manage., 29, 56315647, https://doi.org/10.1007/s11269-015-1138-0.

    • Search Google Scholar
    • Export Citation
  • Wang, Y., L. Duan, T. Liu, J. Li, and P. Feng, 2020: A non-stationary standardized streamflow index for hydrological drought using climate and human-induced indices as covariates. Sci. Total Environ., 699, 134278, https://doi.org/10.1016/j.scitotenv.2019.134278.

    • Search Google Scholar
    • Export Citation
  • Yilmaz, A. G., and B. J. C. Perera, 2014: Extreme rainfall nonstationarity investigation and intensity–frequency–duration relationship. J. Hydrol. Eng., 19, 11601172, https://doi.org/10.1061/(ASCE)HE.1943-5584.0000878.

    • Search Google Scholar
    • Export Citation
  • Yu, J., T.-W. Kim, and D.-H. Park, 2019: Future hydrological drought risk assessment based on nonstationary joint drought management index. Water, 11, 532, https://doi.org/10.3390/w11030532.

    • Search Google Scholar
    • Export Citation
  • Zhang, W., G. Villarini, and M. Wehner, 2019: Contrasting the responses of extreme precipitation to changes in surface air and dew point temperatures. Climatic Change, 154, 257271, https://doi.org/10.1007/s10584-019-02415-8.

    • Search Google Scholar
    • Export Citation
  • Zhang, X., and F. W. Zwiers, 2013: Statistical indices for the diagnosing and detecting changes in extremes. Extremes in a Changing Climate, A. AghaKouchak et al., Eds., Water Science and Technology Library, Vol. 65, Springer, 1–14.

  • Zou, L., J. Xia, L. Ning, D. She, and C. Zhan, 2018: Identification of hydrological drought in eastern China using a time-dependent drought index. Water, 10, 315, https://doi.org/10.3390/w10030315.

    • Search Google Scholar
    • Export Citation
  • Fig. 1.

    Flowchart of the proposed framework.

  • Fig. 2.

    Illustration of the parameter series and the changepoints for a grid located at 21.5°N, 79.25°E for the month of September. (a) Variation of parameters α (blue line) and β (red line) of gamma distribution computed using time sliding window series and their changepoints (solid circles); (b) mean (blue line) and variance (red line) of the gamma distribution and their changepoints (solid circles). Nonstationary modeling of precipitation for the grid: (c) changepoint of α is considered for both α and β; (d) changepoint in the mean is considered as a changepoint for α and β; (e) changepoint of β is considered for both α and β; (f) changepoint in variance is considered for α and β; (g) the respective changepoints of α and β are considered; and (h) without changepoint. The light gray region, dark gray region, and black line represent the area between the 0.05 and 0.95 quantiles, the area between the 0.25 and 0.75 quantiles, and the 0.5 quantile, respectively, while the red dots indicate the observed values.

  • Fig. 3.

    Nonstationarity in the rainfall series of September (1-month time scale) for locations (left) 10.5°N, 78°E and (right) 26.5°N, 72.75°E. (a) Parameter series of α and its trend, (b) parameter series of β and its trend, and (c) PDF of rainfall depth (inset of the right panel is the zoomed portion of PDF of rainfall depth from 0 to 2 mm). Black solid represents the PDF based on 1925 parameter values, gray represents the PDF based on 1985 parameter values, and black dotted represents the PDF based on stationary parameters.

  • Fig. 4.

    Historical drought classification of 3-month time-scale rainfall depth of September for the period 1902–2013 using the modified (a) NSI and (b) SI at grid location 10.5°N, 78°E.

  • Fig. 5.

    Historical drought classification of 3-month time-scale rainfall depth of September for the period 1902–2013 using the modified (a) NSI and (b) SI at grid location 26.5°N, 72.75°E.

  • Fig. 6.

    Spatial comparison of drought classification between NJDI and JDI for (a) September 1918, (b) September 1920, (c) September 1941, (d) September 1965, (e) September 1987, and (f) September 2002.

  • Fig. 7.

    Spatial comparison of drought classification between NJDI and JDI for (top two rows) 1918 and (bottom two rows) 2002 in (a) June, (b) July, (c) August, (d) September, and (e) October.

All Time Past Year Past 30 Days
Abstract Views 778 212 0
Full Text Views 415 210 22
PDF Downloads 456 223 19