A Bayesian Hierarchical Framework for Postprocessing Daily Streamflow Simulations across a River Network

Álvaro Ossandón aDepartment of Civil, Environmental and Architectural Engineering, University of Colorado Boulder, Boulder, Colorado
bDepartamento de Obras Civiles, Universidad Técnica Federico Santa María, Valparaíso, Chile

Search for other papers by Álvaro Ossandón in
Current site
Google Scholar
PubMed
Close
,
Nanditha J. S. cCivil Engineering, Indian Institute of Technology, Gandhinagar, India

Search for other papers by Nanditha J. S. in
Current site
Google Scholar
PubMed
Close
,
Pablo A. Mendoza dDepartment of Civil Engineering, Universidad de Chile, Santiago, Chile
eAdvanced Mining Technology Center, Universidad de Chile, Santiago, Chile

Search for other papers by Pablo A. Mendoza in
Current site
Google Scholar
PubMed
Close
,
Balaji Rajagopalan aDepartment of Civil, Environmental and Architectural Engineering, University of Colorado Boulder, Boulder, Colorado
fCooperative Institute for Research in Environmental Sciences, University of Colorado Boulder, Boulder, Colorado

Search for other papers by Balaji Rajagopalan in
Current site
Google Scholar
PubMed
Close
, and
Vimal Mishra cCivil Engineering, Indian Institute of Technology, Gandhinagar, India

Search for other papers by Vimal Mishra in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

Despite the potential and increasing interest in physically based hydrological models for streamflow forecasting applications, they are constrained in terms of agility to generate ensembles. Hence, we develop and test a Bayesian hierarchical model (BHM) to postprocess physically based hydrologic model simulations at multiple sites on a river network, with the aim to generate probabilistic information (i.e., ensembles) and improve raw model skill. We apply our BHM framework to daily summer (July–August) streamflow simulations at five stations located in the Narmada River basin in central India, forcing the Variable Infiltration Capacity (VIC) model with observed rainfall. In this approach, daily observed streamflow at each station is modeled with a conditionally independent probability density function with time varying distribution parameters, which are modeled as a linear function of potential covariates that include VIC outputs and meteorological variables. Using suitable priors on the parameters, posterior parameters and predictive posterior distributions—and thus ensembles—of daily streamflow are obtained. The best BHM model considers a gamma distribution and uses VIC streamflow and a nonlinear covariate formulated as the product of VIC streamflow and 2-day precipitation spatially averaged across the area between the current and upstream station. The second covariate enables correcting the time delay in flow peaks and nonsystematic biases in VIC streamflow. The results show that the BHM postprocessor increases probabilistic skill in 60% compared to raw VIC simulations, providing reliable ensembles for most sites. This modeling approach can be extended to combine forecasts from multiple sources and provide skillful multimodel ensemble forecasts.

© 2022 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: Álvaro Ossandón, alvaro.ossandon@colorado.edu

Abstract

Despite the potential and increasing interest in physically based hydrological models for streamflow forecasting applications, they are constrained in terms of agility to generate ensembles. Hence, we develop and test a Bayesian hierarchical model (BHM) to postprocess physically based hydrologic model simulations at multiple sites on a river network, with the aim to generate probabilistic information (i.e., ensembles) and improve raw model skill. We apply our BHM framework to daily summer (July–August) streamflow simulations at five stations located in the Narmada River basin in central India, forcing the Variable Infiltration Capacity (VIC) model with observed rainfall. In this approach, daily observed streamflow at each station is modeled with a conditionally independent probability density function with time varying distribution parameters, which are modeled as a linear function of potential covariates that include VIC outputs and meteorological variables. Using suitable priors on the parameters, posterior parameters and predictive posterior distributions—and thus ensembles—of daily streamflow are obtained. The best BHM model considers a gamma distribution and uses VIC streamflow and a nonlinear covariate formulated as the product of VIC streamflow and 2-day precipitation spatially averaged across the area between the current and upstream station. The second covariate enables correcting the time delay in flow peaks and nonsystematic biases in VIC streamflow. The results show that the BHM postprocessor increases probabilistic skill in 60% compared to raw VIC simulations, providing reliable ensembles for most sites. This modeling approach can be extended to combine forecasts from multiple sources and provide skillful multimodel ensemble forecasts.

© 2022 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: Álvaro Ossandón, alvaro.ossandon@colorado.edu

1. Introduction

Floods are among the major causes of yearly infrastructure damage and loss of human lives across the globe (Tanoue et al. 2016; Wallemacq and House 2018). These events are especially impactful in India, where floods driven by summer monsoon (June–September) rainfall events—which contribute to more than 80% of total annual rainfall—are frequently observed and are caused by synoptic-scale cyclonic depressions (Hunt et al. 2016; Hunt and Fletcher 2019). In addition, an increase in frequency and intensity of extreme precipitation events is projected due to anthropogenic climate change (Wasko and Sharma 2017; Ali and Mishra 2018; Papalexiou and Montanari 2019). Thus, skillful daily streamflow forecasts on river networks are crucial for flood mitigation, especially in rainfall-driven river basins, as in the case of India during the monsoon season. However, efforts to develop early flood warning systems in India are limited (CWC 2012).

India’s current operational flood forecasting system is based on statistical models with gauge-to-gauge discharge correlations and coaxial correlations that use discharge and rainfall for most locations (CWC 2015). While gauge-to-gauge correlation-based streamflow forecasting frameworks may be effective for large river basins, they may also yield lower skill in basins with short times of concentration (Ghimire and Krajewski 2020; Krajewski et al. 2020). In view of this, the implementation of physically based hydrological models is an attractive alternative since they can be configured in a spatially distributed fashion, providing detailed information on the spatiotemporal evolution of states and fluxes (e.g., Ercolani and Castelli 2017; Jay-Allemand et al. 2020). Nevertheless, physically based models suffer from structural, parameter, and data errors, yielding biases in hydrologic model output (Liu and Gupta 2007; Clark et al. 2017). Thus, the development of postprocessing techniques (Li et al. 2017b) that allow reducing these errors is needed, although they are rarely implemented in an automated way (Pagano et al. 2014).

Postprocessing methods were originally developed in atmospheric sciences to correct numerical weather prediction (NWP) model output (e.g., Bremnes 2004; Glahn and Lowry 1972; Mendoza et al. 2015b). These methods have evolved for hydrological applications, either to improve the raw skill and accuracy of meteorological variables that force hydrology models (e.g., Clark et al. 2004; Verkade et al. 2013; Grillakis et al. 2018) or for direct application to hydrological models’ output (i.e., streamflow) for operational flood forecasting (e.g., Zalachori et al. 2012; Demargne et al. 2014; Muhammad et al. 2018). In principle, postprocessing techniques seek to estimate the conditional probability of a variable given “raw” (single value or ensemble) simulations or forecasts. In the case of hydrological variables such as streamflow, these methods have to deal with additional difficulties of heteroscedastic errors, poor representation of extreme events due to limited samples (conditional bias), and reduced skill in forecasts of streamflow peak timing, which are important for operational flood forecasting (Li et al. 2017b).

Several postprocessing approaches have been implemented for hydrological simulations or forecasts produced with physically based models, including nonparametric methods, regression-based techniques, ensemble dressing methods, and conditional distribution-based approaches (for a detailed review, see Li et al. 2017b). Nonparametric approaches attempt to map the target variable to historical distributions in order to reduce biases, and examples include quantile mapping (QM; Hashino et al. 2007; Lucatero et al. 2018; Tiwari et al. 2022) and conditional-bias-penalized indicator cokriging (CBP-ICK; Brown and Seo 2010, 2013). Regression-based methods used for hydrological forecasts include the General Linear Model postprocessor (GLMPP; Zhao et al. 2011; Ye et al. 2014, 2015), autoregressive models (AR; Bogner and Pappenberger 2011; McInerney et al. 2017, 2018; Woldemeskel et al. 2018), restricted autoregressive models (restricted AR; Li et al. 2015, 2016, 2017a; Bennett et al. 2021), and quantile regression (QR; Weerts et al. 2011; Tyralis et al. 2019). Conditional distribution-based methods applied for hydrological forecasts include the hydrological uncertainty processor (HUP; Krzysztofowicz and Kelly 2000; Reggiani et al. 2009), model conditional processor (MCP; Todini 2008; Anele et al. 2018), and Bayesian joint probability (BJP; Wang et al. 2009; Pokhrel et al. 2013; Zhao et al. 2015, 2016).

Since physically based hydrological models contain deterministic conservation equations that simulate hydrological processes and their interactions, ensemble simulations or forecasts can be produced when multiple initial conditions, parameterization options, parameter sets, or meteorological traces are provided. An example of the latter is the ensemble streamflow prediction (ESP) system (Day 1985)—used for operational purposes—which initializes a hydrologic model with current land surface hydrologic conditions and drives it with an ensemble of historical meteorological forcings. However, many agencies (e.g., National Weather Service) provide only a few quantiles, instead of the forecast’s distribution. Alternatively, some postprocessing approaches can provide forecast ensembles given deterministic raw forecasts. Among these, AR, restricted AR, BJP, and MCP can deal with the heteroscedasticity of the residuals through the use of variance stabilizing transformation and bivariate normal distributions (Zhao et al. 2015; McInerney et al. 2018; Li et al. 2015) or multivariate truncated normal distributions (Todini 2008; Anele et al. 2018). However, most of these approaches rely on the autocorrelation of residuals (streamflow time persistence), and thus a continuous record of observed streamflow is required for their implementation in real time, which can be an obstacle in basins where recent observations are not readily available.

Given the existing techniques, their limitation for real-time application, and the need for risk-based flood hazard mitigation, we build upon the Bayesian hierarchical network model developed by Ossandón et al. (2021) to propose a Bayesian hierarchical model (BHM) for postprocessing daily streamflow simulations across a river network. Namely, our BHM formulation aims to reduce residual errors arising from historical forcings, model structure, and parametric uncertainties, providing a starting point for future applications to real-time forecasting. Specifically, we seek to test if the proposed BHM framework (i) enhances the skill from a deterministic, physically based hydrological model; (ii) deals with the heteroscedasticity of the residuals; and (iii) is capable of correcting systematic and nonsystematic biases. To this end, we demonstrate the potential of the BHM to postprocess daily monsoon (July–August) streamflow simulations produced with the Variable Infiltration Capacity (VIC; Liang et al. 1994, 1996) model at five gauges in the Narmada River basin network in central India. For the remainder of this paper, we use the term “daily monsoon streamflow” to refer to daily streamflows over the period July–August.

2. Study domain and observational datasets

a. Study region

We used the Narmada River basin in India as a testbed to develop and evaluate the Bayesian hierarchical framework proposed here. The Narmada River basin (97 882 km2) is a narrow and elongated catchment (Fig. 1) that stretches from the headwaters in the Amarkantak hills of central India toward the Arabian Sea. The basin-averaged mean annual rainfall is 1120 mm (period 1951–2018), with most of it arriving during the summer monsoon season (June–September), and the upper portion of the basin receives larger precipitation amounts than the lower basin (Banerjee 2009). Because this basin is an essential water source for the populous states of Madhya Pradesh and Gujarat, several reservoirs have been built within the basin for irrigation during the nonmonsoon season, while others have been designed for flood control. Flood events in the basin occur mainly during July–August, the target season of our application.

Fig. 1.
Fig. 1.

Map of the Narmada River basin in India, showing 200-m elevation bands, the locations of five subbasin outlets (Sandiya, Hoshangabad, Handia, Mandleshwar, and Garudeshwar), and some of the major dams in the basin: Bargi, Tawa, Indirasagar, Jobat, and Sardar Sarovar (from upstream to downstream direction). The gray circle represents an artificial gauge created for the postprocessing model.

Citation: Journal of Hydrometeorology 23, 6; 10.1175/JHM-D-21-0167.1

b. Historical datasets

We used a 0.25° gridded daily precipitation dataset for the period 1951–2018, developed by the India Meteorology Department (IMD; Pai et al. 2014) through an inverse distance weighting scheme that interpolates observations from 6995 meteorological stations across India (Pai et al. 2014). Gridded daily maximum and minimum temperature fields were obtained by spatially interpolating observations from 395 stations from India to the same 0.25° horizontal resolution grid, using the synergraphic mapping approach (SYMAP; Maurer et al. 2002). These daily gridded precipitation and maximum and minimum temperature fields from IMD have been widely used in previous hydrometeorological studies (e.g., Ali et al. 2019; Shah and Mishra 2016). Daily wind speed was obtained from the National Centers for Environmental Prediction (NCEP)–National Centers for Atmospheric Research (NCAR) reanalysis dataset (Kalnay et al. 1996; Kistler et al. 2001), and regridded to the same 0.25° spatial resolution grid to keep consistency with the other variables.

Observed daily streamflow at five gauges in the Narmada River basin—Sandiya, Hoshangabad, Handia, Mandleshwar, and Garudeshwar (Fig. 1)—were obtained from the India Water Resource Information System (IWRIS) for the period 1970–2014.

3. Approach

Our proposed framework has two components: (i) deterministic streamflow simulations from a physically based hydrological model and (ii) a Bayesian hierarchical model to postprocess the deterministic output to provide posterior distribution of streamflow. These two components are described below.

a. Hydrological modeling

1) Model description and configuration

We used the VIC macroscale hydrological model with three soil layers (Liang et al. 1994, 1996; Liang and Xie 2001; Cherkauer et al. 2003), implemented at a 0.25° horizontal resolution. VIC is a semidistributed model that independently solves water and energy balance at each grid cell. The top soil layer is the dynamic layer representing the saturation excess and infiltration excess runoff, while the bottom layers simulated delayed baseflow responses. In the third soil layer, the slow runoff component is estimated nonlinearly using the ARNO method (Franchini and Pacciani 1991). A stand-alone routing model developed by Lohmann et al. (1996) is used for routing the gridded runoff values generated by the VIC model at specific locations along the main river reach. The routing model solves linearized St. Venant equations for channel routing, with a unit hydrograph model for estimating overland flow (Lohmann et al. 1996). The VIC model has been widely used for simulating streamflow, soil moisture, and other water balance components in the United States, India, and other parts of the world (Schumann et al. 2013; Zhou and Guo 2013; Wu et al. 2014; Bohn and Vivoni 2016; Shah and Mishra 2016; Vásquez et al. 2021). Although the construction of multiple reservoirs in our study domain regulates the river flow, we did not consider their influence for simulating streamflow.

In addition to gridded meteorological forcings (i.e., daily precipitation, maximum and minimum temperatures at 2 m, and wind speed at 10-m height from the surface), the VIC model requires land surface characteristics such as soil, vegetation, and topographic parameters as inputs. We obtained the vegetation parameters from Advanced Very High-Resolution Radiometers (AVHRR; Hansen et al. 2000; Sheffield and Wood 2007) and vegetation library from the Land Data Assimilation System (LDAS; Rodell et al. 2004). Soil parameters were obtained from the Harmonized World Soil Database (HWSD; FAO et al. 2012).

2) Calibration and validation

We calibrated and evaluated the VIC model at three locations in the Narmada River basin (Sandiya, Handia, and Garudeshwar) using observed daily streamflow. The model parameters were calibrated manually using data for the period (1980–94) based on the Pearson’s correlation coefficient (R) and the Nash–Sutcliffe efficiency (NSE; Nash and Sutcliffe 1970), using the first five years of the record (1951–55) as spinup. Due to the lack of a common period with streamflow records at all three stations, the model evaluation was conducted during the 1998–2002 period at the Sandiya and Handia stations, and during the 1996–2000 period at Garudeshwar station. Values of the calibrated parameters can be found in Table S2. Scatterplots for the calibration and validation periods; performance metrics R, NSE, and the Kling–Gupta efficiency (KGE; Gupta et al. 2009); and times series with observed and simulated streamflow for the validation period can be found in Fig. S1 in the online supplemental material. We obtained R > 0.8 between observed and simulated daily streamflow for the calibration period and all subbasins, while NSE and KGE values were higher than 0.6 and 0.7, respectively. For the validation period, we obtained R, NSE, and KGE values higher than 0.8, 0.5, and 0.65, respectively. Additionally, a negative bias for high flows is observed at the gauges (see time series in Fig. S1). After calibrating the model, we simulated streamflow at the five locations in the Narmada basin (Fig. 1) for the 1951–2018 period.

b. Bayesian hierarchical model for streamflow postprocessing

Streamflow probability distributions—typically estimated from ensembles—are required to provide risk estimates for operational and management decisions. Although ensembles can be produced with complex, process-based models, their lack of agility (Mendoza et al. 2015a)—in particular, the computational cost associated with high space-time resolution—makes it difficult to implement in real-time forecasting and decision-making applications. To overcome this issue, we propose a Bayesian hierarchical model to postprocess simulated daily streamflow from the VIC model at five gauges of the Narmada River basin: 1) Garudeshwar, 2) Mandleshwar, 3) Handia, 4) Hoshangabad, and 5) Sandiya. The numbers in brackets are used for model formulation, which is based on the Bayesian hierarchical network model proposed by Ossandón et al. (2021). Although we postprocess streamflow simulations in this paper, this framework is envisioned for implementation in real-time forecasting.

1) General model structure

The BHM structure takes advantage of the river network architecture by treating streamflow generation as a spatial Markov process (Ravindranath et al. 2019; Ossandón et al. 2021). In the first layer of the hierarchical model structure, also known as the data layer, the daily streamflow at each location is assumed to follow a conditionally independent probability density function f—typically, a positive distribution that can model skewed data (e.g., gamma, lognormal, Weibull). In the second layer (i.e., the latent or process layer), the density function parameters depend on temporally varying covariates. Suitable priors on the parameters are the final layer that complete the model formulation. Thus, daily observed streamflow at each gauge i and the day t, Qt(i), is expressed as
f{Qt(i)|αt(i),λt(i),Xtk(i),QVIC,tk[jJ(i)]},i=1,2,,n,
where αt(i) and λt(i) are the distribution parameters at the gauge i and day t, Xtk(i), denotes the set of hydrometeorological covariates at gauge i for day tk, and QVIC,tk[jJ(i)] is the VIC streamflow at the gauge j for day tk, where J(i) corresponds to an index vector of length n with only zeros except for the position of the feeders that contain ones [e.g., for Garudeshwar, J(i = 1), only its second element (Mandleshwar) contains a nonzero value, see Fig. 1]. The distribution parameters can be expressed in terms of the expected value [ μt(i)] and variance [ (σt(i))2] of Qt(i):
αt(i)=g1[μt(i),(σt(i))2],λt(i)=g2[μt(i),(σt(i))2]
In the second layer, μt(i), and σt(i) are modeled as linear functions of m temporally varying covariates (including the VIC streamflow). To avoid fitting issues related to zero rainfall values, we consider the following step functions for μt(i) and σt(i):
μt(i)={β1(i)+β2(i)QVIC,tk[jJ(i)]Ptk(i)=0β3(i)+β4(i)QVIC,tk[jJ(i)]+β5(i)X1,tk(i)++βl+3(i)Xl,tk(i)++βm+3(i)Xm1,tk(i)Ptk(i)>0
σt(i)={ϕ1(i)+ϕ2(i)QVIC,tk[jJ(i)]Ptk(i)=0ϕ3(i)+ϕ4(i)QVIC,tk[jJ(i)]+ϕ5(i)X1,tk(i)++ϕl+3(i)Xl,tk(i)++ϕm+3(i)Xm1,tk(i)Ptk(i)>0
where βj(i) and ϕj(i) are the regression coefficients for μt(i) and σt(i), respectively; Ptk(i) is the precipitation covariate at gauge i for day tk [see description in section 3b(2)]; and Xl,tk(i) is the value of covariate l at gauge i for day tk. All model covariates change with time to help capture nonstationarity. Note that even when Eqs. (3) and (4) consider that μt(i), and σt(i) can be modeled with multiple linear regressions for simplicity, nonlinear terms of the covariates can be incorporated to capture nonlinear nonstationarity. Readers can find more details about the likelihood function in Eqs. (S1) and (S2). For the regression coefficients, we considered weakly normal priors centered at 0, with a standard deviation of 10, i.e., p(βj(i))N(0,10) and p(ϕj(i))N(0,10).

2) Potential covariates

We considered three types of potential covariates for the monsoon season: (i) daily streamflow simulated with the VIC model (hereafter referred to as “VIC streamflow”), (ii) spatially averaged precipitation for the area between two consecutive gauges, and (iii) a nonlinear covariate obtained from the product between simulated streamflow and precipitation. We did not consider the lagged observed streamflow as a potential covariate because such measurements are discontinuous of it since 2019 at most of the gauges, precluding the implementation of the BHM framework in real time.

For the first type of covariate (VIC streamflow for station i), we considered both daily VIC streamflow at gauge i ( QVIC,t(i)) and 1-day lagged daily VIC streamflow from the upstream gauge i + 1 ( QVIC,t1(i+1)) as potential covariates. For Sandiya (i.e., the headwater gauge), we also considered the simulated VIC streamflow at an artificial feeder gauge ( QVIC,t1(i+1), gray circle in Fig. 1). For the second type of covariate, we considered spatially averaged precipitation amounts over s days—i.e., sum of precipitation amounts for days t, t − 1, …, and t s + 1—for the area between two successive gauges i and i + 1 ( Psd,t(i)). We tested precipitation durations of s = 1, 2, 3, and 4 days for spatially averaged precipitation.

Even after parameter calibration, conditional biases in VIC streamflow and time delays in simulated peak flows were detected at the four stations. Figure 2 displays the scatterplot of VIC streamflow against observations at the Handia gauge during the monsoon seasons for the period 1978–2014, and their time series for 2013 (wet year). The results show that, although observed streamflow is highly correlated with VIC streamflow for the entire period (R = 0.76, Fig. 2a), there is a conditional bias, i.e., a positive (negative) bias for low (high) flows (see Fig. 3b and Fig. S2 for results in other gauges). Further, a time delay of 1 day is observed for VIC streamflow peaks compared to observed streamflow peaks. To address this issue, we included a nonlinear covariate for gauge i ( NLt(i)), computed as the product between VIC streamflow at gauge i ( QVIC,t(i)) and 2-day spatially averaged precipitation from the area between two gauges ( P2d,t(i)), which is synchronized with observed streamflow at each station.

Fig. 2.
Fig. 2.

(a) Scatterplot with VIC streamflow vs observations at the Handia gauge during monsoon season for the period 1978–2014 and (b) time series for 2013 (high flow year). Light blue ellipses in (b) show examples of conditional bias and time delay of VIC streamflow simulations. The R value is statistically significant (p value < 0.05).

Citation: Journal of Hydrometeorology 23, 6; 10.1175/JHM-D-21-0167.1

Fig. 3.
Fig. 3.

Conceptual diagram of the network Bayesian model for the Narmada River basin; Qt(i) is the observed streamflow at gauge i and day t (predictand), NLt(i) is the product between daily VIC streamflow at gauge i and 2-day spatial average precipitation from the area between stations i and i + 1 at day t, and QVIC,t(i) to the daily VIC streamflow for gauge i and day t.

Citation: Journal of Hydrometeorology 23, 6; 10.1175/JHM-D-21-0167.1

We explored the strength of the relationship between the covariates and daily streamflow at each gauge by computing the Spearman’s rank correlation coefficient (see Table S3), finding statistically significant correlations.

3) Implementation and model fitting

We fitted various candidate BHMs using different combinations of covariates and alternative unconditional distributions of daily streamflow (i.e., gamma and lognormal). For the gamma distribution, rate and shape parameters were formulated as a function of the mean and variance (Wilks 2011). For the lognormal distribution, we applied a log transformation to the predictands and covariates, and the standard deviation was set stationary because of its positive restriction (σ > 0). Since the Garudeshwar gauge has missing periods (summers of 1988, 1989, and 1995), we fitted the BHM including only years with complete records (34 years for Garudeshwar, and 37 years for the other gauges) within the period 1978–2014. To select the best BHM, we used the leave-one-out cross-validation information criteria (LOOIC; Vehtari et al. 2017), which enables the estimation of pointwise out-of-sample prediction accuracy using the log-likelihood evaluated at the posterior simulations of the parameter values. Luo and Al-Harbi (2017) showed that LOOIC performs better than other traditional model selection metrics such as the Akaike information criterion (Akaike 1974) and the deviance information criterion (DIC; Spiegelhalter et al. 2002). For each candidate BHM, we computed the LOOIC using the R package loo (Vehtari et al. 2020) and selected the model with the lowest LOOIC value.

Candidate BHMs were implemented in R (R Core Team 2017) using the program STAN (Stan Development Team 2014) and the R package RStan (Stan Development Team 2020), which provides an interface from R to the STAN library for Bayesian data analysis. Posterior parameter distributions and predictive posterior streamflow distributions (ensembles) were estimated using the No-U-Turn Sampler (NUTS; Hoffman and Gelman 2014) for the Markov chain Monte Carlo method (Gelman and Hill 2006; Robert and Casella 2011), using the priors described in section 3b(1). We ran three parallel chains—each with a length of 4000 simulations (iterations) and a burn-in size of 2000 to ensure convergence—with different initial values. To reduce the sample dependence (autocorrelation), we chose a thinning factor of 2, yielding 3000 samples (1000 samples from each chain). The scale reduction factor R^ (Gelman and Rubin 1992) was used to check model convergence—i.e., R^ values lower than 1.1. The latter condition was achieved in all of our runs for each candidate BHM, indicating model convergence. Consequently, the posterior distributions of the parameters and the predictive posterior distribution of daily streamflows are described with 3000 ensemble members.

4) Cross-validation and verification methods

To test the out-of-sample predictability of the model, we performed leave-one-year-out cross validation by dropping one year (validation year) at a time from 37 (34) years with complete records in the period 1978–2014, and fitting the best BHM using the remaining (training) years for all the gauges (except Garudeshwar). The fitted model is used to predict the dropped year.

To assess the performance of the proposed BHM, we considered deterministic metrics—i.e., the coefficient of efficiency (CE; Briffa et al. 1988), the Pearson correlation coefficient (R), and the relative bias (BIAS)—and probabilistic verification measures, including the continuous ranked probability skill score (CRPSS), rank histogram, reliability diagram, and discrimination diagram. A detailed description of the verification methods is provided in the supplementary material.

In this work, we compute accuracy metrics and rank histograms for different types of predicted events via sample stratification (Bellier et al. 2017). Therefore, we defined the event categories as (strata) “low flow,” “midrange flow,” and “high flow” simulated events, as days when the ensemble mean simulated streamflow is below the observed median (Q50th), between 50th and 80th observed quantiles (Q50th and Q80th) and above the 80th observed quantile (Q80th), respectively. With these categories, we seek to answer the question: how did our postprocessing method perform, given that we simulated a low-/high-flow event?

Reliability and discrimination diagrams were generated for three thresholds obtained from quantiles Q30th, Q50th, and Q80th of streamflow observations at each gauge (see Table S4). Time series of observed daily streamflow at five stations showed that most high-flow events during 1978–2014 are above the upper threshold (Q80th), while low-flow events are between 0 and the lower threshold (Q30th, see Figs. S3 and S4). This is because heavy rainfalls yield significant flows during the monsoon season, but the rest of the days are characterized by low or midrange flows.

4. Results

a. Selection of the best BHM

LOOIC values from candidate BHMs, for which the significance of the regression coefficients (β and ϕ) was checked (i.e., that posterior PDFs do not contain zero in the 95% credible interval), can be found in Table S5. The best BHM (model 11 in Table S5), considered a gamma model for the marginal distribution, daily VIC streamflow ( QVIC,t1(i+1)) and the nonlinear covariate ( NLt(i)) for gauge i as covariates. Note that for models with more than two covariates and that include precipitation, only LOOIC values for the best model are displayed.

In the best BHM model (Fig. 3) daily streamflow at each gauge i, Qt(i), follows a gamma distribution
gamma(Qt(i)|αt(i),λt(i),NLt(i),QVIC,t(i)),i=1,2,,5,
where αt(i)>0 is the shape parameter and λt(i)>0 is the rate parameter for gauge i and day t. These parameters can be expressed as a function of the expected value ( μt(i)) and variance [ (σt(i))2] of Qt(i) (Wilks 2011):
αt(i)=(μt(i))2(σt(i))2,λt(i)=μt(i)(σt(i))2.
Additionally, μt(i) and σt(i) can be expressed as follows:
μt(i)={β1(i)+β2(i)QVIC,t(i)NLt(i)=0β3(i)+β4(i)QVIC,t(i)+β5(i)NLt(i)NLt(i)>0
σt(i)={ϕ1(i)+ϕ2(i)QVIC,t(i)NLt(i)=0ϕ3(i)+ϕ4(i)QVIC,t(i)+ϕ5(i)NLt(i)NLt(i)>0.

Since the nonlinear covariate, NLt(i), is the product between VIC streamflow and precipitation, it was considered in the step restriction of Eqs. (7) and (8).

Figure 4 shows times series of observations, VIC streamflow, and calibrated ensembles of daily streamflow produced with four candidate BHMs for monsoon season 2013, along with scatterplots of BHM ensemble median versus observations for the entire period (1978–2014) at the Handia gauge (i = 3). The BHM model 5 (with Qt(3) as a covariate; Fig. 4a) attenuates the positive bias (i.e., overestimates low flows) but increases the negative bias (i.e., underestimates high flows), being unable to correct the VIC time delay. Although model 8 (with Qt(3) and P2d,t(3) as covariates; Fig. 4b) attenuates the positive bias and can correct the time delay in peak flows, it also increases the negative bias. The results show that the best model (with Qt(3) and NLt(3) as covariates; Fig. 4c) can fix the time delay in peak flow, attenuate the nonsystematic biases (blue line for 2013, and scatterplot with R = 0.86 for the entire period), and produce ensembles that capture the variability in observed values. Finally, the incorporation of P2d,t(3) as an additional covariate to the best model (Model 13; Fig. 4d) yields negligible improvements (i.e., performance for 2013 and R for the entire period similar to the best model). The same features are observed for the other gauges.

Fig. 4.
Fig. 4.

Calibrated daily streamflow ensembles for monsoon season 2013 at the Handia gauge (i = 3), obtained from BHMs calibrated for the full period considering as covariates (a) QVIC,t(3), (b) P2d,t(3) and QVIC,t(3), (c) QVIC,t(3) and NLt(3) (the best model), and (d) QVIC,t(3),NLt(3), and P2d,t(3). Time series of ensembles are presented as boxplots, and outliers are omitted. The red lines represent observed streamflow, dark yellow represents VIC streamflow, and blue lines show the BHM streamflow ensemble mean. Scatterplots of BHM ensemble means vs observed streamflows for the entire period (1978–2014) are also displayed, along with the 1:1 line and statistically significant R values (p value < 0.05).

Citation: Journal of Hydrometeorology 23, 6; 10.1175/JHM-D-21-0167.1

b. Cross validation

Figure 5 shows cross-validated streamflow ensembles from the best BHM model at the terminal (Garudeshwar), middle (Handia), and headwater (Sandiya) gauges. At Garudeshwar, the BHM can fix the time delay and attenuate the positive bias for low and midrange flows in the ensemble mean (Figs. 5a,b). Although a high correlation is obtained for the entire period (R = 0.78 compared to 0.7 for VIC streamflow) and the ensemble spread is able to capture observations, an increase in the negative bias is observed for high flows compared to VIC streamflow. At Handia (Figs. 5c,d) and Sandiya (Figs. 5e,f), the best BHM model is able to fix the time delay and attenuate the conditional bias in terms of the ensemble mean (blue line and scatterplots for the entire period). Further, R values for Handia and Sandiya are 0.86, and observed values are captured by the ensemble spread at both gauges (Figs. 5d,f). Similar performances are observed for Mandleshwar and Hoshangabad gauges (Fig. S5).

Fig. 5.
Fig. 5.

Scatterplots of cross-validated BHM ensemble mean (best model) vs observed streamflow for 1978–2014, and times series with observed, VIC, and cross-validated BHM ensembles of daily monsoon streamflow for 2013 at (a),(b) Garudeshwar, (c),(d) Handia, and (e),(f) Sandiya gauges. Time series of ensembles are presented as boxplots, and outliers are not displayed. Red lines correspond to observed streamflow, dark yellow to VIC streamflow, and blue lines to BHM ensemble streamflow mean. All R values presented here are statistically significant (p value < 0.05).

Citation: Journal of Hydrometeorology 23, 6; 10.1175/JHM-D-21-0167.1

1) Homoscedasticity of the residuals

Homoscedasticity of the residuals—i.e., constant residual (or error) variance for all magnitude of simulated flows is desired. However, heteroscedasticity is not uncommon for higher flow values. To assess the degree of heteroscedasticity, we examine the relationships between simulated streamflow and the corresponding simulated error (historical flow minus simulated flow from the ensembles, Fig. 6) for calibration and cross-validation periods from the BHM at the five gauges. The ensemble spread of the errors is represented by 50% (gray bands) and 95% (light gray bands) confidence intervals, while the thick blue line is the ensemble mean and the red circles the corresponding values from VIC simulated streamflow. For both calibration and cross validation, the BHM simulated flows exhibit very low heteroscedasticity in terms of the ensemble mean (blue lines oscillating around the null error line) compared to VIC simulated streamflow (red points), for all gauges excepting Garudeshwar, where the BHM ensembles overestimate high flows (Figs. 6a,f). In terms of the ensemble spread, the 50% confidence intervals (gray bands) also exhibit very low heteroscedasticity at all five gauges; however, the 95% confidence intervals (light gray bands) tend to increase with streamflow magnitude. This feature can be explained by the nonstationary variance assumption of the BHM model at all gauges, being also present in other approaches that handle the heteroscedasticity of the residuals (BJP; Zhao et al. 2015).

Fig. 6.
Fig. 6.

Relationships between simulated streamflow and simulated error. Rows are associated with error uncertainty obtained from BHM calibration and cross validation for the entire period (1978–2014), and columns are associated with different gauges. Blue lines correspond to the ensemble mean; gray bands to 50% confidence interval; light gray bands to 95% confidence interval; and the red dots to errors in VIC simulated streamflow.

Citation: Journal of Hydrometeorology 23, 6; 10.1175/JHM-D-21-0167.1

2) Deterministic accuracy metrics

Figure 7 displays boxplots with performance metrics (CE, R, and BIAS) of VIC and BHM models for the entire period (1978–2014) and the three strata (low, midrange, and high simulated events) defined in section 3b(4). For CE (left panels), VIC streamflow was considered as the reference model. For the entire period, BHM outperforms VIC significantly based on all accuracy metrics. Indeed, the distribution of CE (95% interval) and medians are above 0.38 and 0.48, respectively, for all gauges (Fig. 7a). The increase of median R relative to VIC forecast is roughly 0.1 for each gauge (Fig. 7b). Although negative biases are obtained at all gauges, their distributions are close to zero and absolute reductions range from 48% to 78% for the median BIAS, and their distributions do not overlap (Fig. 7c). For R, a nonparametric test (Gibbons and Chakraborti 1992) showed that differences between BHM and VIC medians are statistically significant at a 95% confidence level.

Fig. 7.
Fig. 7.

Boxplots of (left) CE, (center) R, and (right) BIAS at the five stations of the Narmada River basin for (a)–(c) the entire period, (d)–(f) low [ Q^(t)<Q50th], (g)–(i) midrange [ Q50thQ^(t)Q80th], and (j)–(l) high [ Q^(t)>Q80th] flow simulated events strata. For CE (in the left panels), VIC streamflow was considered as the reference model. Each boxplot comprises results from 3000 sample bootstraps. The horizontal bold line in the boxplot is the median, the body of a boxplot shows the interquartile range (Q25thQ75th) and the whiskers represent the sample 95% confidence interval. Differences between the BHM and VIC median for R were significant at a 95% confidence level—based on a two-tailed nonparametric test (Gibbons and Chakraborti 1992)—for the entire period and strata.

Citation: Journal of Hydrometeorology 23, 6; 10.1175/JHM-D-21-0167.1

For the strata associated with simulated low-flow events, considerable improvements over VIC streamflow are observed at all station gauges for CE and BIAS, since the entire CE distribution is above 0.36 (Fig. 7d). The absolute median reduction in BIAS is larger than 50% (Fig. 7f). For all stations but Mandleshwar (Fig. 7e), we obtained a higher R for VIC streamflow, being Garudeshwar the only site where BHM yielded a high reduction in R. These results indicate that BHM could reduce the positive bias in VIC streamflow (Fig. 2) when a low-flow event is simulated. Figures 7g–I show that BHM outperforms VIC in all metrics (except R for Garudeshwar) when a midrange-flow event is simulated, with bias close to zero and CE distributions above 0.6. Finally, BHM outperforms VIC at all gauges for high-flow simulated events strata (Figs. 7j–l): the CE distributions are above 0, median R improvements range 0.22–0.3, and absolute median BIAS reductions reach ∼20% with distributions close to zero, except for Garudeshwar (median BIAS of ∼−20%). These results indicate that when a high-flow event is simulated at each gauge, the postprocessing model provides improvements over raw VIC output in terms of time delay (higher R values) and bias reduction (lower absolute BIAS values).

3) Probabilistic accuracy metrics

Figure 8 shows CRPSS results from BHM model output—using (i) VIC streamflow and (ii) climatology as the reference—for the entire period (1978–2014) and the three strata (low, midrange, and high-flow simulated events strata). Compared to VIC streamflow (Fig. 8a), the proposed BHM framework yields a substantial increase in skill (95% of the distributions above 0.56) at all gauges considering the entire period, low- and midrange-flow events strata, excepting Mandleshwar for low-flow simulated events (95% of the distribution above 0.48). The smallest skill improvements are achieved for the high-flow simulated events strata (95% of the distributions are between 0.35 and 0.65). On the other hand, BHM yields a significant increase in skill (95% of the distribution above 0.35) at all gauges compared to observed climatology (Fig. 8b) considering the entire period, low- and high-flow events strata; even for the low-flow events strata, 95% of the CRPSS distribution at all gauges is above 0.72. The lowest increases in skill are obtained for the midrange-flow events strata (95% of the distributions are between 0.22 and 0.36).

Fig. 8.
Fig. 8.

Boxplots of CRPSS considering (a) VIC simulated streamflow and (b) observed climatology as the reference forecast at the five gauges for the entire period, low [ Q^(t)<Q50th], midrange [ Q50thQ^(t)Q80th], and high [ Q^(t)>Q80th] flow simulated events strata. Each boxplot represents the same as in Fig. 7.

Citation: Journal of Hydrometeorology 23, 6; 10.1175/JHM-D-21-0167.1

4) Reliability

Figure 9 displays rank histograms for cross-validated BHM streamflow ensembles along with discrepancy indices (DI) for each station (columns), the entire period and the three strata (displayed in different rows). For Garudeshwar (first column of Fig. 9), the shape of rank histograms for the entire period, midrange and high-flow simulated events strata reveals a combination of over dispersive (underconfident) and systematic positive bias shapes. Rank histograms of the low-flow simulated events strata shows a slight under dispersive (overconfident) shape. DI values range 33%–41% for the entire period, low- and midrange-flow simulated strata, and increases up to 66% for the high-flow simulated strata. At the remaining sites, most rank histograms show a close to uniform shape for the entire period, and low and midrange-flow simulated events strata, with DI values generally below 18%. An exception is the high-flow simulated strata at all gauges, where rank histograms exhibit a slight undersimulation shape, with DI values below 33%. Additionally, a slight positive bias is observed at Handia and Hoshangabad gauges for the low-flow simulated strata. Similar features are observed in reliability diagrams at the other four stations (Fig. S6).

Fig. 9.
Fig. 9.

Rank histograms for cross-validated BHM daily streamflow ensembles during July–August for the entire period, low [ Q^(t)<Q50th], midrange [ Q50thQ^(t)Q80th], and high [ Q^(t)>Q80th] flow simulated events strata. Horizontal red lines correspond to the median frequency.

Citation: Journal of Hydrometeorology 23, 6; 10.1175/JHM-D-21-0167.1

5) Discrimination

Figure 10 displays discrimination diagrams with BHM results for each station (rows) and three streamflow thresholds (columns). The metrics included in each panel are discrimination distance (d), overlapping area (A), and standard deviations (σ0 and σ1) from the two conditional likelihoods. A good discrimination is given by sharp (i.e., small σ0 and σ1), well separated (large d), and no overlapping (small A) likelihoods. In this case, we omit confidence limits in discrimination diagrams because there were only minor differences in the resampled PDFs. At Garudeshwar (the first column of Fig. 10), conditional likelihoods are well separated (high d), with little overlap and increased PDF sharpness for nonoccurrence events (i.e., decrease of σ0) as the threshold increases. No clear relationships with thresholds are observed for the standard deviation of event occurrence (σ1). For Q30th the shape of nonoccurrence PDF (o = 0) is not skewed to the left, which implies a failure of the BHM system to detect nonoccurrence of low-flow events. The opposite is observed during midrange and high-flow events, for which PDFs shape of occurrence (o = 1, red lines) for Q50th and Q80th are not skewed to the right, showing difficulties to detect midrange and high-flow events.

Fig. 10.
Fig. 10.

Discrimination diagrams for cross-validated BHM streamflows computed over July–August days during the period 1978–2014. Rows are associated with different daily streamflow thresholds, and columns are associated to different gauges. The PDF for cases where o = 0 (no occurrence) is represented by the dark line (diamonds), and the PDF for cases with o = 1 (occurrence) is represented by the red line (circles). The metrics displayed in each panel are discrimination distance (d), overlapping area (A), and standard deviations (σ0 and σ1) from the two conditional likelihoods.

Citation: Journal of Hydrometeorology 23, 6; 10.1175/JHM-D-21-0167.1

At the remaining four stations (columns 2–5 of Fig. 10), the proposed BHM framework provides good discrimination, with a decrease (increase) in PDF sharpness of event occurrence (nonoccurrence) for larger thresholds—i.e., increase of σ1 and decrease of σ0 for larger thresholds. The PDFs shapes of event occurrence (o = 1, red lines) and nonoccurrence (o = 0, black lines) for Q30th and Q50th thresholds are skewed to the right and left, respectively, reflecting success in detecting low- and midrange-flow events. For the Q80th threshold, the PDF of occurrence at three gauges (Mandleshwar, Hoshangabad, and Sandiya; Figs. 10l,n,o) is not skewed to the right, reflecting a failure to assign high probabilities for high-streamflow events.

5. Discussion

The BHM postprocessing framework presented here has several appealing features. First, it can be easily adapted for different types of hydrometeorological variables by selecting suitable marginal distributions (e.g., extreme value, Weibull, binomial, Poisson) at the data layer. Nonlinearities can be incorporated into the process layer by adding nonlinear terms in the regression equations. The approach also enables handling the heteroscedasticity of the residuals, similar to MCP (Todini 2008) and BJP (Zhao et al. 2015). The BHM approach can be implemented with single trace or ensemble forecasts from multiple sources (i.e., models) as potential covariates, enabling multimodel combinations (e.g., Mendoza et al. 2014). Qualitative information about the flows can also be incorporated in the process layers as covariates or conditions of the steps functions. For example, information about reservoirs can be incorporated into the BHM as conditions for the steps functions or covariates.

Initial implementations of the BHM considered other VIC outputs as potential covariates, such as the soil moisture from the three layers. Nevertheless, this variable was finally discarded because complex nonlinear transformations were needed to achieve acceptable skill. Additionally, soil moisture is a local storage within the watershed, while streamflow results from the combination of different hydrological processes at various spatiotemporal scales.

Although previous studies showed that LOOIC performs better than other traditional model selection metrics such as the AIC and DIC (Luo and Al-Harbi 2017), we noticed that the LOOIC metric may not necessarily help to avoid an unnecessary increase of model complexity. For example, a previous implementation of this framework considered to increase the model structure complexity using three-step functions for the mean and variance, with the hypothesis that including a third step for high-flow events in the step functions would handle the sampling error. Then, we selected a BHM model with three-step functions as the best model according to the LOOIC criteria. However, a comprehensive cross-validation performance assessment of both models (BHM implemented here and one with three-step functions) revealed that the improvement in performance obtained by adding a step was marginal to justify an increase in model complexity (6 additional coefficients). Thus, we recommend for future applications of the framework to carefully examine tradeoffs between model complexity and gain in performance—and not be constrained with a single metric.

In this work, we did not compare BHM results against generalized linear models (GLMs) as a reference since they yield lower performance than BHM in terms of probabilistic accuracy metrics (Ossandón et al. 2021). In comparison with postprocessing methods that rely on streamflow persistence (Li et al. 2015, 2016, 2017a; McInerney et al. 2018; Bennett et al. 2021), BHM can reach similar skill improvements without incorporating observed streamflow as a covariate, which is helpful for basins (e.g., Narmada) where discontinuous streamflow records are a problem for model development. A detailed comparison of BHM against other postprocessing methods (Wood and Schaake 2008; Weerts et al. 2011) is out of the scope of this paper, being left for future work.

A key limitation of this study is the exclusion of reservoir operations in the Narmada River basin from hydrologic model simulations, since we did not have reliable information about operation policies. We speculate that such limitation could affect some of the results presented here. For example, although BHM provides a bias reduction and increased skill and correlation relative to VIC at Garudeshwar, it also yields a negative bias for high-flow simulated events strata, poor reliability (combination of over dispersive and positive bias shapes) for the entire period, midrange and high-flow simulated events strata, and poor discrimination. At this gauge, a poor performance of VIC simulated streamflow was obtained during the monsoon season, with large biases likely attributed to reservoir effects that are not captured by VIC. This problem is also observed during July 2013 (Fig. 5b)—when VIC overestimates observed low flows—and other water years (not shown). Future work could explore, besides the incorporation of reservoir operation rules, other parameter estimation strategies, including alternative calibration algorithms (e.g., Gharari et al. 2013; Fenicia et al. 2018) and objective functions (e.g., Shafii and Tolson 2015; Fowler et al. 2018). We expect that improved hydrologic model performance should translate into enhanced skill in postprocessed ensembles.

The proposed framework could be easily implemented in other major flood-affected basins worldwide, with different river network topologies, and using other conceptual or physically based hydrological models, suitable parametric distributions, and covariates. Also, in the case that hydrologic model simulations cannot capture well the spatiotemporal dependence of the observed data, copulas can be included in the BHM for simultaneous multisite postprocessing.

6. Summary and conclusions

We developed a Bayesian hierarchical model (BHM) for the simultaneous postprocessing of daily summer (July–August) streamflow obtained from a physically based model at multiple sites. The framework was tested at five streamflow gauges in central India’s Narmada River Basin network, using outputs from the Variable Infiltration Capacity (VIC) model for the summer season 1978–2014. Daily streamflow at each location was modeled using a Gamma distribution with time-varying parameters, which were modeled as a linear function of VIC streamflow and a nonlinear covariate—the product of VIC streamflow and 2-day spatial average precipitation from the area between two consecutive gauges. This nonlinear covariate could fix VIC streamflow peaks’ conditional bias and time delay. The best model (i.e., gamma distribution and the covariates) was selected using LOOIC—a cross-validation criterion. Finally, the inclusion of suitable priors for the regression coefficients helped to correctly simulate posterior parameter distributions and predictive posterior distribution of streamflows for each day and location.

Deterministic accuracy metrics showed enhanced performance of BHM ensembles compared to raw VIC simulated streamflow—i.e., absolute bias reductions of at least 83% and skill improvements of at least 38% and 46% for CE and CRPSS, respectively, across the basin for the entire period. The BHM skill results were consistently better than VIC streamflow for the entire period and the three strata (low-, midrange-, and high-flow events strata) at all gauges. Further, rank histograms and discrimination diagrams showed that the BHM frameworks provide streamflow ensembles with a good spread and discrimination at most sites, except for the terminal gauge (Garudeshwar). Overall, the results presented here illustrate the enormous potential of this framework for short- to medium-range streamflow forecasting in the Narmada River basin, complementing and enhancing the current operational flood forecasting system (CWC 2015). Additionally, the proposed approach could be implemented and tested in other river basin networks across the globe, contributing to improved flood hazard mitigation.

Acknowledgments.

This research was funded by the Monsoon Mission project of the Ministry of Earth Sciences, India. We also acknowledge the support from the Fulbright Foreign Student Program and the National Agency for Research and Development (ANID) Scholarship Program/DOCTORADO BECAS CHILE/2015-56150013 to the first author. P. A. Mendoza acknowledges support from Fondecyt projects 11200142 and CONICYT/PIA Project AFB180004.

Data availability statement.

The dataset used in this study, which consists of time series of potential covariates and daily monsoon period (July–August) streamflow for four station gauges in the Narmada River basin, can be downloaded from HydroShare https://www.hydroshare.org/resource/e2e4924a3ed2416b831b0e6a0cb0bdaf/.

REFERENCES

  • Akaike, H., 1974: A new look at the statistical model identification. IEEE Trans. Autom. Control, 19, 716723, https://doi.org/10.1109/TAC.1974.1100705.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ali, H., and V. Mishra, 2018: Increase in subdaily precipitation extremes in India under 1.5° and 2.0°C warming worlds. Geophys. Res. Lett., 45, 69726982, https://doi.org/10.1029/2018GL078689.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ali, H., P. Modi, and V. Mishra, 2019: Increased flood risk in Indian sub-continent under the warming climate. Wea. Climate Extremes, 25, 100212, https://doi.org/10.1016/j.wace.2019.100212.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Anele, A. O., E. Todini, Y. Hamam, and A. M. Abu-Mahfouz, 2018: Predictive uncertainty estimation in water demand forecasting using the model conditional processor. Water, 10, 475, https://doi.org/10.3390/w10040475.

    • Crossref
    • Export Citation
  • Banerjee, R., 2009: Review of water governance in the Narmada river basin. Society for Promotion of Wastelands Development, Tech. Rep., 40 pp., http://www.indiaenvironmentportal.org.in/files/Narmada_Basin_Review.pdf.

    • Crossref
    • Export Citation
  • Bellier, J., I. Zin, and G. Bontron, 2017: Sample stratification in verification of ensemble forecasts of continuous scalar variables: Potential benefits and pitfalls. Mon. Wea. Rev., 145, 35293544, https://doi.org/10.1175/MWR-D-16-0487.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Bennett, J. C., Q. J. Wang, D. E. Robertson, R. Bridgart, J. Lerat, M. Li, and K. Michael, 2021: An error model for long-range ensemble forecasts of ephemeral rivers. Adv. Water Resour., 151, 103891, https://doi.org/10.1016/j.advwatres.2021.103891.

    • Crossref
    • Export Citation
  • Bogner, K., and F. Pappenberger, 2011: Multiscale error analysis, correction, and predictive uncertainty estimation in a flood forecasting system. Water Resour. Res., 47, 7524, https://doi.org/10.1029/2010WR009137.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Bohn, T. J., and E. R. Vivoni, 2016: Process-based characterization of evapotranspiration sources over the North American monsoon region. Water Resour. Res., 52, 358384, https://doi.org/10.1002/2015WR017934.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Bremnes, J. B., 2004: Probabilistic forecasts of precipitation in terms of quantiles using NWP model output. Mon. Wea. Rev., 132, 338347, https://doi.org/10.1175/1520-0493(2004)132<0338:PFOPIT>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Briffa, K. R., P. D. Jones, J. R. Pilcher, and M. K. Hughes, 1988: Reconstructing summer temperatures in northern Fennoscandinavia back to AD 1700 using tree-ring data from Scots pine. Arct. Alp. Res., 20, 385394, https://doi.org/10.2307/1551336.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Brown, J. D., and D. J. Seo, 2010: A nonparametric postprocessor for bias correction of hydrometeorological and hydrologic ensemble forecasts. J. Hydrometeor., 11, 642665, https://doi.org/10.1175/2009JHM1188.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Brown, J. D., and D.-J. Seo, 2013: Evaluation of a nonparametric post-processor for bias correction and uncertainty estimation of hydrologic predictions. Hydrol. Processes, 27, 83105, https://doi.org/10.1002/hyp.9263.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Cherkauer, K. A., L. C. Bowling, and D. P. Lettenmaier, 2003: Variable infiltration capacity cold land process model updates. Global Planet. Change, 38, 151159, https://doi.org/10.1016/S0921-8181(03)00025-0.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Clark, M. P., S. Gangopadhyay, L. Hay, B. Rajagopalan, and R. Wilby, 2004: The Schaake Shuffle: A method for reconstructing space–time variability in forecasted precipitation and temperature fields. J. Hydrometeor., 5, 243262, https://doi.org/10.1175/1525-7541(2004)005<0243:TSSAMF>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Clark, M. P., and Coauthors, 2017: The evolution of process-based hydrologic models: Historical challenges and the collective quest for physical realism. Hydrol. Earth Syst. Sci., 21, 34273440, https://doi.org/10.5194/hess-21-3427-2017.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • CWC, 2012: Flood forecasting and warning network performance: Appraisal Report 2011. Central Water Commission, Tech. Rep., 90 pp., http://www.cwc.gov.in/sites/default/files/Final_FFWNPAR20112_For_Printing.pdf.

    • Crossref
    • Export Citation
  • CWC, 2015: Flood forecasting and warning system in India. Proc. Regional Flood Early Warning System Workshop, Bangkok, Thailand, Regional Integrated Multi-Hazard Early Warning System, 78, https://documents1.worldbank.org/curated/en/431281468000591916/pdf/103879-WP-PUBLIC-Rimes-Workshop-Proceedings-13Jan2016.pdf.

    • Crossref
    • Export Citation
  • Day, G. N., 1985: Extended streamflow forecasting using NWSRFS. J. Water Resour. Plann. Manage., 111, 157170, https://doi.org/10.1061/(ASCE)0733-9496(1985)111:2(157).

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Demargne, J., and Coauthors, 2014: The science of NOAA’s operational hydrologic ensemble forecast service. Bull. Amer. Meteor. Soc., 95, 7998, https://doi.org/10.1175/BAMS-D-12-00081.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ercolani, G., and F. Castelli, 2017: Variational assimilation of streamflow data in distributed flood forecasting. Water Resour. Res., 53, 158183, https://doi.org/10.1002/2016WR019208.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • FAO, IIASA, ISRIC, ISSCAS, and JRC, 2012: H Harmonized World Soil Database (version 1.2). FAO and IIASA, http://webarchive.iiasa.ac.at/Research/LUC/External-World-soil-database/HTML/index.html?sb=1.

    • Crossref
    • Export Citation
  • Fenicia, F., D. Kavetski, P. Reichert, and C. Albert, 2018: Signature-domain calibration of hydrological models using approximate Bayesian computation: Empirical analysis of fundamental properties. Water Resour. Res., 54, 39583987, https://doi.org/10.1002/2017WR021616.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Fowler, K., M. Peel, A. Western, and L. Zhang, 2018: Improved rainfall-runoff calibration for drying climate: Choice of objective function. Water Resour. Res., 54, 33923408, https://doi.org/10.1029/2017WR022466.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Franchini, M., and M. Pacciani, 1991: Comparative analysis of several conceptual rainfall-runoff models. J. Hydrol., 122, 161219, https://doi.org/10.1016/0022-1694(91)90178-K.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Gelman, A., and D. B. Rubin, 1992: Inference from iterative simulation using multiple sequences. Stat. Sci., 7, 457472, https://doi.org/10.1214/ss/1177011136.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Gelman, A., and J. Hill, 2006: Data Analysis Using Regression and Multilevel/Hierarchical Models. Cambridge University Press, 625 pp.

    • Crossref
    • Export Citation
  • Gharari, S., M. Hrachowitz, F. Fenicia, and H. H. G. Savenije, 2013: An approach to identify time consistent model parameters: Sub-period calibration. Hydrol. Earth Syst. Sci., 17, 149161, https://doi.org/10.5194/hess-17-149-2013.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ghimire, G. R., and W. F. Krajewski, 2020: Exploring persistence in streamflow forecasting. J. Amer. Water Resour. Assoc., 56, 542550, https://doi.org/10.1111/1752-1688.12821.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Gibbons, J. D., and S. Chakraborti, 1992: Nonparametric Statistical Inference. 4th ed. Marcel Dekker, Inc., 645 pp.

    • Crossref
    • Export Citation
  • Glahn, H. R., and D. A. Lowry, 1972: The Use of Model Output Statistics (MOS) in objective weather forecasting. J. Appl. Meteor., 11, 12031211, https://doi.org/10.1175/1520-0450(1972)011<1203:TUOMOS>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Grillakis, M., A. Koutroulis, and I. Tsanis, 2018: Improving seasonal forecasts for basin scale hydrological applications. Water, 10, 1593, https://doi.org/10.3390/w10111593.

    • Crossref
    • Export Citation
  • Gupta, H. V., H. Kling, K. K. Yilmaz, and G. F. Martinez, 2009: Decomposition of the mean squared error and NSE performance criteria: Implications for improving hydrological modelling. J. Hydrol., 377, 8091, https://doi.org/10.1016/j.jhydrol.2009.08.003.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hansen, M. C., R. Sohlberg, R. S. Defries, and J. R. Townshend, 2000: Global land cover classification at 1 km spatial resolution using a classification tree approach. Int. J. Remote Sens., 21, 13311364, https://doi.org/10.1080/014311600210209.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hashino, T., A. A. Bradley, and S. S. Schwartz, 2007: Evaluation of bias-correction methods for ensemble streamflow volume forecasts. Hydrol. Earth Syst. Sci., 11, 939950, https://doi.org/10.5194/hess-11-939-2007.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hoffman, M. D., and A. Gelman, 2014: The No-U-turn sampler: Adaptively setting path lengths in Hamiltonian Monte Carlo. J. Mach. Learn. Res., 15, 15931623, https://www.jmlr.org/papers/volume15/hoffman14a/hoffman14a.pdf.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hunt, K. M., and J. K. Fletcher, 2019: The relationship between Indian monsoon rainfall and low-pressure systems. Climate Dyn., 53, 18591871, https://doi.org/10.1007/s00382-019-04744-x.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hunt, K. M. R., A. G. Turner, and D. E. Parker, 2016: The spatiotemporal structure of precipitation in Indian monsoon depressions. Quart. J. Roy. Meteor. Soc., 142, 31953210, https://doi.org/10.1002/qj.2901.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Jay-Allemand, M., P. Javelle, I. Gejadze, P. Arnaud, P. O. Malaterre, J. A. Fine, and D. Organde, 2020: On the potential of variational calibration for a fully distributed hydrological model: Application on a Mediterranean catchment. Hydrol. Earth Syst. Sci., 24, 55195538, https://doi.org/10.5194/hess-24-5519-2020.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Kalnay, E., M. Kanamitsu, R. Kistler, W. Collins, D. Deaven, and L. Gandin, 1996: The NCEP/NCAR 40-Year Reanalysis Project. Bull. Amer. Meteor. Soc., 77, 437472, https://doi.org/10.1175/1520-0477(1996)077<0437:TNYRP>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Kistler, R., and Coauthors, 2001: The NCEP–NCAR 50-year reanalysis: Monthly means CD-ROM and documentation. Bull. Amer. Meteor. Soc., 82, 247267, https://doi.org/10.1175/1520-0477(2001)082<0247:TNNYRM>2.3.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Krajewski, W. F., G. R. Ghimire, and F. Quintero, 2020: Streamflow forecasting without models. J. Hydrometeor., 21, 16891704, https://doi.org/10.1175/JHM-D-19-0292.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Krzysztofowicz, R., and K. S. Kelly, 2000: Hydrologic uncertainty processor for probabilistic river stage forecasting. Water Resour. Res., 36, 32653277, https://doi.org/10.1029/2000WR900108.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Li, M., Q. J. Wang, J. C. Bennett, and D. E. Robertson, 2015: A strategy to overcome adverse effects of autoregressive updating of streamflow forecasts. Hydrol. Earth Syst. Sci., 19, 115, https://doi.org/10.5194/hess-19-1-2015.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Li, M., Q. J. Wang, J. C. Bennett, and D. E. Robertson, 2016: Error Reduction and Representation In Stages (ERRIS) in hydrological modelling for ensemble streamflow forecasting. Hydrol. Earth Syst. Sci., 20, 35613579, https://doi.org/10.5194/hess-20-3561-2016.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Li, M., Q. J. Wang, D. E. Robertson, and J. C. Bennett, 2017a: Improved error modelling for streamflow forecasting at hourly time steps by splitting hydrographs into rising and falling limbs. J. Hydrol., 555, 586599, https://doi.org/10.1016/j.jhydrol.2017.10.057.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Li, W., Q. Duan, C. Miao, A. Ye, W. Gong, and Z. Di, 2017b: A Review on Statistical Postprocessing Methods for Hydrometeorological Ensemble Forecasting. John Wiley & Sons Inc., 1246 pp.

    • Crossref
    • Export Citation
  • Liang, X., and Z. Xie, 2001: A new surface runoff parameterization with subgrid-scale soil heterogeneity for land surface models. Adv. Water Resour., 24, 11731193, https://doi.org/10.1016/S0309-1708(01)00032-X.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Liang, X., D. P. Lettenmaier, E. F. Wood, and S. J. Burges, 1994: A simple hydrologically based model of land surface water and energy fluxes for general circulation models. J. Geophys. Res., 99, 14 415–14 429, https://doi.org/10.1029/94JD00483.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Liang, X., E. F. Wood, and D. P. Lettenmaier, 1996: Surface soil moisture parameterization of the VIC-2L model: Evaluation and modification. Global Planet. Change, 13, 195206, https://doi.org/10.1016/0921-8181(95)00046-1.

    • Search Google Scholar
    • Export Citation
  • Liu, Y., and H. V. Gupta, 2007: Uncertainty in Hydrologic Modeling: Toward an Integrated Data Assimilation Framework. John Wiley & Sons, Ltd, 7401 pp.

    • Crossref
    • Export Citation
  • Lohmann, D., R. Nolte-Holube, and E. Raschke, 1996: A large-scale horizontal routing model to be coupled to land surface parametrization schemes. Tellus, 48A, 708721, https://doi.org/10.3402/tellusa.v48i5.12200.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lucatero, D., H. Madsen, J. C. Refsgaard, J. Kidmose, and K. H. Jensen, 2018: Seasonal streamflow forecasts in the Ahlergaarde catchment, Denmark: The effect of preprocessing and post-processing on skill and statistical consistency. Hydrol. Earth Syst. Sci., 22, 36013617, https://doi.org/10.5194/hess-22-3601-2018.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Luo, Y., and K. Al-Harbi, 2017: Performances of LOO and WAIC as IRT model selection methods. Psychol. Test Assess. Model., 59, 183205, https://www.psychologie-aktuell.com/fileadmin/download/ptam/2-2017_20170627/03_Luo_.pdf.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Maurer, E. P., A. W. Wood, J. C. Adam, D. P. Lettenmaier, and B. Nijssen, 2002: A long-term hydrologically based dataset of land surface fluxes and states for the conterminous United States. J. Climate, 15, 32373251, https://doi.org/10.1175/1520-0442(2002)015<3237:ALTHBD>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • McInerney, D., M. Thyer, D. Kavetski, J. Lerat, and G. Kuczera, 2017: Improving probabilistic prediction of daily streamflow by identifying Pareto optimal approaches for modeling heteroscedastic residual errors. Water Resour. Res., 53, 21992239, https://doi.org/10.1002/2016WR019168.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • McInerney, D., M. Thyer, D. Kavetski, B. Bennett, J. Lerat, M. Gibbs, and G. Kuczera, 2018: A simplified approach to produce probabilistic hydrological model predictions. Environ. Modell. Software, 109, 306314, https://doi.org/10.1016/j.envsoft.2018.07.001.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Mendoza, P. A., B. Rajagopalan, M. P. Clark, G. Cortés, and J. McPhee, 2014: A robust multimodel framework for ensemble seasonal hydroclimatic forecasts. Water Resour. Res., 50, 60306052, https://doi.org/10.1002/2014WR015426.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Mendoza, P. A., M. P. Clark, M. Barlage, B. Rajagopalan, L. Samaniego, G. Abramowitz, and H. Gupta, 2015a: Are we unnecessarily constraining the agility of complex process-based models? Water Resour. Res., 51, 716728, https://doi.org/10.1002/2014WR015820.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Mendoza, P. A., B. Rajagopalan, M. P. Clark, K. Ikeda, and R. M. Rasmussen, 2015b: Statistical postprocessing of high-resolution regional climate model output. Mon. Wea. Rev., 143, 15331553, https://doi.org/10.1175/MWR-D-14-00159.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Muhammad, A., T. A. Stadnyk, F. Unduche, and P. Coulibaly, 2018: Multi-model approaches for improving seasonal ensemble streamflow prediction scheme with various statistical post-processing techniques in the Canadian Prairie region. Water, 10, 1604, https://doi.org/10.3390/w10111604.

    • Crossref
    • Export Citation
  • Nash, J. E., and J. V. Sutcliffe, 1970: River flow forecasting through conceptual models Part I – A discussion of principles. J. Hydrol., 10, 282290, https://doi.org/10.1016/0022-1694(70)90255-6.

    • Search Google Scholar
    • Export Citation
  • Ossandón, Á., B. Rajagopalan, U. Lall, J. S. Nanditha, and V. Mishra, 2021: A Bayesian hierarchical network model for daily streamflow ensemble forecasting. Water Resour. Res., 57, e2021WR029920, https://doi.org/10.1029/2021WR029920.

    • Crossref
    • Export Citation
  • Pagano, T. C., and Coauthors, 2014: Challenges of operational river forecasting. J. Hydrometeor., 15, 16921707, https://doi.org/10.1175/JHM-D-13-0188.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Pai, D., L. Sridhar, M. Rajeevan, O. P. Sreejith, N. S. Satbhai, and B. Mukhopadhyay, 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, 1–18, https://doi.org/10.54302/mausam.v65i1.851.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Papalexiou, S. M., and A. Montanari, 2019: Global and regional increase of precipitation extremes under global warming. Water Resour. Res., 55, 49014914, https://doi.org/10.1029/2018WR024067.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Pokhrel, P., D. E. Robertson, and Q. J. Wang, 2013: A Bayesian joint probability post-processor for reducing errors and quantifying uncertainty in monthly streamflow predictions. Hydrol. Earth Syst. Sci., 17, 795804, https://doi.org/10.5194/hess-17-795-2013.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • R Core Team, 2017: R: A language and environment for statistical computing. R Foundation for Statistical Computing, https://www.R-project.org/.

    • Crossref
    • Export Citation
  • Ravindranath, A., N. Devineni, U. Lall, E. R. Cook, G. Pederson, J. Martin, and C. Woodhouse, 2019: Streamflow reconstruction in the upper Missouri River basin using a novel Bayesian network model. Water Resour. Res., 55, 76947716, https://doi.org/10.1029/2019WR024901.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Reggiani, P., M. Renner, A. H. Weerts, and P. A. Van Gelder, 2009: Uncertainty assessment via Bayesian revision of ensemble streamflow predictions in the operational river Rhine forecasting system. Water Resour. Res., 45, W02428, https://doi.org/10.1029/2007WR006758.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Robert, C., and G. Casella, 2011: A short history of Markov chain Monte Carlo: Subjective recollections from incomplete data. Stat. Sci., 26, 102115, https://doi.org/10.1214/10-STS351.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Rodell, M., and Coauthors, 2004: The Global Land Data Assimilation System. Bull. Amer. Meteor. Soc., 85, 381394, https://doi.org/10.1175/BAMS-85-3-381.

    • Search Google Scholar
    • Export Citation
  • Schumann, G. J., J. C. Neal, N. Voisin, K. M. Andreadis, F. Pappenberger, N. Phanthuwongpakdee, A. C. Hall, and P. D. Bates, 2013: A first large-scale flood inundation forecasting model. Water Resour. Res., 49, 62486257, https://doi.org/10.1002/wrcr.20521.

    • Search Google Scholar
    • Export Citation
  • Shafii, M., and B. A. Tolson, 2015: Optimizing hydrological consistency by incorporating hydrological signatures into model calibration objectives. Water Resour. Res., 51, 37963814, https://doi.org/10.1002/2014WR016520.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Shah, R. D., and V. Mishra, 2016: Utility of Global Ensemble Forecast System (GEFS) reforecast for medium-range drought prediction in India. J. Hydrometeor., 17, 17811800, https://doi.org/10.1175/JHM-D-15-0050.1.

    • Search Google Scholar
    • Export Citation
  • Sheffield, J., and E. F. Wood, 2007: Characteristics of global and regional drought, 1950–2000: Analysis of soil moisture data from off-line simulation of the terrestrial hydrologic cycle. J. Geophys. Res., 112, D17115, https://doi.org/10.1029/2006JD008288.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Spiegelhalter, D. J., N. G. Best, B. P. Carlin, and A. van der Linde, 2002: Bayesian measures of model complexity and fit. J. Roy. Stat. Soc., 64, 583639, https://doi.org/10.1111/1467-9868.00353.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Stan Development Team, 2014: Stan Modeling Language User’s Guide and Reference Manual. Stan Development Team, 408 pp.

    • Crossref
    • Export Citation
  • Stan Development Team, 2020: RStan: The R interface to Stan. Stan Development Team, http://mc-stan.org/.

  • Tanoue, M., Y. Hirabayashi, and H. Ikeuchi, 2016: Global-scale river flood vulnerability in the last 50 years. Sci. Rep., 6, 36021, https://doi.org/10.1038/srep36021.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Tiwari, A. D., M. Parthasarathi, and V. Mishra, 2022: Influence of bias correction of meteorological and streamflow forecast on hydrological prediction in India. J. Hydrometeor., https://doi.org/10.1175/JHM-D-20-0235.1, in press.

    • Search Google Scholar
    • Export Citation
  • Todini, E., 2008: A model conditional processor to assess predictive uncertainty in flood forecasting. Int. J. River Basin Manage., 6, 123137, https://doi.org/10.1080/15715124.2008.9635342.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Tyralis, H., G. Papacharalampous, A. Burnetas, and A. Langousis, 2019: Hydrological post-processing using stacked generalization of quantile regression algorithms: Large-scale application over CONUS. J. Hydrol., 577, 123957, https://doi.org/10.1016/j.jhydrol.2019.123957.

    • Crossref
    • Export Citation
  • Vásquez, N., J. Cepeda, T. Gómez, P. A. Mendoza, M. Lagos, J. P. Boisier, C. Álvarez-Garretón, and X. Vargas, 2021: Catchment-scale natural water balance in Chile. Water Resources of Chile, Springer International Publishing, 189–208.

    • Crossref
    • Export Citation
  • Vehtari, A., A. Gelman, and J. Gabry, 2017: Practical Bayesian model evaluation using leave-one-out cross-validation and WAIC. Stat. Comput., 27, 14131432, https://doi.org/10.1007/s11222-016-9696-4.

    • Search Google Scholar
    • Export Citation
  • Vehtari, A., J. Gabry, M. Magnusson, Y. Yao, P.-C. Bürkner, T. Paananen, and A. Gelman, 2020: loo: Efficient leave-one-out cross-validation and WAIC for Bayesian models. R package version 2.4.1, Stan Development Team, https://mc-stan.org/loo/.

    • Crossref
    • Export Citation
  • Verkade, J. S., J. D. Brown, P. Reggiani, and A. H. Weerts, 2013: Post-processing ECMWF precipitation and temperature ensemble reforecasts for operational hydrologic forecasting at various spatial scales. J. Hydrol., 501, 7391, https://doi.org/10.1016/j.jhydrol.2013.07.039.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wallemacq, P., and R. House, 2018: Economic losses, poverty and disasters: 1998–2017. UNISDR and CRED, Tech. Rep., 31 pp., https://reliefweb.int/report/world/economic-losses-poverty-disasters-1998-2017.

    • Crossref
    • Export Citation
  • Wang, Q. J., D. E. Robertson, and F. H. Chiew, 2009: A Bayesian joint probability modeling approach for seasonal forecasting of streamflows at multiple sites. Water Resour. Res., 45, 5407, https://doi.org/10.1029/2008WR007355.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wasko, C., and A. Sharma, 2017: Continuous rainfall generation for a warmer climate using observed temperature sensitivities. J. Hydrol., 544, 575590, https://doi.org/10.1016/j.jhydrol.2016.12.002.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Weerts, A. H., H. C. Winsemius, and J. S. Verkade, 2011: Estimation of predictive hydrological uncertainty using quantile regression: Examples from the National Flood Forecasting System (England and Wales). Hydrol. Earth Syst. Sci., 15, 255265, https://doi.org/10.5194/hess-15-255-2011.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wilks, D. S., 2011: Statistical Methods in the Atmospheric Sciences. 3rd ed. International Geophysics Series, Vol. 100, Academic Press, 704 pp.

    • Crossref
    • Export Citation
  • Woldemeskel, F., D. McInerney, J. Lerat, M. Thyer, D. Kavetski, D. Shin, N. Tuteja, and G. Kuczera, 2018: Evaluating post-processing approaches for monthly and seasonal streamflow forecasts. Hydrol. Earth Syst. Sci., 22, 62576278, https://doi.org/10.5194/hess-22-6257-2018.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wood, A. W., and J. C. Schaake, 2008: Correcting errors in streamflow forecast ensemble mean and spread. J. Hydrometeor., 9, 132148, https://doi.org/10.1175/2007JHM862.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wu, H., R. F. Adler, Y. Tian, G. J. Huffman, H. Li, and J. Wang, 2014: Real-time global flood estimation using satellite-based precipitation and a coupled land surface and routing model. Water Resour. Res., 50, 26932717, https://doi.org/10.1002/2013WR014710.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ye, A., Q. Duan, X. Yuan, E. F. Wood, and J. Schaake, 2014: Hydrologic post-processing of MOPEX streamflow simulations. J. Hydrol., 508, 147156, https://doi.org/10.1016/j.jhydrol.2013.10.055.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ye, A., Q. Duan, J. Schaake, J. Xu, X. Deng, Z. Di, C. Miao, and W. Gong, 2015: Post-processing of ensemble forecasts in low-flow period. Hydrol. Processes, 29, 24382453, https://doi.org/10.1002/hyp.10374.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Zalachori, I., M.-H. Ramos, R. Garçon, T. Mathevet, and J. Gailhard, 2012: Statistical processing of forecasts for hydrological ensemble prediction: A comparative study of different bias correction strategies. Adv. Sci. Res., 8, 135141, https://doi.org/10.5194/asr-8-135-2012.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Zhao, L., Q. Duan, J. Schaake, A. Ye, and J. Xia, 2011: A hydrologic post-processor for ensemble streamflow predictions. Adv. Geosci., 29, 5159, https://doi.org/10.5194/adgeo-29-51-2011.

    • Search Google Scholar
    • Export Citation
  • Zhao, T., Q. J. Wang, J. C. Bennett, D. E. Robertson, Q. Shao, and J. Zhao, 2015: Quantifying predictive uncertainty of streamflow forecasts based on a Bayesian joint probability model. J. Hydrol., 528, 329340, https://doi.org/10.1016/j.jhydrol.2015.06.043.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Zhao, T., A. Schepen, and Q. J. Wang, 2016: Ensemble forecasting of sub-seasonal to seasonal streamflow by a Bayesian joint probability modelling approach. J. Hydrol., 541, 839849, https://doi.org/10.1016/j.jhydrol.2016.07.040.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Zhou, Y., and S. Guo, 2013: Incorporating ecological requirement into multipurpose reservoir operating rule curves for adaptation to climate change. J. Hydrol., 498, 153164, hhttps://doi.org/10.1016/j.jhydrol.2013.06.028.

    • Crossref
    • Search Google Scholar
    • Export Citation

Supplementary Materials

Save
  • Akaike, H., 1974: A new look at the statistical model identification. IEEE Trans. Autom. Control, 19, 716723, https://doi.org/10.1109/TAC.1974.1100705.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ali, H., and V. Mishra, 2018: Increase in subdaily precipitation extremes in India under 1.5° and 2.0°C warming worlds. Geophys. Res. Lett., 45, 69726982, https://doi.org/10.1029/2018GL078689.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ali, H., P. Modi, and V. Mishra, 2019: Increased flood risk in Indian sub-continent under the warming climate. Wea. Climate Extremes, 25, 100212, https://doi.org/10.1016/j.wace.2019.100212.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Anele, A. O., E. Todini, Y. Hamam, and A. M. Abu-Mahfouz, 2018: Predictive uncertainty estimation in water demand forecasting using the model conditional processor. Water, 10, 475, https://doi.org/10.3390/w10040475.

    • Crossref
    • Export Citation
  • Banerjee, R., 2009: Review of water governance in the Narmada river basin. Society for Promotion of Wastelands Development, Tech. Rep., 40 pp., http://www.indiaenvironmentportal.org.in/files/Narmada_Basin_Review.pdf.

    • Crossref
    • Export Citation
  • Bellier, J., I. Zin, and G. Bontron, 2017: Sample stratification in verification of ensemble forecasts of continuous scalar variables: Potential benefits and pitfalls. Mon. Wea. Rev., 145, 35293544, https://doi.org/10.1175/MWR-D-16-0487.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Bennett, J. C., Q. J. Wang, D. E. Robertson, R. Bridgart, J. Lerat, M. Li, and K. Michael, 2021: An error model for long-range ensemble forecasts of ephemeral rivers. Adv. Water Resour., 151, 103891, https://doi.org/10.1016/j.advwatres.2021.103891.

    • Crossref
    • Export Citation
  • Bogner, K., and F. Pappenberger, 2011: Multiscale error analysis, correction, and predictive uncertainty estimation in a flood forecasting system. Water Resour. Res., 47, 7524, https://doi.org/10.1029/2010WR009137.

    • Search Google Scholar
    • Export Citation
  • Bohn, T. J., and E. R. Vivoni, 2016: Process-based characterization of evapotranspiration sources over the North American monsoon region. Water Resour. Res., 52, 358384, https://doi.org/10.1002/2015WR017934.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Bremnes, J. B., 2004: Probabilistic forecasts of precipitation in terms of quantiles using NWP model output. Mon. Wea. Rev., 132, 338347, https://doi.org/10.1175/1520-0493(2004)132<0338:PFOPIT>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Briffa, K. R., P. D. Jones, J. R. Pilcher, and M. K. Hughes, 1988: Reconstructing summer temperatures in northern Fennoscandinavia back to AD 1700 using tree-ring data from Scots pine. Arct. Alp. Res., 20, 385394, https://doi.org/10.2307/1551336.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Brown, J. D., and D. J. Seo, 2010: A nonparametric postprocessor for bias correction of hydrometeorological and hydrologic ensemble forecasts. J. Hydrometeor., 11, 642665, https://doi.org/10.1175/2009JHM1188.1.

    • Search Google Scholar
    • Export Citation
  • Brown, J. D., and D.-J. Seo, 2013: Evaluation of a nonparametric post-processor for bias correction and uncertainty estimation of hydrologic predictions. Hydrol. Processes, 27, 83105, https://doi.org/10.1002/hyp.9263.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Cherkauer, K. A., L. C. Bowling, and D. P. Lettenmaier, 2003: Variable infiltration capacity cold land process model updates. Global Planet. Change, 38, 151159, https://doi.org/10.1016/S0921-8181(03)00025-0.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Clark, M. P., S. Gangopadhyay, L. Hay, B. Rajagopalan, and R. Wilby, 2004: The Schaake Shuffle: A method for reconstructing space–time variability in forecasted precipitation and temperature fields. J. Hydrometeor., 5, 243262, https://doi.org/10.1175/1525-7541(2004)005<0243:TSSAMF>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Clark, M. P., and Coauthors, 2017: The evolution of process-based hydrologic models: Historical challenges and the collective quest for physical realism. Hydrol. Earth Syst. Sci., 21, 34273440, https://doi.org/10.5194/hess-21-3427-2017.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • CWC, 2012: Flood forecasting and warning network performance: Appraisal Report 2011. Central Water Commission, Tech. Rep., 90 pp., http://www.cwc.gov.in/sites/default/files/Final_FFWNPAR20112_For_Printing.pdf.

    • Crossref
    • Export Citation
  • CWC, 2015: Flood forecasting and warning system in India. Proc. Regional Flood Early Warning System Workshop, Bangkok, Thailand, Regional Integrated Multi-Hazard Early Warning System, 78, https://documents1.worldbank.org/curated/en/431281468000591916/pdf/103879-WP-PUBLIC-Rimes-Workshop-Proceedings-13Jan2016.pdf.

    • Crossref
    • Export Citation
  • Day, G. N., 1985: Extended streamflow forecasting using NWSRFS. J. Water Resour. Plann. Manage., 111, 157170, https://doi.org/10.1061/(ASCE)0733-9496(1985)111:2(157).

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Demargne, J., and Coauthors, 2014: The science of NOAA’s operational hydrologic ensemble forecast service. Bull. Amer. Meteor. Soc., 95, 7998, https://doi.org/10.1175/BAMS-D-12-00081.1.

    • Search Google Scholar
    • Export Citation
  • Ercolani, G., and F. Castelli, 2017: Variational assimilation of streamflow data in distributed flood forecasting. Water Resour. Res., 53, 158183, https://doi.org/10.1002/2016WR019208.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • FAO, IIASA, ISRIC, ISSCAS, and JRC, 2012: H Harmonized World Soil Database (version 1.2). FAO and IIASA, http://webarchive.iiasa.ac.at/Research/LUC/External-World-soil-database/HTML/index.html?sb=1.

    • Crossref
    • Export Citation
  • Fenicia, F., D. Kavetski, P. Reichert, and C. Albert, 2018: Signature-domain calibration of hydrological models using approximate Bayesian computation: Empirical analysis of fundamental properties. Water Resour. Res., 54, 39583987, https://doi.org/10.1002/2017WR021616.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Fowler, K., M. Peel, A. Western, and L. Zhang, 2018: Improved rainfall-runoff calibration for drying climate: Choice of objective function. Water Resour. Res., 54, 33923408, https://doi.org/10.1029/2017WR022466.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Franchini, M., and M. Pacciani, 1991: Comparative analysis of several conceptual rainfall-runoff models. J. Hydrol., 122, 161219, https://doi.org/10.1016/0022-1694(91)90178-K.

    • Search Google Scholar
    • Export Citation
  • Gelman, A., and D. B. Rubin, 1992: Inference from iterative simulation using multiple sequences. Stat. Sci., 7, 457472, https://doi.org/10.1214/ss/1177011136.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Gelman, A., and J. Hill, 2006: Data Analysis Using Regression and Multilevel/Hierarchical Models. Cambridge University Press, 625 pp.

    • Crossref
    • Export Citation
  • Gharari, S., M. Hrachowitz, F. Fenicia, and H. H. G. Savenije, 2013: An approach to identify time consistent model parameters: Sub-period calibration. Hydrol. Earth Syst. Sci., 17, 149161, https://doi.org/10.5194/hess-17-149-2013.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ghimire, G. R., and W. F. Krajewski, 2020: Exploring persistence in streamflow forecasting. J. Amer. Water Resour. Assoc., 56, 542550, https://doi.org/10.1111/1752-1688.12821.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Gibbons, J. D., and S. Chakraborti, 1992: Nonparametric Statistical Inference. 4th ed. Marcel Dekker, Inc., 645 pp.

    • Crossref
    • Export Citation
  • Glahn, H. R., and D. A. Lowry, 1972: The Use of Model Output Statistics (MOS) in objective weather forecasting. J. Appl. Meteor., 11, 12031211, https://doi.org/10.1175/1520-0450(1972)011<1203:TUOMOS>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Grillakis, M., A. Koutroulis, and I. Tsanis, 2018: Improving seasonal forecasts for basin scale hydrological applications. Water, 10, 1593, https://doi.org/10.3390/w10111593.

    • Crossref
    • Export Citation
  • Gupta, H. V., H. Kling, K. K. Yilmaz, and G. F. Martinez, 2009: Decomposition of the mean squared error and NSE performance criteria: Implications for improving hydrological modelling. J. Hydrol., 377, 8091, https://doi.org/10.1016/j.jhydrol.2009.08.003.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hansen, M. C., R. Sohlberg, R. S. Defries, and J. R. Townshend, 2000: Global land cover classification at 1 km spatial resolution using a classification tree approach. Int. J. Remote Sens., 21, 13311364, https://doi.org/10.1080/014311600210209.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hashino, T., A. A. Bradley, and S. S. Schwartz, 2007: Evaluation of bias-correction methods for ensemble streamflow volume forecasts. Hydrol. Earth Syst. Sci., 11, 939950, https://doi.org/10.5194/hess-11-939-2007.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hoffman, M. D., and A. Gelman, 2014: The No-U-turn sampler: Adaptively setting path lengths in Hamiltonian Monte Carlo. J. Mach. Learn. Res., 15, 15931623, https://www.jmlr.org/papers/volume15/hoffman14a/hoffman14a.pdf.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hunt, K. M., and J. K. Fletcher, 2019: The relationship between Indian monsoon rainfall and low-pressure systems. Climate Dyn., 53, 18591871, https://doi.org/10.1007/s00382-019-04744-x.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hunt, K. M. R., A. G. Turner, and D. E. Parker, 2016: The spatiotemporal structure of precipitation in Indian monsoon depressions. Quart. J. Roy. Meteor. Soc., 142, 31953210, https://doi.org/10.1002/qj.2901.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Jay-Allemand, M., P. Javelle, I. Gejadze, P. Arnaud, P. O. Malaterre, J. A. Fine, and D. Organde, 2020: On the potential of variational calibration for a fully distributed hydrological model: Application on a Mediterranean catchment. Hydrol. Earth Syst. Sci., 24, 55195538, https://doi.org/10.5194/hess-24-5519-2020.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Kalnay, E., M. Kanamitsu, R. Kistler, W. Collins, D. Deaven, and L. Gandin, 1996: The NCEP/NCAR 40-Year Reanalysis Project. Bull. Amer. Meteor. Soc., 77, 437472, https://doi.org/10.1175/1520-0477(1996)077<0437:TNYRP>2.0.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Kistler, R., and Coauthors, 2001: The NCEP–NCAR 50-year reanalysis: Monthly means CD-ROM and documentation. Bull. Amer. Meteor. Soc., 82, 247267, https://doi.org/10.1175/1520-0477(2001)082<0247:TNNYRM>2.3.CO;2.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Krajewski, W. F., G. R. Ghimire, and F. Quintero, 2020: Streamflow forecasting without models. J. Hydrometeor., 21, 16891704, https://doi.org/10.1175/JHM-D-19-0292.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Krzysztofowicz, R., and K. S. Kelly, 2000: Hydrologic uncertainty processor for probabilistic river stage forecasting. Water Resour. Res., 36, 32653277, https://doi.org/10.1029/2000WR900108.

    • Search Google Scholar
    • Export Citation
  • Li, M., Q. J. Wang, J. C. Bennett, and D. E. Robertson, 2015: A strategy to overcome adverse effects of autoregressive updating of streamflow forecasts. Hydrol. Earth Syst. Sci., 19, 115, https://doi.org/10.5194/hess-19-1-2015.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Li, M., Q. J. Wang, J. C. Bennett, and D. E. Robertson, 2016: Error Reduction and Representation In Stages (ERRIS) in hydrological modelling for ensemble streamflow forecasting. Hydrol. Earth Syst. Sci., 20, 35613579, https://doi.org/10.5194/hess-20-3561-2016.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Li, M., Q. J. Wang, D. E. Robertson, and J. C. Bennett, 2017a: Improved error modelling for streamflow forecasting at hourly time steps by splitting hydrographs into rising and falling limbs. J. Hydrol., 555, 586599, https://doi.org/10.1016/j.jhydrol.2017.10.057.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Li, W., Q. Duan, C. Miao, A. Ye, W. Gong, and Z. Di, 2017b: A Review on Statistical Postprocessing Methods for Hydrometeorological Ensemble Forecasting. John Wiley & Sons Inc., 1246 pp.

    • Crossref
    • Export Citation
  • Liang, X., and Z. Xie, 2001: A new surface runoff parameterization with subgrid-scale soil heterogeneity for land surface models. Adv. Water Resour., 24, 11731193, https://doi.org/10.1016/S0309-1708(01)00032-X.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Liang, X., D. P. Lettenmaier, E. F. Wood, and S. J. Burges, 1994: A simple hydrologically based model of land surface water and energy fluxes for general circulation models. J. Geophys. Res., 99, 14 415–14 429, https://doi.org/10.1029/94JD00483.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Liang, X., E. F. Wood, and D. P. Lettenmaier, 1996: Surface soil moisture parameterization of the VIC-2L model: Evaluation and modification. Global Planet. Change, 13, 195206, https://doi.org/10.1016/0921-8181(95)00046-1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Liu, Y., and H. V. Gupta, 2007: Uncertainty in Hydrologic Modeling: Toward an Integrated Data Assimilation Framework. John Wiley & Sons, Ltd, 7401 pp.

    • Crossref
    • Export Citation
  • Lohmann, D., R. Nolte-Holube, and E. Raschke, 1996: A large-scale horizontal routing model to be coupled to land surface parametrization schemes. Tellus, 48A, 708721, https://doi.org/10.3402/tellusa.v48i5.12200.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lucatero, D., H. Madsen, J. C. Refsgaard, J. Kidmose, and K. H. Jensen, 2018: Seasonal streamflow forecasts in the Ahlergaarde catchment, Denmark: The effect of preprocessing and post-processing on skill and statistical consistency. Hydrol. Earth Syst. Sci., 22, 36013617, https://doi.org/10.5194/hess-22-3601-2018.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Luo, Y., and K. Al-Harbi, 2017: Performances of LOO and WAIC as IRT model selection methods. Psychol. Test Assess. Model., 59, 183205, https://www.psychologie-aktuell.com/fileadmin/download/ptam/2-2017_20170627/03_Luo_.pdf.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Maurer, E. P., A. W. Wood, J. C. Adam, D. P. Lettenmaier, and B. Nijssen, 2002: A long-term hydrologically based dataset of land surface fluxes and states for the conterminous United States. J. Climate, 15, 32373251, https://doi.org/10.1175/1520-0442(2002)015<3237:ALTHBD>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • McInerney, D., M. Thyer, D. Kavetski, J. Lerat, and G. Kuczera, 2017: Improving probabilistic prediction of daily streamflow by identifying Pareto optimal approaches for modeling heteroscedastic residual errors. Water Resour. Res., 53, 21992239, https://doi.org/10.1002/2016WR019168.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • McInerney, D., M. Thyer, D. Kavetski, B. Bennett, J. Lerat, M. Gibbs, and G. Kuczera, 2018: A simplified approach to produce probabilistic hydrological model predictions. Environ. Modell. Software, 109, 306314, https://doi.org/10.1016/j.envsoft.2018.07.001.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Mendoza, P. A., B. Rajagopalan, M. P. Clark, G. Cortés, and J. McPhee, 2014: A robust multimodel framework for ensemble seasonal hydroclimatic forecasts. Water Resour. Res., 50, 60306052, https://doi.org/10.1002/2014WR015426.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Mendoza, P. A., M. P. Clark, M. Barlage, B. Rajagopalan, L. Samaniego, G. Abramowitz, and H. Gupta, 2015a: Are we unnecessarily constraining the agility of complex process-based models? Water Resour. Res., 51, 716728, https://doi.org/10.1002/2014WR015820.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Mendoza, P. A., B. Rajagopalan, M. P. Clark, K. Ikeda, and R. M. Rasmussen, 2015b: Statistical postprocessing of high-resolution regional climate model output. Mon. Wea. Rev., 143, 15331553, https://doi.org/10.1175/MWR-D-14-00159.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Muhammad, A., T. A. Stadnyk, F. Unduche, and P. Coulibaly, 2018: Multi-model approaches for improving seasonal ensemble streamflow prediction scheme with various statistical post-processing techniques in the Canadian Prairie region. Water, 10, 1604, https://doi.org/10.3390/w10111604.

    • Crossref
    • Export Citation
  • Nash, J. E., and J. V. Sutcliffe, 1970: River flow forecasting through conceptual models Part I – A discussion of principles. J. Hydrol., 10, 282290, https://doi.org/10.1016/0022-1694(70)90255-6.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ossandón, Á., B. Rajagopalan, U. Lall, J. S. Nanditha, and V. Mishra, 2021: A Bayesian hierarchical network model for daily streamflow ensemble forecasting. Water Resour. Res., 57, e2021WR029920, https://doi.org/10.1029/2021WR029920.

    • Crossref
    • Export Citation
  • Pagano, T. C., and Coauthors, 2014: Challenges of operational river forecasting. J. Hydrometeor., 15, 16921707, https://doi.org/10.1175/JHM-D-13-0188.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Pai, D., L. Sridhar, M. Rajeevan, O. P. Sreejith, N. S. Satbhai, and B. Mukhopadhyay, 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, 1–18, https://doi.org/10.54302/mausam.v65i1.851.

    • Search Google Scholar
    • Export Citation
  • Papalexiou, S. M., and A. Montanari, 2019: Global and regional increase of precipitation extremes under global warming. Water Resour. Res., 55, 49014914, https://doi.org/10.1029/2018WR024067.

    • Search Google Scholar
    • Export Citation
  • Pokhrel, P., D. E. Robertson, and Q. J. Wang, 2013: A Bayesian joint probability post-processor for reducing errors and quantifying uncertainty in monthly streamflow predictions. Hydrol. Earth Syst. Sci., 17, 795804, https://doi.org/10.5194/hess-17-795-2013.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • R Core Team, 2017: R: A language and environment for statistical computing. R Foundation for Statistical Computing, https://www.R-project.org/.

  • Ravindranath, A., N. Devineni, U. Lall, E. R. Cook, G. Pederson, J. Martin, and C. Woodhouse, 2019: Streamflow reconstruction in the upper Missouri River basin using a novel Bayesian network model. Water Resour. Res., 55, 76947716, https://doi.org/10.1029/2019WR024901.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Reggiani, P., M. Renner, A. H. Weerts, and P. A. Van Gelder, 2009: Uncertainty assessment via Bayesian revision of ensemble streamflow predictions in the operational river Rhine forecasting system. Water Resour. Res., 45, W02428, https://doi.org/10.1029/2007WR006758.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Robert, C., and G. Casella, 2011: A short history of Markov chain Monte Carlo: Subjective recollections from incomplete data. Stat. Sci., 26, 102115, https://doi.org/10.1214/10-STS351.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Rodell, M., and Coauthors, 2004: The Global Land Data Assimilation System. Bull. Amer. Meteor. Soc., 85, 381394, https://doi.org/10.1175/BAMS-85-3-381.

    • Search Google Scholar
    • Export Citation
  • Schumann, G. J., J. C. Neal, N. Voisin, K. M. Andreadis, F. Pappenberger, N. Phanthuwongpakdee, A. C. Hall, and P. D. Bates, 2013: A first large-scale flood inundation forecasting model. Water Resour. Res., 49, 62486257, https://doi.org/10.1002/wrcr.20521.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Shafii, M., and B. A. Tolson, 2015: Optimizing hydrological consistency by incorporating hydrological signatures into model calibration objectives. Water Resour. Res., 51, 37963814, https://doi.org/10.1002/2014WR016520.

    • Search Google Scholar
    • Export Citation
  • Shah, R. D., and V. Mishra, 2016: Utility of Global Ensemble Forecast System (GEFS) reforecast for medium-range drought prediction in India. J. Hydrometeor., 17, 17811800, https://doi.org/10.1175/JHM-D-15-0050.1.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Sheffield, J., and E. F. Wood, 2007: Characteristics of global and regional drought, 1950–2000: Analysis of soil moisture data from off-line simulation of the terrestrial hydrologic cycle. J. Geophys. Res., 112, D17115, https://doi.org/10.1029/2006JD008288.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Spiegelhalter, D. J., N. G. Best, B. P. Carlin, and A. van der Linde, 2002: Bayesian measures of model complexity and fit. J. Roy. Stat. Soc., 64, 583639, https://doi.org/10.1111/1467-9868.00353.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Stan Development Team, 2014: Stan Modeling Language User’s Guide and Reference Manual. Stan Development Team, 408 pp.

  • Stan Development Team, 2020: RStan: The R interface to Stan. Stan Development Team, http://mc-stan.org/.

    • Crossref
    • Export Citation
  • Tanoue, M., Y. Hirabayashi, and H. Ikeuchi, 2016: Global-scale river flood vulnerability in the last 50 years. Sci. Rep., 6, 36021, https://doi.org/10.1038/srep36021.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Tiwari, A. D., M. Parthasarathi, and V. Mishra, 2022: Influence of bias correction of meteorological and streamflow forecast on hydrological prediction in India. J. Hydrometeor., https://doi.org/10.1175/JHM-D-20-0235.1, in press.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Todini, E., 2008: A model conditional processor to assess predictive uncertainty in flood forecasting. Int. J. River Basin Manage., 6, 123137, https://doi.org/10.1080/15715124.2008.9635342.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Tyralis, H., G. Papacharalampous, A. Burnetas, and A. Langousis, 2019: Hydrological post-processing using stacked generalization of quantile regression algorithms: Large-scale application over CONUS. J. Hydrol., 577, 123957, https://doi.org/10.1016/j.jhydrol.2019.123957.

    • Crossref
    • Export Citation
  • Vásquez, N., J. Cepeda, T. Gómez, P. A. Mendoza, M. Lagos, J. P. Boisier, C. Álvarez-Garretón, and X. Vargas, 2021: Catchment-scale natural water balance in Chile. Water Resources of Chile, Springer International Publishing, 189–208.

    • Crossref
    • Export Citation
  • Vehtari, A., A. Gelman, and J. Gabry, 2017: Practical Bayesian model evaluation using leave-one-out cross-validation and WAIC. Stat. Comput., 27, 14131432, https://doi.org/10.1007/s11222-016-9696-4.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Vehtari, A., J. Gabry, M. Magnusson, Y. Yao, P.-C. Bürkner, T. Paananen, and A. Gelman, 2020: loo: Efficient leave-one-out cross-validation and WAIC for Bayesian models. R package version 2.4.1, Stan Development Team, https://mc-stan.org/loo/.

    • Crossref
    • Export Citation
  • Verkade, J. S., J. D. Brown, P. Reggiani, and A. H. Weerts, 2013: Post-processing ECMWF precipitation and temperature ensemble reforecasts for operational hydrologic forecasting at various spatial scales. J. Hydrol., 501, 7391, https://doi.org/10.1016/j.jhydrol.2013.07.039.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wallemacq, P., and R. House, 2018: Economic losses, poverty and disasters: 1998–2017. UNISDR and CRED, Tech. Rep., 31 pp., https://reliefweb.int/report/world/economic-losses-poverty-disasters-1998-2017.

    • Crossref
    • Export Citation
  • Wang, Q. J., D. E. Robertson, and F. H. Chiew, 2009: A Bayesian joint probability modeling approach for seasonal forecasting of streamflows at multiple sites. Water Resour. Res., 45, 5407, https://doi.org/10.1029/2008WR007355.

    • Search Google Scholar
    • Export Citation
  • Wasko, C., and A. Sharma, 2017: Continuous rainfall generation for a warmer climate using observed temperature sensitivities. J. Hydrol., 544, 575590, https://doi.org/10.1016/j.jhydrol.2016.12.002.

    • Search Google Scholar
    • Export Citation
  • Weerts, A. H., H. C. Winsemius, and J. S. Verkade, 2011: Estimation of predictive hydrological uncertainty using quantile regression: Examples from the National Flood Forecasting System (England and Wales). Hydrol. Earth Syst. Sci., 15, 255265, https://doi.org/10.5194/hess-15-255-2011.

    • Search Google Scholar
    • Export Citation
  • Wilks, D. S., 2011: Statistical Methods in the Atmospheric Sciences. 3rd ed. International Geophysics Series, Vol. 100, Academic Press, 704 pp.

  • Woldemeskel, F., D. McInerney, J. Lerat, M. Thyer, D. Kavetski, D. Shin, N. Tuteja, and G. Kuczera, 2018: Evaluating post-processing approaches for monthly and seasonal streamflow forecasts. Hydrol. Earth Syst. Sci., 22, 62576278, https://doi.org/10.5194/hess-22-6257-2018.

    • Search Google Scholar
    • Export Citation
  • Wood, A. W., and J. C. Schaake, 2008: Correcting errors in streamflow forecast ensemble mean and spread. J. Hydrometeor., 9, 132148, https://doi.org/10.1175/2007JHM862.1.

    • Search Google Scholar
    • Export Citation
  • Wu, H., R. F. Adler, Y. Tian, G. J. Huffman, H. Li, and J. Wang, 2014: Real-time global flood estimation using satellite-based precipitation and a coupled land surface and routing model. Water Resour. Res., 50, 26932717, https://doi.org/10.1002/2013WR014710.

    • Search Google Scholar
    • Export Citation
  • Ye, A., Q. Duan, X. Yuan, E. F. Wood, and J. Schaake, 2014: Hydrologic post-processing of MOPEX streamflow simulations. J. Hydrol., 508, 147156, https://doi.org/10.1016/j.jhydrol.2013.10.055.

    • Search Google Scholar
    • Export Citation
  • Ye, A., Q. Duan, J. Schaake, J. Xu, X. Deng, Z. Di, C. Miao, and W. Gong, 2015: Post-processing of ensemble forecasts in low-flow period. Hydrol. Processes, 29, 24382453, https://doi.org/10.1002/hyp.10374.

    • Search Google Scholar
    • Export Citation
  • Zalachori, I., M.-H. Ramos, R. Garçon, T. Mathevet, and J. Gailhard, 2012: Statistical processing of forecasts for hydrological ensemble prediction: A comparative study of different bias correction strategies. Adv. Sci. Res., 8, 135141, https://doi.org/10.5194/asr-8-135-2012.

    • Search Google Scholar
    • Export Citation
  • Zhao, L., Q. Duan, J. Schaake, A. Ye, and J. Xia, 2011: A hydrologic post-processor for ensemble streamflow predictions. Adv. Geosci., 29, 5159, https://doi.org/10.5194/adgeo-29-51-2011.

    • Search Google Scholar
    • Export Citation
  • Zhao, T., Q. J. Wang, J. C. Bennett, D. E. Robertson, Q. Shao, and J. Zhao, 2015: Quantifying predictive uncertainty of streamflow forecasts based on a Bayesian joint probability model. J. Hydrol., 528, 329340, https://doi.org/10.1016/j.jhydrol.2015.06.043.

    • Search Google Scholar
    • Export Citation
  • Zhao, T., A. Schepen, and Q. J. Wang, 2016: Ensemble forecasting of sub-seasonal to seasonal streamflow by a Bayesian joint probability modelling approach. J. Hydrol., 541, 839849, https://doi.org/10.1016/j.jhydrol.2016.07.040.

    • Search Google Scholar
    • Export Citation
  • Zhou, Y., and S. Guo, 2013: Incorporating ecological requirement into multipurpose reservoir operating rule curves for adaptation to climate change. J. Hydrol., 498, 153164, hhttps://doi.org/10.1016/j.jhydrol.2013.06.028.

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

    Map of the Narmada River basin in India, showing 200-m elevation bands, the locations of five subbasin outlets (Sandiya, Hoshangabad, Handia, Mandleshwar, and Garudeshwar), and some of the major dams in the basin: Bargi, Tawa, Indirasagar, Jobat, and Sardar Sarovar (from upstream to downstream direction). The gray circle represents an artificial gauge created for the postprocessing model.

  • Fig. 2.

    (a) Scatterplot with VIC streamflow vs observations at the Handia gauge during monsoon season for the period 1978–2014 and (b) time series for 2013 (high flow year). Light blue ellipses in (b) show examples of conditional bias and time delay of VIC streamflow simulations. The R value is statistically significant (p value < 0.05).

  • Fig. 3.

    Conceptual diagram of the network Bayesian model for the Narmada River basin; Qt(i) is the observed streamflow at gauge i and day t (predictand), NLt(i) is the product between daily VIC streamflow at gauge i and 2-day spatial average precipitation from the area between stations i and i + 1 at day t, and QVIC,t(i) to the daily VIC streamflow for gauge i and day t.

  • Fig. 4.

    Calibrated daily streamflow ensembles for monsoon season 2013 at the Handia gauge (i = 3), obtained from BHMs calibrated for the full period considering as covariates (a) QVIC,t(3), (b) P2d,t(3) and QVIC,t(3), (c) QVIC,t(3) and NLt(3) (the best model), and (d) QVIC,t(3),NLt(3), and P2d,t(3). Time series of ensembles are presented as boxplots, and outliers are omitted. The red lines represent observed streamflow, dark yellow represents VIC streamflow, and blue lines show the BHM streamflow ensemble mean. Scatterplots of BHM ensemble means vs observed streamflows for the entire period (1978–2014) are also displayed, along with the 1:1 line and statistically significant R values (p value < 0.05).

  • Fig. 5.

    Scatterplots of cross-validated BHM ensemble mean (best model) vs observed streamflow for 1978–2014, and times series with observed, VIC, and cross-validated BHM ensembles of daily monsoon streamflow for 2013 at (a),(b) Garudeshwar, (c),(d) Handia, and (e),(f) Sandiya gauges. Time series of ensembles are presented as boxplots, and outliers are not displayed. Red lines correspond to observed streamflow, dark yellow to VIC streamflow, and blue lines to BHM ensemble streamflow mean. All R values presented here are statistically significant (p value < 0.05).

  • Fig. 6.

    Relationships between simulated streamflow and simulated error. Rows are associated with error uncertainty obtained from BHM calibration and cross validation for the entire period (1978–2014), and columns are associated with different gauges. Blue lines correspond to the ensemble mean; gray bands to 50% confidence interval; light gray bands to 95% confidence interval; and the red dots to errors in VIC simulated streamflow.

  • Fig. 7.

    Boxplots of (left) CE, (center) R, and (right) BIAS at the five stations of the Narmada River basin for (a)–(c) the entire period, (d)–(f) low [ Q^(t)<Q50th], (g)–(i) midrange [ Q50thQ^(t)Q80th], and (j)–(l) high [ Q^(t)>Q80th] flow simulated events strata. For CE (in the left panels), VIC streamflow was considered as the reference model. Each boxplot comprises results from 3000 sample bootstraps. The horizontal bold line in the boxplot is the median, the body of a boxplot shows the interquartile range (Q25thQ75th) and the whiskers represent the sample 95% confidence interval. Differences between the BHM and VIC median for R were significant at a 95% confidence level—based on a two-tailed nonparametric test (Gibbons and Chakraborti 1992)—for the entire period and strata.

  • Fig. 8.

    Boxplots of CRPSS considering (a) VIC simulated streamflow and (b) observed climatology as the reference forecast at the five gauges for the entire period, low [ Q^(t)<Q50th], midrange [ Q50thQ^(t)Q80th], and high [ Q^(t)>Q80th] flow simulated events strata. Each boxplot represents the same as in Fig. 7.

  • Fig. 9.

    Rank histograms for cross-validated BHM daily streamflow ensembles during July–August for the entire period, low [ Q^(t)<Q50th], midrange [ Q50thQ^(t)Q80th], and high [ Q^(t)>Q80th] flow simulated events strata. Horizontal red lines correspond to the median frequency.

  • Fig. 10.

    Discrimination diagrams for cross-validated BHM streamflows computed over July–August days during the period 1978–2014. Rows are associated with different daily streamflow thresholds, and columns are associated to different gauges. The PDF for cases where o = 0 (no occurrence) is represented by the dark line (diamonds), and the PDF for cases with o = 1 (occurrence) is represented by the red line (circles). The metrics displayed in each panel are discrimination distance (d), overlapping area (A), and standard deviations (σ0 and σ1) from the two conditional likelihoods.

All Time Past Year Past 30 Days
Abstract Views 672 0 0
Full Text Views 2481 1711 362
PDF Downloads 575 183 21