Hostname: page-component-54dcc4c588-rz4zl Total loading time: 0 Render date: 2025-09-29T11:31:06.458Z Has data issue: false hasContentIssue false

Precipitation prediction over the upper Indus Basin from large-scale circulation patterns using Gaussian processes

Published online by Cambridge University Press:  23 September 2025

Kenza Tazi*
Affiliation:
Department of Engineering, https://ror.org/013meh722 University of Cambridge , Cambridge, UK https://ror.org/01rhff309 British Antarctic Survey , Cambridge, UK
Andrew Orr
Affiliation:
https://ror.org/01rhff309 British Antarctic Survey , Cambridge, UK
J. Scott Hosking
Affiliation:
https://ror.org/01rhff309 British Antarctic Survey , Cambridge, UK https://ror.org/035dkdb55 The Alan Turing Institute , London, UK
Richard E. Turner
Affiliation:
Department of Engineering, https://ror.org/013meh722 University of Cambridge , Cambridge, UK https://ror.org/035dkdb55 The Alan Turing Institute , London, UK
*
Corresponding author: Kenza Tazi; Email: kt484@cam.ac.uk

Abstract

Water resources from the Indus Basin sustain over 270 million people. However, water security in this region is threatened by climate change. This is especially the case for the upper Indus Basin, where most frozen water reserves are expected to decrease significantly by the end of the century, leaving rainfall as the main driver of river flow. However, future precipitation estimates from global climate models differ greatly for this region. To address this uncertainty, this paper explores the feasibility of using probabilistic machine learning to map large-scale circulation fields, better represented by global climate models, to local precipitation over the upper Indus Basin. More specifically, Gaussian processes are trained to predict monthly ERA5 precipitation data over a 15-year horizon. This paper also explores different Gaussian process model designs, including a non-stationary covariance function to learn complex spatial relationships in the data. Going forward, this approach could be used to make more accurate predictions from global climate model outputs and better assess the probability of future precipitation extremes.

Information

Type
Application Paper
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

Impact Statement

Future precipitation in the upper Indus Basin and, by extension, the water security of downstream communities remains difficult to predict. New approaches quantifying the probabilities of extreme precipitation events, which lead to floods or droughts, are needed to inform more sustainable policy and climate change adaptation. This paper presents one such method.

1. Introduction

The Indus is one of the longest rivers in Asia. From the Tibetan Plateau, it meanders through northern India and along the entire length of Pakistan before flowing into the Arabian Sea. Over its 3.000 km journey shown in Figure 1, it sustains the livelihoods of more than 270 million people, including the world’s largest contiguous irrigation system (Shrestha et al., Reference Shrestha, Wagle, Rajbhandari, Khan and Adams2019). Of all the rivers originating in the Hindu-Kush Karakoram Himalaya (HKKH) region, the Indus depends the most on high-altitude water resources (Immerzeel et al., Reference Immerzeel, Van Beek and Bierkens2010; Lutz et al., Reference Lutz, Immerzeel, Shrestha and Bierkens2014). Over 60% of its total flow is attributed to glacier and snow melt in an area known as the Upper Indus Basin (UIB) (Immerzeel and Bierkens, Reference Immerzeel and Bierkens2012; Lutz et al., Reference Lutz, Immerzeel, Kraaijenbrink, Shrestha and Bierkens2016). The Indus River is therefore one of the most sensitive to climate change. However, there is considerable uncertainty around the future amount, seasonality, and composition of the Indus’ flow (Krishnan et al., Reference Krishnan, Shrestha, Ren, Rajbhandari, Saeed, Sanjay, Syed, Vellore, Xu, You and Ren2019; Lutz et al., Reference Lutz, Immerzeel, Kraaijenbrink, Shrestha and Bierkens2016; Orr et al., Reference Orr, Ahmad, Alam, Appadurai, Bharucha, Biemans, Bolch, Chaulagain, Dhaubanjar, Dimri, Dixon, Fowler, Gioli, Halvorson, Hussain, Jeelani, Kamal, Khalid, Liu, Lutz, Mehra, Miles, Momblanch, Muccione, Mukherji, Mustafa, Najmuddin, Nasimi, Nüsser, Pandey, Parveen, Pellicciotti, Pollino, Potter, Qazizada, Ray, Romshoo, Sarkar, Sawas, Sen, Shah, Shah, Shea, Sheikh, Shrestha, Tayal, Tigala, Virk, Wester and Wescoat2022; Sabin et al., Reference Sabin, Krishnan, Vellore, Priya, Borgaonkar, Singh, Sagar, Krishnan, Sanjay, Gnanaseelan, Mujumdar, Kulkarni and Chakraborty2020). This includes predictions of extremes such as floods and droughts. Coupled with an increasing demand for freshwater, future water security in this area is at risk. This security is predominantly undermined by the river’s largest source of current and future hydrological uncertainty: precipitation (Immerzeel et al., Reference Immerzeel, Lutz, Andrade, Bahl, Biemans, Bolch, Hyde, Brumby, Davies, Elmore, Emmer, Feng, Fernández, Haritashya, Kargel, Koppes, Kraaijenbrink, Kulkarni, Mayewski, Nepal, Pacheco, Painter, Pellicciotti, Rajaram, Rupper, Sinisalo, Shrestha, Viviroli, Wada, Xiao, Yao and Baillie2020; Orr et al., Reference Orr, Ahmad, Alam, Appadurai, Bharucha, Biemans, Bolch, Chaulagain, Dhaubanjar, Dimri, Dixon, Fowler, Gioli, Halvorson, Hussain, Jeelani, Kamal, Khalid, Liu, Lutz, Mehra, Miles, Momblanch, Muccione, Mukherji, Mustafa, Najmuddin, Nasimi, Nüsser, Pandey, Parveen, Pellicciotti, Pollino, Potter, Qazizada, Ray, Romshoo, Sarkar, Sawas, Sen, Shah, Shah, Shea, Sheikh, Shrestha, Tayal, Tigala, Virk, Wester and Wescoat2022; Smolenaars et al., Reference Smolenaars, Lutz, Biemans, Dhaubanjar, Immerzeel and Ludwig2021).

Figure 1. Map of Indus Basin showing elevation (TROPOMI, 2019), the Indus River (Kelso and Patterson, Reference Kelso and Patterson2010, in blue) and the UIB (NASA JPL, 2013, watershed boundary shown with dashed line). The inset shows the watershed’s location with respect to the Asian continent.

In the UIB, precipitation largely originates from two major atmospheric events: the Indian summer monsoon (ISM) and the western disturbances (WDs). The ISM brings precipitation from June to September through the displacement of humid air from the Indian Ocean, and in particular the Bay of Bengal (Bookhagen and Burbank, Reference Bookhagen and Burbank2010). The strength of the ISM is primarily modulated by the El Niño-Southern Oscillation (Ding and Wang, Reference Ding and Wang2005; Di Capua et al., Reference Di Capua, Kretschmer, Donner, Van Den Hurk, Vellore, Krishnan and Coumou2020). WDs are extratropical storms that originate over the Mediterranean region. They travel as frontal systems, collecting moisture from the Mediterranean, Black, and Caspian seas until they intensify upon reaching South Asia (Midhuna et al., Reference Midhuna, Kumar and Dimri2020). They are primarily modulated by the North Atlantic Oscillation and El Niño-Southern Oscillation (Hunt and Zaz, Reference Hunt and Zaz2023; Bharati et al., Reference Bharati, Deb, Hunt, Orr and Dash2025) and are strongest in winter from December to April, peaking in March over the UIB (Filippi et al., Reference Filippi, Palazzi, von Hardenberg and Provenzale2014; Dahri et al., Reference Dahri, Ludwig, Moors, Ahmad, Khan and Kabat2016; Bharati et al., Reference Bharati, Deb, Hunt, Orr and Dash2025). Consequently, the eastern UIB receives up to 70% of its annual precipitation in the summer, whereas the western UIB receives 40–60% of its precipitation during winter (Dahri et al., Reference Dahri, Ludwig, Moors, Ahmad, Khan and Kabat2016; Immerzeel et al., Reference Immerzeel, Wanders, Lutz, Shea and Bierkens2015). The complex orography of the UIB also strongly influences precipitation, resulting in large localized spatial–temporal variations, including extremes (Bookhagen and Burbank, Reference Bookhagen and Burbank2006; Lutz et al., Reference Lutz, Immerzeel, Kraaijenbrink, Shrestha and Bierkens2016).

Global climate models (GCMs) often struggle to accurately simulate synoptic and local precipitation patterns (Stephens et al., Reference Stephens, L’Ecuyer, Forbes, Gettelmen, Golaz, Bodas-Salcedo, Suzuki, Gabriel and Haynes2010; Rocheta et al., Reference Rocheta, Sugiyanto, Johnson, Evans and Sharma2014). Their coarse spatial resolution and limited physics cannot resolve complex hydro-meteorological dynamics, including over the HKKH region (Palazzi et al., Reference Palazzi, Von Hardenberg and Provenzale2013; Panday et al., Reference Panday, Thibeault and Frey2015). To overcome these issues, regional climate models (RCMs) are used to dynamically downscale GCM outputs. These models better represent precipitation variability, including extremes. However, RCMs assume that large-scale driving fields, such as temperature and geopotential, are reliably simulated (Chokkavarapu and Mandla, Reference Chokkavarapu and Mandla2019). RCMs are also computationally expensive, lack error estimates, propagate considerable model-dependent uncertainty (Hawkins and Sutton, Reference Hawkins and Sutton2009), and are generally not optimized for mountainous regions, especially the HKKH (Cannon et al., Reference Cannon, Carvalho, Jones, Hoell, Norris, Kiladis and Tahir2017; Norris et al., Reference Norris, Carvalho, Jones, Cannon, Bookhagen, Palazzi and Tahir2017, Reference Norris, Carvalho, Jones and Cannon2019; Orr et al., Reference Orr, Listowski, Couttet, Collier, Immerzeel, Deb and Bannister2017). For example, RCMs from the Coordinated Regional Climate Downscaling Experiment (CORDEX) for South Asia regularly report seasonal errors of over 100% for present-day precipitation when compared to gauge-based gridded precipitation (Sanjay et al., Reference Sanjay, Krishnan, Shrestha, Rajbhandari and Ren2017). On the other hand, large-scale circulation variables, such as horizontal wind, temperature, and pressure fields, are typically better simulated by GCMs (Tapiador et al., Reference Tapiador, Roca, Del Genio, Dewitte, Petersen and Zhang2019; Arias et al., Reference Arias, Bellouin, Coppola, Jones, Krinner, Marotzke, Naik, Palmer, Plattner, Rogelj, Rojas, Sillmann, Storelvmo, Trewin, Rao, Adhikary, Allan, Armour, Bala and Zickfeld2023). Furthermore, these large-scale features have previously been used to statistically downscale monthly precipitation and seasonal precipitation forecasts over the Tian Shan and Pamir mountains in Central Asia using support vector regression (Umirbekov et al., Reference Umirbekov, Peña-Guerrero and Müller2022) and to downscale monthly cumulative and daily extreme precipitation over High Mountain Asia using neural networks (Gerlitz et al., Reference Gerlitz, Conrad and Böhner2015).

This paper explores the feasibility of using variables better represented by GCMs to predict future precipitation over the UIB. For this study, monthly fifth-generation ECMWF reanalysis (ERA5) precipitation data (Hersbach et al., Reference Hersbach, Bell, Berrisford, Hirahara, Horányi, Muñoz-Sabater, Nicolas, Peubey, Radu, Schepers, Simmons, Soci, Abdalla, Abellan, Balsamo, Bechtold, Biavati, Bidlot, Bonavita, de Chiara, Dahlgren, Dee, Diamantakis, Dragani, Flemming, Forbes, Fuentes, Geer, Haimberger, Healy, Hogan, Hólm, Janisková, Keeley, Laloyaux, Lopez, Lupu, Radnoti, de Rosnay, Rozum, Vamborg, Villaume and Thépaut2020) is used to train a machine learning model for temporal prediction over a 15-year horizon. The following experiments serve as a proof-of-concept for applying this approach to predict future precipitation from GCMs. If ERA5 precipitation is assumed to be close to reality, this application can be considered a proof-of-concept for heterogeneous model output statistics downscaling and bias-correction (Maraun and Widmann, Reference Maraun and Widmann2017). If ERA5 precipitation is assumed to be different from reality, this application can simply be considered an emulation of ERA5.

A large variety of methods are applicable to this task. Traditional methods such as multiple linear regression or distribution shift methods are more interpretable (Schoof and Pryor, Reference Schoof and Pryor2001; Gudmundsson et al., Reference Gudmundsson, Bremnes, Haugen and Engen-Skaugen2012; Maraun and Widmann, Reference Maraun and Widmann2017). However, scalable machine learning methods, such as convolutional neural networks, recurrent neural networks, or transformers (Wang et al., Reference Wang, Tian, Lowe, Kalin and Lehrter2021; Zhong et al., Reference Zhong, Du, Chen, Wang and Li2024), and in particular probabilistic machine learning methods, such as generative adversarial networks (Harris et al., Reference Harris, McRae, Chantry, Dueben and Palmer2022; Kumar et al., Reference Kumar, Atey, Singh, Chattopadhyay, Acharya, Singh, Nanjundiah and Rao2023; Leinonen et al., Reference Leinonen, Nerini and Berne2021) or diffusion models (Ling et al., Reference Ling, Lu, Luo, Bai, Behera, Jin, Pan, Jiang and Yamagata2024; Srivastava et al., Reference Srivastava, Yang, Kerrigan, Dresdner, McGibbon, Bretherton and Mandt2024; Mardani et al., Reference Mardani, Brenowitz, Cohen, Pathak, Chen, Liu, Vahdat, Nabian, Ge, Subramaniam, Kashinath, Kautz and Pritchard2025), have become strong choices due to their accuracy in modeling complex spatiotemporal distributions.

In this work, we choose to instead explore Gaussian processes (GP) (Rasmussen and Williams, Reference Rasmussen and Williams2005). This Bayesian machine learning method retains a high level of interpretability while providing a mechanism for incorporating domain knowledge into the model and returning uncertainty estimates through predictive probability distributions of future precipitation. These distributions can then be used to quantify the probability of extreme precipitation events. Different GP covariance functions are explored, including a non-stationary Gibbs kernel, building on work from Lalchand et al. (Reference Lalchand, Tazi, Cheema, Turner and Hosking2022). The paper is structured as follows. The data are first presented in Section 2 and the methodology in Section 3. Model performance is presented in Section 4. Finally, the study’s limitations and further work are discussed in Section 5.

2. ERA5 precipitation and predictive variables

Previous research has found that ERA5 climate reanalysis, although generally exhibiting a wet bias for the HKKH region (Baudouin et al., Reference Baudouin, Herzog and Petrie2020; Chen et al., Reference Chen, Sharma, Zhou, Yang, Li, Niu, Hu and Khadka2021; Kumar et al., Reference Kumar, Hodnebrog, Daloz, Sen, Badiger and Krishnaswamy2021), offers precipitation estimates close to observations, in terms of amounts, seasonality, and variability, from a daily to multi-annual scale, compared to other reanalysis and RCM products over the UIB (Baudouin et al., Reference Baudouin, Herzog and Petrie2020; Dahri et al., Reference Dahri, Ludwig, Moors, Ahmad, Ahmad, Ahmad, Riaz and Kabat2021; Khanal et al., Reference Khanal, Tiwari, Lutz, Hurk and Immerzeel2023). In this work, monthly averages are chosen to learn the UIB’s climatology rather than using daily averages that capture its more stochastic weather patterns. Smoother monthly data also avoids dealing with consecutive days of zero precipitation, which would require the more involved modeling of a Bernoulli-gamma distribution (Martinez-Villalobos and Neelin, Reference Martinez-Villalobos and Neelin2019; Girona-Mata et al., Reference Girona-Mata, Orr, Widmann, Bannister, Dars, Hosking, Norris, Ocio, Phillips, Steiner and Richard2024).

Figure 2 shows ERA5 monthly precipitation data over the UIB from 1970 to 2004. The precipitation has an approximately log-normal distribution with a mean of $ 1.83\;\mathrm{mm}\;{\mathrm{d}}^{-1} $ , a standard deviation of $ 1.84\;\mathrm{mm}\;{\mathrm{d}}^{-1} $ and values ranging from 0.00 to $ 23.40\;\mathrm{mm}\;{\mathrm{d}}^{-1} $ . Soft $ k $ -means clustering is applied to the precipitation data using different time intervals and cluster numbers, excluding latitude and longitude. Three distinct precipitation regimes, shown in Figure 3, are identified and subsequently referred to as the Gilgit, Ngari, and Khyber regimes. Figure 2 shows that the Ngari regime is characterized by a clear periodic pattern, while the Gilgit and Khyber regimes have more stochastic time series (Figure 2).

Figure 2. ERA5 monthly precipitation over the UIB between 1970 and 2004. The average annual and seasonal precipitation over the UIB (left) and time series for three locations in the UIB (right) are shown. The time series shows the typical precipitation regimes in the basin (Figure 3) and are fit with a linear regression model (dotted green lines) to identify trends over the studied time period.

Figure 3. UIB precipitation regimes. Clusters are generated by applying soft $ k $ -means to ERA5 monthly precipitation data from 1970 to 2004. The figure shows the clusters for a weighting threshold of 0.6, following the approach of Lalchand et al. (Reference Lalchand, Tazi, Cheema, Turner and Hosking2022).

Potential variables for predicting future precipitation are compiled from existing literature. They are summarized in Table 1. Large-scale geopotential fields are first considered. The strength and variability of the ISM and WDs over the UIB have previously been linked to large-scale geopotential fields outside the UIB via empirical orthogonal function (EOF) analysis (Filippi et al., Reference Filippi, Palazzi, von Hardenberg and Provenzale2014; Mölg et al., Reference Mölg, Maussion and Scherer2014). EOF analysis, also known as geographically weighted principal component analysis, is the decomposition of a signal or dataset in terms of its orthogonal basis functions. EOFs reduce the dimensionality and noise of large-scale fields and are therefore commonly used to determine atmospheric patterns (Weare and Nasstrom, Reference Weare and Nasstrom1982). Here, monthly EOFs are generated from daily ERA5 geopotentials at 200 hPa, 500 hPa, and 850 hPa over the UIB. Spatially averaged EOFs over the Bay of Bengal (10–16°N, 83–93°E) and east of the Caspian Sea (35–40°N, 60–70°E) are also calculated to capture large-scale pressure systems entering the basin.

In addition to geopotential fields, monthly historical climate indices from NOAA (2011) are investigated to train the model. Climate indices are numerical measures used to describe specific patterns or anomalies in the Earth’s atmosphere and oceans. These indices are typically derived from observational data. The NAO, Niño 3.4, and Niño 4 indices, which provide a representation of the North Atlantic and El Niño-Southern oscillations, are considered. The NAO index is calculated by performing a rotated principal component analysis (RPCA) to the geopotential anomaly at 500 hPa (National Weather Service Climate Prediction Center, 2008). The Niño 3.4 and 4 indices measure anomalous sea surface temperatures over overlapping areas in the tropical Pacific Ocean. Finally, ERA5 surface-level variables are studied. These include static fields such as latitude, longitude, altitude, slope of sub-grid orography, and angle of sub-grid orography as well as local meteorological variables such as 2 m dew point temperature and total column water vapour. Although the latter two variables are not large-scale fields, they can be more faithfully modelled by GCMs. This is a result of the variables being more directly related to the thermodynamic properties of the atmosphere and less influenced by the complex microphysical processes that affect precipitation, especially over mountains (Houze, Reference Houze2012; Tapiador et al., Reference Tapiador, Roca, Del Genio, Dewitte, Petersen and Zhang2019). The monthly 2 m temperature averaged over 20–40°N 65–85°E is also investigated as a proxy for seasonality with annual periodicity similar to precipitation.

3. Method

This section first introduces GPs and methods to address non-stationary and non-Gaussian data. We then present the experimental setup, including the validation protocol and the specific models used.

3.1. Gaussian processes

Consider the data set $ {\left\{{y}_i\right\}}_{i=1}^N $ representing $ N $ monthly ERA5 precipitation values where $ {y}_i\in \mathrm{\mathbb{R}} $ . The values are given at inputs $ {\left\{{\boldsymbol{x}}_i\right\}}_{i=1}^N\in {\mathrm{\mathbb{R}}}^D $ , where $ D $ is the number of input dimensions (e.g., time, latitude, longitude, or elevation). We assume $ {y}_i $ can be generated by a function $ f $ at $ {\boldsymbol{x}}_i $ :

(1) $$ {y}_i=f\left({\boldsymbol{x}}_i\right)+{\varepsilon}_i, $$

where $ {\varepsilon}_i $ is a noise term and is assumed to be normally distributed with zero mean and standard deviation $ {\sigma}_n $ , $ {\varepsilon}_i\sim \mathcal{N}\left(0;{\sigma}_n^2\right) $ .

The function $ f $ can be modeled with a GP. We refer the reader to Rasmussen and Williams (Reference Rasmussen and Williams2005) for an introduction to GPs and follow their notation. A GP is a stochastic process for which any finite collection of its random variables is distributed according to a multivariate normal distribution. Similarly to a multivariate normal distribution, a GP is defined by a mean function $ \mu \left(\boldsymbol{x}\right) $ and a covariance function or kernel $ k\left(\boldsymbol{x},{\boldsymbol{x}}^{\prime}\right) $ :

(2) $$ f\left(\boldsymbol{x}\right)\sim GP\left(\mu \left(\boldsymbol{x}\right),k\Big(\boldsymbol{x},{\boldsymbol{x}}^{\prime}\Big)\right), $$

where $ \boldsymbol{x} $ and $ {\boldsymbol{x}}^{\prime } $ are two arbitrary inputs. Practically, the mean function $ \mu \left(\boldsymbol{x}\right) $ is used to model behaviour common to all data, such as linear trends. The covariance function $ k\left(\boldsymbol{x},{\boldsymbol{x}}^{\prime}\right) $ captures the correlation between the outputs and encodes properties such as smoothness and periodicity. The mean and covariance function can therefore be used to sample representative functions from the GP. The spread of these functions is interpreted as the uncertainty of the model.

The covariance function makes GPs well-suited for highly correlated geophysical datasets. If the covariance function is stationary, the correlation of the outputs depends only on the distance between the inputs $ \boldsymbol{x} $ and $ {\boldsymbol{x}}^{\prime } $ . The Gibbs kernel (Reference Gibbs1998) introduces non-stationarity through input-dependent covariance lengthscales in each dimension. This flexibility should allow the function samples to adapt to the varying precipitation patterns across the UIB. Instead of a single point estimate of the lengthscale $ {\mathrm{\ell}}_d $ per dimension $ d $ , we can consider $ {\mathrm{\ell}}_d\left(\boldsymbol{x}\right) $ , a lengthscale function. The Gibbs kernel for multi-dimensional inputs is given by,

(3) $$ {k}_{\mathrm{Gibbs}}\left(\boldsymbol{x},{\boldsymbol{x}}^{\prime}\right)=\prod \limits_{d=1}^D\sqrt{\frac{2{\mathrm{\ell}}_d\left(\boldsymbol{x}\right){\mathrm{\ell}}_d\left({\boldsymbol{x}}^{\prime}\right)}{{\mathrm{\ell}}_d^2\left(\boldsymbol{x}\right)+{\mathrm{\ell}}_d^2\left({\boldsymbol{x}}^{\prime}\right)}}\exp \left\{-\sum \limits_{d=1}^D\frac{{\left({x}_d-{x}_d^{\prime}\right)}^2}{{\mathrm{\ell}}_d^2\left(\boldsymbol{x}\right)+{\mathrm{\ell}}_d^2\left({\boldsymbol{x}}^{\prime}\right)}\right\}. $$

We model these lengthscale functions non-parametrically with latent GPs, building on work from Lalchand et al. (Reference Lalchand, Tazi, Cheema, Turner and Hosking2022). In this paper, our aim is to design a covariance function using knowledge of precipitation and large-scale circulation patterns and enhance it by learning non-stationary spatial relationships between the data.

The distribution of GP is, by definition, Gaussian. Making the distribution of the target variable (i.e., precipitation) more normal can therefore help with inference. Warped GPs (Snelson et al., Reference Snelson, Ghahramani and Rasmussen2003) allow us to handle non-Gaussian distributions by transforming the GP inputs and outputs. For precipitation in the UIB, a Box–Cox function is used as the warping function:

(4) $$ {y}_i\left(\lambda \right)=\left\{\begin{array}{ll}\frac{y_i^{\lambda }-1}{\lambda },& \mathrm{if}\hskip1em \lambda \ne 0\\ {}\log {y}_i,& \mathrm{if}\hskip1em \lambda =0\end{array}\right. $$

where $ {y}_i $ is the $ i $ th observation and is assumed to be positive and $ \lambda $ is the scaling factor found through maximum likelihood estimation. We note that the inverse transformation of the predicted mean yields the median of the resulting distribution.

3.2. Experimental setup

The design and validation of the following models adhere to the guidelines proposed by Tazi et al. (Reference Tazi, Lin, Viljoen, Gardner, John, Ge and Turner2023). To train and evaluate the GPs, ERA5 monthly precipitation data are divided chronologically. The training set represents 35 years between 1970 and 2004. The model is evaluated on the next 15 years of precipitation data between 2005 and 2019. Ideally, the model would be trained on all the data in the UIB. However, to overcome the high computational complexity of GPs for large datasets, the data are sampled. Unless otherwise stated, the training set contains 10,000 samples from the training period, while the validation and test sets are each comprised of 3500 samples from the validation period. The sampling for each set is random, stratified by location, and without replacement. The training set represents 6% of the total available data over the training period. Performance is evaluated by computing the root mean square error (RMSE), the coefficient of determination $ {R}^2 $ , and the mean log loss (MLL) as defined in Appendix A.

For the first model investigated, the temporal distribution of the precipitation is captured by fitting a GP to the time series of each grid point within the UIB. This collection of GPs is spatially independent and is subsequently referred to as the single-location model. A constant mean function is used here and in the following models, as no obvious trends are shared between the data, as shown in Figure 2. For precipitation in the UIB, an approximately annual periodicity is expected. We therefore use an additive kernel structure with local periodicity defined as

(5) $$ {k}_{\mathrm{SLM}}={k}_{\mathrm{Per}}\left({x}_t\right)\cdot {k}_{\mathrm{Mat}32}\left({x}_t\right)+\sum \limits_{d=1}^D{k}_{\mathrm{Mat}32}\left({\boldsymbol{x}}_{\mathrm{other}}\right), $$

where $ {k}_{\mathrm{Per}} $ is the periodic kernel and $ {k}_{\mathrm{Mat}32} $ is the Matérn 3/2 kernel. We use a Matérn kernel rather than a smooth squared exponential kernel to better model discontinuities in time and space. The periodic and Matérn kernels follow their standard definitions as presented by Duvenaud (Reference Duvenaud2014). The other variables, excluding time and spatial information such as elevation, are represented by $ {\boldsymbol{x}}_{\mathrm{other}} $ . The $ {\boldsymbol{x}}_{\mathrm{other}} $ inputs are determined through greedy feature selection as detailed in Appendix B and could include local or large-scale features from Table 1. To establish a baseline, we first train the single-location model with all data at each location for the full training period (full model). A data-limited model is then trained using the previously described training set, approximately 34 points per location.

For the second model, we apply a GP to each cluster shown in Figure 3. The cluster model uses a similar kernel design:

(6) $$ {k}_{\mathrm{STAT}}={k}_{\mathrm{Per}}\left({x}_t\right)\cdot {k}_{\mathrm{Mat}32}\left({x}_t\right)+{k}_{\mathrm{Mat}32-\mathrm{ARD}}\left({\boldsymbol{x}}_{\mathrm{lat},\mathrm{lon}}\right)+\sum \limits_{d=1}^D\;{k}_{\mathrm{Mat}32}\left({\boldsymbol{x}}_{\mathrm{other}\ \mathrm{spatial}}\right), $$

where the subscript ‘STAT’ refers to the stationarity of the kernel, $ {k}_{\mathrm{Mat}32\hbox{-} \mathrm{ARD}} $ is the Matérn 3/2 kernel with automatic relevance determination (MacKay, Reference MacKay1996), $ {\boldsymbol{x}}_{\mathrm{lat},\mathrm{lon}} $ are the latitude and longitude inputs, and $ {\boldsymbol{x}}_{\mathrm{other}\ \mathrm{spatial}} $ are the other inputs, excluding time, latitude, and longitude, but including other spatial variables such as elevation or slope. These inputs are again chosen through greedy feature selection (Appendix B).

Finally, GPs are fit to the entire UIB. Two whole-basin models are applied. One model is defined using the stationary kernel from Equation 6. The other model uses a non-stationary kernel with the Gibbs function described in Equation 3 for the spatial inputs $ {\boldsymbol{x}}_{\mathrm{lat},\mathrm{lon}} $ . The non-stationary kernel is defined as

(7) $$ {k}_{\mathrm{NONSTAT}}={k}_{\mathrm{Per}}\left({x}_t\right)\cdot {k}_{\mathrm{Mat}32}\left({x}_t\right)+{k}_{\mathrm{Gibbs}}\left({\boldsymbol{x}}_{\mathrm{lat},\mathrm{lon}}\right)+\sum \limits_{d=1}^D\;{k}_{\mathrm{Mat}32\hbox{-} \mathrm{ARD}}\left({\boldsymbol{x}}_{\mathrm{other}\ \mathrm{spatial}}\right). $$

4. Results

4.1. Single-location model

The single-location model is created by training a GP at each gridpoint. Following the results shown in Figure B1, the inputs from Table 1 for this model are time, total column water vapour, 2 m dew point temperature, UIB-averaged 2 m temperature, and the first EOF components of the geopotential at 200 and 500 hPa over the UIB. The outcome of the selection is consistent with features that have a high linear correlation with precipitation. Although large-scale circulation is known to influence precipitation, this link is not established for many of the single-location models during the feature selection process.

For the full single-location model, the GPs have an average predictive skill of $ {R}^2 $ = 0.50, $ \mathrm{RMSE}=0.71\;\mathrm{mm}\;{\mathrm{d}}^{-1} $ and MLL = 0.87 for the test set as shown in Table 2. The performance of the full model with respect to location, evaluated on all data between 2005 and 2019 is plotted in Figure 4. The southeast UIB presents the best $ {R}^2 $ scores. In these cases, the GPs faithfully predict most precipitation values but sometimes struggle to match the monsoon highs. Several GPs in the centre and west of the basin capture precipitation patterns with less accuracy. These regions roughly correspond to the boundaries between the clusters shown in Figure 3. In these more difficult cases, the GPs increase their uncertainty bounds accordingly, indicated by a higher MLL. The RMSE values scale with the average annual precipitation (Figure 2). Spatial differences in $ {R}^2 $ scores could suggest that one or more important variables are missing. The differences could also point to an inappropriate kernel design, such as constraining the precipitation to follow a periodic pattern. The data-limited model yields predictive skill of $ {R}^2 $ = − 0.29, $ \mathrm{RMSE}=1.15\;\mathrm{mm}\;{\mathrm{d}}^{-1} $ and MLL = 2.46 over the 15-year horizon (Table 2). However, including information from neighbouring locations and tailoring the input features improves performance, as shown with the cluster model in the next section.

Table 2. Performance of the single-location(SLM), cluster (Khyber, Gilgit, Ngari), and whole-basin models (WBM). The RMSE is given in $ \mathrm{mm}\;{\mathrm{d}}^{-1} $

Figure 4. Test $ {R}^2 $ (left), RMSE (centre), and MLL (right) for the full single-location model across the UIB evaluated on all precipitation data between 2005 and 2019.

4.2. Cluster model

Next, GPs are trained on the precipitation clusters. Input features are selected for each precipitation regime in Figure B2. For the cluster models, 2 m dew point temperature is favoured over total column water vapour for most of the basin in terms of local climatic variables. The cluster models are also able to better leverage relationships between precipitation and large-scale features. The predictive skill of the models is also minimally improved (and even occasionally worsened) when including orographic features such as slope, elevation, or the angle of the sub-grid orography, despite their well-established influence on precipitation (Houze, Reference Houze2012; Dahri et al., Reference Dahri, Ludwig, Moors, Ahmad, Khan and Kabat2016; Wolvin et al., Reference Wolvin, Strong, Rupper and Steenburgh2024).

The features’ contributed variance is then compared in Table C1. The 2 m dew point temperature is a strong predictor for all precipitation regimes, in particular for Ngari model. The total column water vapour and the El Nino 3.4 index contribute strongly to the Khyber model, whereas orography is more important for the Gilgit model. Large-scale features generally contribute more to the Gilgit and Khyber regimes. These results are consistent with known precipitation drivers for the UIB. The Khyber cluster is located at a lower altitude but is more exposed to frontal systems driven by large-scale circulation patterns. On the other hand the Ngari cluster is more sheltered and mainly receives precipitation from the ISM. The Gilgit cluster sits between the two and is characterized by the largest elevation gradients (Wolvin et al., Reference Wolvin, Strong, Rupper and Steenburgh2024).

In this setting, the cluster model benefits from far less data than the full single-location model, but is still able to infer information about the distribution of precipitation. Using an area weighted average for this model, Table 2 shows $ {R}^2 $  = 0.49,  $ \mathrm{RMSE}=0.96\;\mathrm{mm}\;{\mathrm{d}}^{-1} $ and MLL = 1.03 for the test set. The uncertainty calibration for the cluster model therefore improves on the full single-location model. However, the GP distribution medians generally underestimate high precipitation values at both training and test time.

4.3. Whole-basin models

GPs with stationary and non-stationary kernels are then applied to the whole basin. Following the results from Figure B3, the stationary model is trained with time, latitude, longitude, 2 m dew point temperature, UIB-averaged 2 m temperature, the first EOF components of the geopotential at 500 and 850 hPa over the UIB, the averaged first EOF component of the geopotential at 200 hPa over the Bay of Bengal and the averaged first and second EOF components of the geopotential at 500 hPa east of the Caspian Sea. These results are consistent with the trends observed in the previous section, with the stationary whole-basin model selecting the most important features from the cluster models.

For the stationary model, the 2 m dew point temperature and the UIB-averaged 2 m temperature contribute the most to the prediction of precipitation (Table C1). The same features are also used to train the non-stationary GP. The learnt lengthscales for latitude and longitude are presented in Figure 5. The optimized lengthscale values match the annual distribution of precipitation shown in Figure 2 with on average longer lengthscales for longitude and shorter lengthscales for latitude.

Figure 5. Learnt non-stationary kernel lengthscales over the UIB for longitude, latitude and their magnitude.

The whole-basin model with a stationary kernel has a similar $ {R}^2 $ (0.49), a worse RMSE ( $ 1.25\;\mathrm{mm}\;{\mathrm{d}}^{-1} $ ) but a better MLL (0.91) compared to the cluster model (Table 2). On the other hand, the non-stationary model struggles to generalize as well to new times. This more flexible model yields large predictive variances and thus MLL scores, including at training time. This behaviour suggests that this kernel design is not appropriate and may be more suitable for low-dimensional interpolation tasks such as downscaling. More data or simpler lengthscale functions could also improve model accuracy and uncertainty calibration.

5. Discussion

In this paper, we have shown that GPs can generate uncertainty-aware predictions of precipitation with moderate skill. With further analysis, the stochastic nature of these models can help us better capture the climate change signal in the means, extremes, and seasonal cycles. By producing interpretable outputs similar to a perturbed-physics ensemble for a single RCM, GPs set themselves apart from other traditional downscaling and bias-correction methods.

However, this study suffers from three main limitations. First, the application of the cluster and whole-basin models is restricted by the scalability of GPs. For these models, 6% of the training data is used for training. This amount of data requires 60 GB of RAM to train the stationary whole-basin model, which is generally not possible without a high-memory GPU (e.g., NVIDIA A100 or V100 with 80 GB HBM2 memory) to load and optimize the covariance matrix. On a single Intel Xeon Gold 5317 CPU, training the model would take 62 hours. Therefore, methods to select representative data samples (Liu et al., Reference Liu, Motoda and Yu2004) and scale GPs such as products of experts, low-rank approximations, and inducing points (Liu et al., Reference Liu, Ong, Shen and Cai2020; Tazi et al., Reference Tazi, Lin, Viljoen, Gardner, John, Ge and Turner2023) will need to be implemented to keep future inference and hyperparameter learning tractable. Moving to even larger sets (> $ {10}^6 $ data points) would require probabilistic deep learning, such as neural processes (Vaughan et al., Reference Vaughan, Tebbutt, Hosking and Turner2022; Andersson et al., Reference Andersson, Bruinsma, Markou, Requeima, Coca-Castro, Vaughan, Ellis, Lazzara, Jones and Hosking2023) or hybrid deep learning-GP approaches (Putra et al., Reference Putra, Ishikawa and Tatsubori2022; Sun et al., Reference Sun, Jiang, Yang, Xie and Chen2022).

Second, our knowledge of past precipitation in the UIB is incomplete. The limited in situ observations make capturing the true underlying distribution of precipitation and evaluating models challenging. Accessibility and harsh weather conditions mean that most precipitation gauges are placed in relatively low-altitude locations (Dahri et al., Reference Dahri, Ludwig, Moors, Ahmad, Khan and Kabat2016). Alternatives, such as remote sensing, exist, but these tools alone cannot determine precipitation at a fine enough spatiotemporal scale to resolve its distribution accurately (Yin et al., Reference Yin, Zhang, Liu, Colella and Chen2008; Andermann et al., Reference Andermann, Bonnet and Gloaguen2011; Meng et al., Reference Meng, Li, Hao, Wang and Shao2014; Shukla et al., Reference Shukla, Ojha, Singh, Pal and Fu2019).

Third, most features, including large-scale circulation fields and indices, do not provide clear enough signals to improve the GP predictions. This behaviour is due to the features being heavily cross-correlated: only a few variables are needed for efficient inference. Principal component analysis could be a useful tool to improve feature selection, model skill, and the GP’s numerical stability during inference, but at the expense of losing model interpretability. There is also growing evidence that long-distance atmospheric interactions, including those studied in this paper, are prone to analytical artefacts and are not as strong or simple climate predictors as once thought (Runge et al., Reference Runge, Petoukhov and Kurths2014; McGraw and Barnes, Reference McGraw and Barnes2018).

The studied temporal resolution and timescales could also limit the establishment of correlations between precipitation and these large-scale variables, and thus model accuracy. Previous work has shown that correlations with atmospheric oscillations are highly seasonal and sometimes predictive up to a year in advance (Umirbekov et al., Reference Umirbekov, Peña-Guerrero and Müller2022). Building more seasonal structure into the covariance function could therefore improve the viability of using large-scale features. Furthermore, weekly or daily fluctuations in the ISM and WDs cannot be resolved on a monthly scale. Including these fluctuations could allow the model to capture intra-seasonal variability crucial to predicting extreme precipitation events in the UIB. Future work could also investigate meteorological conditions during these extreme events in more depth and test more climate variables, such as large-scale humidity at different pressure levels.

Going forward, the GP models could also be refined by explicitly incorporating physics linked to precipitation over the UIB. Multiple GPs could be trained to learn the relationships between different variables, e.g., temperature and elevation. The outputs of these models could then be used to predict precipitation. For this purpose, a setup similar to the GP auto-regressive (GPAR) framework by Requeima et al. (Reference Requeima, Tebbutt, Bruinsma and Turner2019) could be considered. More faithful results could also be achieved by incorporating training data from other sources, such as remote sensing or in situ precipitation gauges using multi-fidelity GPs (Tazi et al., Reference Tazi, Orr, Hernandez-González, Hosking and Turner2024).

Additionally, a better scoring function could be used for the feature selection experiments. The MLL is calculated in the warped precipitation space and likely underestimates the loss for large precipitation values. This score also does not penalize for model complexity, which could lead to overly complicated models with too many features. Future model outputs could also be assessed with respect to long-term variability to better reflect their sensitivity to climate change.

Finally, the eventual aim of this work is to generate local precipitation predictions from GCM outputs. To achieve this goal, a GP model could be re-trained on coarsened ERA5 input data and make predictions from GCM projections. This procedure should be straightforward, as the ensemble members of the Coupled Model Intercomparison Project (CMIP) report all the variables needed to derive the input features presented in this paper (Eyring et al., Reference Eyring, Bony, Meehl, Senior, Stevens, Stouffer and Taylor2016; Hersbach et al., Reference Hersbach, Bell, Berrisford, Hirahara, Horányi, Muñoz-Sabater, Nicolas, Peubey, Radu, Schepers, Simmons, Soci, Abdalla, Abellan, Balsamo, Bechtold, Biavati, Bidlot, Bonavita, de Chiara, Dahlgren, Dee, Diamantakis, Dragani, Flemming, Forbes, Fuentes, Geer, Haimberger, Healy, Hogan, Hólm, Janisková, Keeley, Laloyaux, Lopez, Lupu, Radnoti, de Rosnay, Rozum, Vamborg, Villaume and Thépaut2020). The results of this process could enable more accurate predictions of local precipitation distributions over the UIB.

6. Conclusion

This paper has demonstrated that it is possible to predict certain aspects of precipitation over complex orography using variables that are better represented by GCMs, including large-scale circulation fields. With only 6% of the available data for the training period, the whole-basin model with a stationary covariance function is able to predict monthly precipitation with an average $ {R}^2 $ of 0.49, an RMSE of $ 1.25\;\mathrm{mm}\;{\mathrm{d}}^{-1} $ and an MLL of 0.92. The 2 m dew point temperature and the UIB-averaged 2 m temperature were identified as the most predictive features for precipitation. Using a cluster basin model with more tailored input features produces a smaller RMSE ( $ 0.97\;\mathrm{mm}\;{\mathrm{d}}^{-1} $ ) but a higher MLL (1.02). These results improve on the data-limited single-location model, but not the full single-location model (RMSE = 0.50, $ \mathrm{RMSE}=0.71\;\mathrm{mm}\;{\mathrm{d}}^{-1} $ , MLL = 0.87). The stationary whole-basin model also improves on the non-stationary whole-basin model in this data-sparse setting. The non-stationary model could therefore be more suitable for better-sampled or interpolative tasks. In summary, these experiments highlight the trade-offs between learning spatial covariance, incorporating domain knowledge, and increasing computational complexity with GPs. More importantly, this work paves the way for more accurate precipitation predictions with uncertainty distributions and, in turn, more sustainable climate adaptation at a local level.

Open peer review

To view the open peer review materials for this article, please visit http://doi.org/10.1017/eds.2025.10020.

Data availability statement

The code for this paper is available on GitHub and Zenodo (Tazi, Reference Tazi2024). ERA5 data are available through the Copernicus Data Store (Hersbach et al., Reference Hersbach, Bell, Berrisford, Biavati, Horányi, Sabater, Nicolas, Peubey, Radu, Rozum, Schepers, Simmons, Soci, Dee and Thépaut2023a, Reference Hersbach, Bell, Berrisford, Biavati, Horányi, Sabater, Nicolas, Peubey, Radu, Rozum, Schepers, Simmons, Soci, Dee and Thépaut2023b) and climate indices from the US National Ocean and Atmospheric Administration (NOAA, 2011).

Acknowledgements

The authors thank Tony Phillips for providing the upper Indus Basin shapefiles.

Author contribution

Conceptualization: K.T; S.H. Methodology: K.T; R.T. Data curation: K.T. Data visualization: K.T. Original draft: K.T.; R.E.T; A.O.; S.H. All authors approved the final submitted draft.

Funding statement

This work was supported by the Engineering and Physical Sciences Research Council through the AI for Environmental Risk (AI4ER) CDT [grant number: EP/S022961/1; award number: 2270379].

Competing interests

The authors declare none.

Ethical statement

The research meets all ethical guidelines, including adherence to the legal requirements of the studied countries.

A. Metric definitions

A.1. Root mean square error

The root mean square error (RMSE) represents the typical distance of the model from the data. For multiple input–output pairs $ \left\{\boldsymbol{X},\boldsymbol{Y}\right\} $ , the RMSE is given by:

(8) $$ \mathrm{RMSE}=\sqrt{\left\langle {\left(\boldsymbol{Y}-\boldsymbol{f}\right)}^2\right\rangle } $$

where $ \boldsymbol{f} $ is the set of predicted values at $ \boldsymbol{X} $ . We use $ \left\langle \cdot \right\rangle $ here and in the following definitions as a shorthand for the average over the $ N $ data points. The RMSE is more sensitive to outliers and systematic errors than the mean absolute error or bias.

A.2. Coefficient of determination

The coefficient of determination or $ {R}^2 $ represents the fraction of the data’s variance explained by the model. It is given by:

(9) $$ {R}^2=1-\frac{{\mathrm{SS}}_{\mathrm{res}}}{{\mathrm{SS}}_{\mathrm{tot}}}=1-\frac{\sum_i^N{\left({y}_i-{f}_i\right)}^2}{\sum_i^N{\left({y}_i-\left\langle \boldsymbol{Y}\right\rangle \right)}^2} $$

where $ {\mathrm{SS}}_{\mathrm{res}} $ is the sum of the squared residuals and $ {\mathrm{SS}}_{\mathrm{tot}} $ the total sum of squares. An $ {R}^2 $ of 1 indicates a perfect fit, while a negative $ {R}^2 $ means the model performs worse than the mean.

A.3. Mean log loss

The mean log loss (MLL) can be interpreted as measure of uncertainty calibration for probabilistic models. Using the predictive distribution at the test inputs $ \boldsymbol{X} $ , the probability of the target data $ \boldsymbol{Y} $ given the model $ \mathrm{\mathcal{M}} $ , $ p\left(\boldsymbol{Y}|\mathrm{\mathcal{M}}\right) $ , can be calculated. The log loss is given by taking the negative logarithm of this probability (Rasmussen and Williams, Reference Rasmussen and Williams2005). Taking the mean over all inputs gives the MLL:

(10) $$ \mathrm{MLL}=-\left\langle \log p\left(\boldsymbol{Y}|\mathrm{\mathcal{M}}\right)\right\rangle =\left\langle \frac{1}{2}\log \left(2\pi {\boldsymbol{\sigma}}^2\right)+\frac{{\left(\boldsymbol{Y}-\boldsymbol{\mu} \right)}^2}{2{\boldsymbol{\sigma}}^2}\right\rangle, $$

where $ \boldsymbol{\sigma} $ and $ \boldsymbol{\mu} $ are the predicted means and variance at the inputs $ \boldsymbol{X} $ using the model $ \mathrm{\mathcal{M}} $ . Smaller values imply more skill. The MLL is applied to the warped GP predictions.

B. Feature selection experiments

The models input features are selected via the MLL on the validation set. Figures B1, B2, and B3 show the results of greedy feature selection experiments for the single-location, cluster, and stationary whole-basin models trained on the first 5000 points from the training set. Under the assumption that we would like the models to be consistent in time and space, the covariance function are initialized with time for the single-location model and time, latitude and longitude for the cluster and whole-basin models.

Figure B1. Model MLL as a function of selected input features for the single-location GPs. The plot illustrates the greedy feature selection process, where the best previous kernel design is used to test the next most likely feature. Negative changes imply the feature improves model predictive skill. Features associated with positive changes decrease model predictive skill and are not included in subsequent covariance functions.

Figure B2. As Figure B1 for the cluster GPs. Although EOF500C2 reduces the validation MLL of the Ngari GP, we exclude this feature in the final model as it sets the variance of previously selected features to be very low.

Figure B3. As Figure B1 for the stationary whole-basin GP.

Feature rank for the feature selection experiments is established through recursive feature elimination using a random forest ensemble. The recursive feature elimination shows total column water vapour and 2 m dew point temperature to be most important when predicting precipitation. Of the climate indices, the NAO index is most closely linked to precipitation. The first components of the empirical orthogonal functions (EOFs) of the geopotentials over the UIB are the most predictive of the geopotential field variables. The other atmospheric fields are found be to less connected to precipitation than local input features such as slope or elevation. Time-lagged atmospheric indices and large-scale features were also investigated (not shown), but presented no additional correlation with precipitation.

C. Feature variance contribution

Table C1 details the contributed variance of the input features for the cluster and whole-basin models.

Table C1. Feature variance contribution for the cluster (Khyber, Gilgit, Ngari) and stationary whole-basin (UIB) models. The variances are normalized with respect to the training set distribution, with higher values representing larger contributions towards predicting precipitation. Empty entries correspond to features that were not chosen during the feature selection experiments

Footnotes

This research article was awarded Open Data and Open Materials badges for transparent practices. See the Data Availability Statement for details.

References

Andermann, C, Bonnet, S and Gloaguen, R (2011) Evaluation of precipitation data sets along the Himalayan front. Geochemistry, Geophysics, Geosystems 12(7). https://doi.org/10.1029/2011GC003513.CrossRefGoogle Scholar
Andersson, TR, Bruinsma, WP, Markou, S, Requeima, J, Coca-Castro, A, Vaughan, A, Ellis, A-L, Lazzara, MA, Jones, D, Hosking, Set al (2023) Environmental sensor placement with convolutional Gaussian neural processes. Environmental Data Science 2, e32. https://doi.org/10.1017/eds.2023.22CrossRefGoogle Scholar
Arias, P, Bellouin, N, Coppola, E, Jones, R, Krinner, G, Marotzke, J, Naik, V, Palmer, M, Plattner, G-K, Rogelj, J, Rojas, M, Sillmann, J, Storelvmo, T, Trewin, PTB, Rao, KA, Adhikary, B, Allan, R, Armour, K, Bala, G, … Zickfeld, K (2023) Technical summary. In Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change (pp. 35144). Cambridge University Press. https://doi.org/10.1017/9781009157896Google Scholar
Baudouin, J-P, Herzog, M and Petrie, CA (2020) Cross-validating precipitation datasets in the Indus River basin. Hydrology and Earth System Sciences 24(1), 427450. https://doi.org/10.5194/hess-24-427-2020.CrossRefGoogle Scholar
Bharati, P, Deb, P, Hunt, KMR, Orr, A and Dash, MK (2025) ENSO-induced latitudinal variation of the subtropical jet modulates extreme winter precipitation over the Western Himalaya. Adavances in Atmospheric Sciences 42, 111. https://doi.org/10.1007/s00376-024-4057-2.Google Scholar
Bookhagen, B and Burbank, DW (2006) Topography, relief, and TRMM-derived rainfall variations along the Himalaya. Geophysical Research Letters 33(8). https://doi.org/10.1029/2006GL02603.Google Scholar
Bookhagen, B and Burbank, DW (2010) Toward a complete Himalayan hydrological budget: Spatiotemporal distribution of snowmelt and rainfall and their impact on river discharge. Journal of Geophysical Research: Earth Surface 115(F3). https://doi.org/10.1029/2009JF001426.CrossRefGoogle Scholar
Cannon, F, Carvalho, LM, Jones, C, Hoell, A, Norris, J, Kiladis, GN and Tahir, AA (2017) The influence of tropical forcing on extreme winter precipitation in the western Himalaya. Climate Dynamics 48, 12131232. https://doi.org/10.1007/s00382-016-3137-0.CrossRefGoogle Scholar
Chen, Y, Sharma, S, Zhou, X, Yang, K, Li, X, Niu, X, Hu, X and Khadka, N (2021) Spatial performance of multiple reanalysis precipitation datasets on the southern slope of central Himalaya. Atmospheric Research 250, 105365. https://doi.org/10.1016/j.atmosres.2020.105365.CrossRefGoogle Scholar
Chokkavarapu, N and Mandla, VR (2019) Comparative study of GCMs, RCMs, downscaling and hydrological models: A review toward future climate change impact estimation. SN Applied Sciences 1(12), 1698. https://doi.org/10.1007/s42452-019-1764-x.CrossRefGoogle Scholar
Dahri, ZH, Ludwig, F, Moors, E, Ahmad, B, Khan, A and Kabat, P (2016) An appraisal of precipitation distribution in the high-altitude catchments of the Indus basin. Science of the Total Environment 548, 289306. https://doi.org/10.1016/j.scitotenv.2016.01.001.CrossRefGoogle ScholarPubMed
Dahri, ZH, Ludwig, F, Moors, E, Ahmad, S, Ahmad, B, Ahmad, S, Riaz, M and Kabat, P (2021) Climate change and hydrological regime of the high-altitude Indus basin under extreme climate scenarios. Science of the Total Environment 768, 144467. https://doi.org/10.1016/j.scitotenv.2020.144467.CrossRefGoogle ScholarPubMed
Di Capua, G, Kretschmer, M, Donner, RV, Van Den Hurk, B, Vellore, R, Krishnan, R and Coumou, D (2020) Tropical and mid-latitude teleconnections interacting with the Indian summer monsoon rainfall: A theory-guided causal effect network approach. Earth System Dynamics 11(1), 1734. https://doi.org/10.5194/esd-11-17-2020.CrossRefGoogle Scholar
Ding, Q and Wang, B (2005) Circumglobal teleconnection in the northern hemisphere summer. Journal of Climate 18(17), 34833505. https://doi.org/10.1175/JCL13473.1.CrossRefGoogle Scholar
Duvenaud, D (2014) Automatic Model Construction with Gaussian Processes. Doctoral dissertation, University of Cambridge. https://doi.org/10.17863/CAM.14087.CrossRefGoogle Scholar
Eyring, V, Bony, S, Meehl, GA, Senior, CA, Stevens, B, Stouffer, RJ and Taylor, KE (2016) Overview of the coupled model Intercomparison project phase 6 (CMIP6) experimental design and organization. Geoscientific Model Development 9(5), 19371958. https://doi.org/10.5194/gmd-9-1937-2016.CrossRefGoogle Scholar
Filippi, L, Palazzi, E, von Hardenberg, J and Provenzale, A (2014) Multidecadal variations in the relationship between the NAO and winter precipitation in the Hindu Kush–Karakoram. Journal of Climate 27(20), 78907902. https://doi.org/10.1175/JCLI-D-14-00286.1.CrossRefGoogle Scholar
Gerlitz, L, Conrad, O and Böhner, J (2015) Large-scale atmospheric forcing and topographic modification of precipitation rates over high Asia–a neural-network-based approach. Earth System Dynamics 6(1), 6181. https://doi.org/10.5194/esd-6-61-2015.CrossRefGoogle Scholar
Gibbs, MN (1998) Bayesian Gaussian Processes for Regression and Classification. Doctoral dissertation, University of Cambridge. (accessed 11 Nov 2024) https://citeseerx.ist.psu.edu/document?repid=rep1&type=pdf&doi=b5a0c62c8d7cf51137bfb079947b8393c00ed169Google Scholar
Girona-Mata, M, Orr, A, Widmann, M, Bannister, D, Dars, GH, Hosking, S, Norris, J, Ocio, D, Phillips, T, Steiner, J and Richard, ET (2024) Probabilistic precipitation downscaling for ungauged mountain sites: A pilot study for the Hindu Kush Karakoram Himalaya. EGUsphere 2024, 133. https://doi.org/10.5194/egusphere-2024-2805.Google Scholar
Gudmundsson, L, Bremnes, JB, Haugen, JE and Engen-Skaugen, T (2012) Techinal note: Downscaling RCM precipitation to the station scale using statistical transformations and a comparison of methods. Hydrology and Earth System Sciences 16(9), 33833390. https://doi.org/10.5194/hess-16-3383-2012.CrossRefGoogle Scholar
Harris, L, McRae, ATT, Chantry, M, Dueben, PD and Palmer, TN (2022) A generative deep learning approach to stochastic downscaling of precipitation forecasts. Journal of Advances in Modeling Earth Systems 14(10), e2022MS003120. https://doi.org/10.1029/2022MS003120.CrossRefGoogle ScholarPubMed
Hawkins, E and Sutton, R (2009) The potential to narrow uncertainty in regional climate predictions. Bulletin of the American Meteorological Society 90(8), 10951108. https://doi.org/10.1175/2009BAMS2607.1.CrossRefGoogle Scholar
Hersbach, H, Bell, B, Berrisford, P, Hirahara, S, Horányi, A, Muñoz-Sabater, J, Nicolas, J, Peubey, C, Radu, R, Schepers, D, Simmons, A, Soci, C, Abdalla, S, Abellan, X, Balsamo, G, Bechtold, P, Biavati, G, Bidlot, J, Bonavita, M, de Chiara, G, Dahlgren, P, Dee, D, Diamantakis, M, Dragani, R, Flemming, J, Forbes, R, Fuentes, M, Geer, A, Haimberger, L, Healy, S, Hogan, RJ, Hólm, E, Janisková, M, Keeley, S, Laloyaux, P, Lopez, P, Lupu, C, Radnoti, G, de Rosnay, P, Rozum, I, Vamborg, F, Villaume, S and Thépaut, J-N (2020) The ERA5 global reanalysis. Quarterly Journal of the Royal Meteorological Society 146(730), 19992049. https://doi.org/10.1002/qj.3803.CrossRefGoogle Scholar
Hersbach, H, Bell, B, Berrisford, P, Biavati, G, Horányi, A, Sabater, MJ, Nicolas, J, Peubey, C, Radu, R, Rozum, I, Schepers, D, Simmons, A, Soci, C, Dee, D and Thépaut, J-N (2023a) ERA5 Monthly Averaged Data on Pressure Levels from 1940 to Present [Data Set]. https://doi.org/10.24381/cds.6860a573CrossRefGoogle Scholar
Hersbach, H, Bell, B, Berrisford, P, Biavati, G, Horányi, A, Sabater, MJ, Nicolas, J, Peubey, C, Radu, R, Rozum, I, Schepers, D, Simmons, A, Soci, C, Dee, D and Thépaut, J-N (2023b) ERA5 Monthly Averaged Data on Single Levels from 1940 to Present [Data Set]. https://doi.org/10.24381/cds.f17050d7CrossRefGoogle Scholar
Houze, RA (2012) Orographic effects on precipitating clouds. Reviews of Geophysics 50(1). https://doi.org/10.1029/2011RG000365.CrossRefGoogle Scholar
Hunt, KM and Zaz, SN (2023) Linking the North Atlantic Oscillation to winter precipitation over the Western Himalaya through disturbances of the subtropical jet. Climate Dynamics 60(7), 23892403. https://doi.org/10.1007/s00382-022-06450-7.CrossRefGoogle Scholar
Immerzeel, W and Bierkens, M (2012) Asia’s water balance. Nature Geoscience 5(12), 841842. https://doi.org/10.1038/ngeo1643.CrossRefGoogle Scholar
Immerzeel, WW, Van Beek, LP and Bierkens, MF (2010) Climate change will affect the Asian water towers. Science 328(5984), 13821385. https://doi.org/10.1126/science.1183188.CrossRefGoogle ScholarPubMed
Immerzeel, W, Wanders, N, Lutz, A, Shea, J and Bierkens, M (2015) Reconciling high-altitude precipitation in the upper Indus basin with glacier mass balances and runoff. Hydrology and Earth System Sciences 19(11), 46734687. https://doi.org/10.5194/hess-19-4673-2015.CrossRefGoogle Scholar
Immerzeel, WW, Lutz, AF, Andrade, M, Bahl, A, Biemans, H, Bolch, T, Hyde, S, Brumby, SP, Davies, BJ, Elmore, AC, Emmer, A, Feng, M, Fernández, A, Haritashya, UK, Kargel, JS, Koppes, MN, Kraaijenbrink, PDA, Kulkarni, AV, Mayewski, PA, Nepal, S, Pacheco, P, Painter, TH, Pellicciotti, F, Rajaram, H, Rupper, S, Sinisalo, A, Shrestha, AB, Viviroli, D, Wada, Y, Xiao, C, Yao, T and Baillie, JEM (2020) Importance and vulnerability of the world’s water towers. Nature 577(7790), 364369. https://doi.org/10.1038/s41586-019-1822-y.CrossRefGoogle ScholarPubMed
Kelso, NV and Patterson, T (2010) Introducing natural earth data -naturalearthdata.com. Geographia Technica 5(82–89), 25. Accessed 11 November 2024. https://technicalgeography.org/pdf/sp_i_2010/12_introducing_natural_earth_data_naturaleart.pdfGoogle Scholar
Khanal, S, Tiwari, S, Lutz, A, Hurk, B and Immerzeel, W (2023) Historical climate trends over High Mountain Asia derived from ERA5 reanalysis data. Journal of Applied Meteorology and Climatology 62(2), 263288. https://doi.org/10.1175/JAMC-D-21-0045.1.CrossRefGoogle Scholar
Krishnan, R, Shrestha, AB, Ren, G, Rajbhandari, R, Saeed, S, Sanjay, J, Syed, MA, Vellore, R, Xu, Y, You, Q and Ren, Y (2019) Unravelling climate change in the Hindu Kush Himalaya: Rapid warming in the mountains and increasing extremes. In The Hindu Kush Himalaya Assessment: Mountains, Climate Change, Sustainability and People. Springer, pp. 5797. https://doi.org/10.1007/978-3-319-92288-1_3CrossRefGoogle Scholar
Kumar, M, Hodnebrog, Ø, Daloz, AS, Sen, S, Badiger, S and Krishnaswamy, J (2021) Measuring precipitation in Eastern Himalaya: Ground validation of eleven satellite, model and gauge interpolated gridded products. Journal of Hydrology 599, 126252. https://doi.org/10.1016/j.jhydrol.2021.126252.CrossRefGoogle Scholar
Kumar, B, Atey, K, Singh, BB, Chattopadhyay, R, Acharya, N, Singh, M, Nanjundiah, RS and Rao, SA (2023) On the modern deep learning approaches for precipitation downscaling. Earth Science Informatics 16(2), 14591472. https://doi.org/10.1007/s12145-023-00970-4.CrossRefGoogle Scholar
Lalchand, V, Tazi, K, Cheema, TM, Turner, RE and Hosking, S (2022) Kernel learning for explainable climate science. In 16th Bayesian Modelling Applications Workshop at UAI. https://doi.org/10.48550/arXiv.2209.04947CrossRefGoogle Scholar
Leinonen, J, Nerini, D and Berne, A (2021) Stochastic super-resolution for downscaling time-evolving atmospheric fields with a generative adversarial network. IEEE Transactions on Geoscience and Remote Sensing 59(9), 72117223. https://doi.org/10.1109/TGRS.2020.3032790.CrossRefGoogle Scholar
Ling, F, Lu, Z, Luo, J-J, Bai, L, Behera, SK, Jin, D, Pan, B, Jiang, H and Yamagata, T (2024) Diffusion model-based probabilistic downscaling for 180-year East Asian climate reconstruction. npj Climate and Atmospheric Science 7(1), 131. https://doi.org/10.1038/s41612-024-00679-1.CrossRefGoogle Scholar
Liu, H, Motoda, H and Yu, L (2004) A selective sampling approach to active feature selection. Artificial Intelligence 159(1–2), 4974. https://doi.org/10.1016/j.artint.2004.05.009.Google Scholar
Liu, H, Ong, Y-S, Shen, X and Cai, J (2020) When Gaussian process meets big data: A review of scalable GPs. IEEE Transactions on Neural Networks and Learning Systems 31(11), 44054423. https://doi.org/10.1109/TNNLS.2019.2957109.CrossRefGoogle ScholarPubMed
Lutz, A, Immerzeel, W, Shrestha, A and Bierkens, M (2014) Consistent increase in High Asia’s runoff due to increasing glacier melt and precipitation. Nature Climate Change 4(7), 587592. https://doi.org/10.1038/nclimate2237.CrossRefGoogle Scholar
Lutz, AF, Immerzeel, WW, Kraaijenbrink, PD, Shrestha, AB and Bierkens, MF (2016) Climate change impacts on the upper Indus hydrology: Sources, shifts and extremes. PLoS One 11(11), e0165630. https://doi.org/10.1371/journal.pone.0165630.CrossRefGoogle ScholarPubMed
MacKay, DJC (1996) Bayesian non-linear modeling for the prediction competition. In Maximum Entropy and Bayesian Methods: Santa Barbara, California, USA, 1993. Springer, pp. 221234. https://doi.org/10.1007/978-94-015-8729-7_18CrossRefGoogle Scholar
Maraun, D and Widmann, M (2017) Statistical Downscaling and Bias Correction for Climate Research. Cambridge University Press. https://doi.org/10.1017/9781107588783.Google Scholar
Mardani, M, Brenowitz, N, Cohen, Y, Pathak, J, Chen, C-Y, Liu, C-C, Vahdat, A, Nabian, MA, Ge, T, Subramaniam, A, Kashinath, K, Kautz, J and Pritchard, M (2025) Residual corrective diffusion modeling for km-scale atmospheric downscaling. Communications Earth & Environment 6(1), 124. https://doi.org/10.1038/s43247-025-02042-5.CrossRefGoogle Scholar
Martinez-Villalobos, C and Neelin, JD (2019) Why do precipitation intensities tend to follow gamma distributions? Journal of the Atmospheric Sciences 76(11), 36113631. https://doi.org/10.1175/JAS-D-18-0343.1.Google Scholar
McGraw, MC and Barnes, EA (2018) Memory matters: A case for granger causality in climate variability studies. Journal of Climate 31(8), 32893300. https://doi.org/10.1175/JCLI-D-17-0334.1.CrossRefGoogle Scholar
Meng, J, Li, L, Hao, Z, Wang, J and Shao, Q (2014) Suitability of TRMM satellite rainfall in driving a distributed hydrological model in the source region of Yellow River. Journal of Hydrology 509, 320332. https://doi.org/10.1016/j.jhydrol.2013.11.049.CrossRefGoogle Scholar
Midhuna, T, Kumar, P and Dimri, A (2020) A new Western disturbance index for the Indian winter monsoon. Journal of Earth System Science 129, 114. https://doi.org/10.1007/s12040-019-1324-1.CrossRefGoogle Scholar
Mölg, T, Maussion, F and Scherer, D (2014) Mid-latitude westerlies as a driver of glacier variability in monsoonal High Asia. Nature Climate Change 4(1), 6873. https://doi.org/10.1038/nclimate2055.CrossRefGoogle Scholar
NASA JPL (2013) NASA Shuttle Radar Topography Mission Global 1 Arc Second. Data Set. https://doi.org/10.5067/MEASURES/SRTM/SRTMGL1.003.CrossRefGoogle Scholar
National Weather Service Climate Prediction Center (2008) RPCA Technique for Calculating Monthly Teleconnection Indices. Accessed 27 July 2020. https://origin.cpc.ncep.noaa.gov/data/teledoc/teleindcalc.shtml.Google Scholar
NOAA (2011) Climate and Weather Linkage. Accessed 13 October 2023 https://origin.cpc.ncep.noaa.gov/products/precip/CWlink/MJO/climwx.shtml.Google Scholar
Norris, J, Carvalho, LM, Jones, C, Cannon, F, Bookhagen, B, Palazzi, E and Tahir, AA (2017) The spatiotemporal variability of precipitation over the Himalaya: Evaluation of one-year WRF model simulation. Climate Dynamics 49, 21792204. https://doi.org/10.1007/s00382-016-3414-y.CrossRefGoogle Scholar
Norris, J, Carvalho, LM, Jones, C and Cannon, F (2019) Deciphering the contrasting climatic trends between the central Himalaya and Karakoram with 36 years of WRF simulations. Climate Dynamics 52, 159180. https://doi.org/10.1007/s00382-018-4133-3.CrossRefGoogle Scholar
Orr, A, Listowski, C, Couttet, M, Collier, E, Immerzeel, W, Deb, P and Bannister, D (2017) Sensitivity of simulated summer monsoonal precipitation in Langtang Valley, Himalaya, to cloud microphysics schemes in WRF. Journal of Geophysical Research: Atmospheres 122(12), 62986318. https://doi.org/10.1002/2016JD025801.CrossRefGoogle Scholar
Orr, A, Ahmad, B, Alam, U, Appadurai, A, Bharucha, ZP, Biemans, H, Bolch, T, Chaulagain, NP, Dhaubanjar, S, Dimri, AP, Dixon, H, Fowler, HJ, Gioli, G, Halvorson, SJ, Hussain, A, Jeelani, G, Kamal, S, Khalid, IS, Liu, S, Lutz, A, Mehra, MK, Miles, E, Momblanch, A, Muccione, V, Mukherji, A, Mustafa, D, Najmuddin, O, Nasimi, MN, Nüsser, M, Pandey, VP, Parveen, S, Pellicciotti, F, Pollino, C, Potter, E, Qazizada, MR, Ray, S, Romshoo, S, Sarkar, SK, Sawas, A, Sen, S, Shah, A, Shah, MAA, Shea, JM, Sheikh, AT, Shrestha, AB, Tayal, S, Tigala, S, Virk, ZT, Wester, P and Wescoat, JL Jr (2022) Knowledge priorities on climate change and water in the Upper Indus Basin: A horizon scanning exercise to identify the top 100 research questions in social and natural sciences. Earth’s Future 10(4), e2021EF002619. https://doi.org/10.1029/2021EF002619.CrossRefGoogle Scholar
Palazzi, E, Von Hardenberg, J and Provenzale, A (2013) Precipitation in the Hindu-Kush Karakoram Himalaya: Observations and future scenarios. Journal of Geophysical Research: Atmospheres 118(1), 85100. https://doi.org/10.1029/2012JD018697.CrossRefGoogle Scholar
Panday, PK, Thibeault, J and Frey, KE (2015) Changing temperature and precipitation extremes in the Hindu Kush-Himalayan region: An analysis of CMIP3 and CMIP5 simulations and projections. International Journal of Climatology 35(10), 30583077. https://doi.org/10.1002/joc.4192.CrossRefGoogle Scholar
Putra, RI, Ishikawa, T and Tatsubori, M (2022) Spatiotemporal interpolation of ungauged river discharge via deep kernel learning. In 2022 IEEE International Conference on Big Data (Big Data). IEEE, pp. 48764880. https://doi.org/10.1109/BigData55660.2022.10020965CrossRefGoogle Scholar
Rasmussen, CE and Williams, CK (2005) Gaussian Processes for Machine Learning, Vol. 1. The MIT Press. https://doi.org/10.7551/mitpress/3206.001.0001.CrossRefGoogle Scholar
Requeima, J, Tebbutt, W, Bruinsma, W and Turner, RE (2019) The Gaussian process autoregressive regression model (GPAR). In The 22nd International Conference on Artificial Intelligence and Statistics, vol. 89, pp. 18601869. https://doi.org/10.48550/arXiv.1802.07182CrossRefGoogle Scholar
Rocheta, E, Sugiyanto, M, Johnson, F, Evans, J and Sharma, A (2014) How well do general circulation models represent low-frequency rainfall variability? Water Resources Research 50(3), 21082123. https://doi.org/10.1002/2012WR013085.CrossRefGoogle Scholar
Runge, J, Petoukhov, V and Kurths, J (2014) Quantifying the strength and delay of climatic interactions: The ambiguities of cross correlation and a novel measure based on graphical models. Journal of Climate 27(2), 720739. https://doi.org/10.1175/JCLI-D-13-00159.1.CrossRefGoogle Scholar
Sabin, TP, Krishnan, R, Vellore, R, Priya, P, Borgaonkar, HP, Singh, BB and Sagar, A (2020) Climate change over the Himalayas. In Krishnan, R, Sanjay, J, Gnanaseelan, C, Mujumdar, M, Kulkarni, A and Chakraborty, S (eds), Assessment of Climate Change over the Indian Region: A Report of the Ministry of Earth Sciences (MoES), Government of India. Springer Singapore, pp. 207222. https://doi.org/10.1007/978-981-15-4327-2_11.CrossRefGoogle Scholar
Sanjay, J, Krishnan, R, Shrestha, AB, Rajbhandari, R and Ren, G-Y (2017) Downscaled climate change projections for the Hindu Kush Himalayan region using CORDEX South Asia regional climate models. Advances in Climate Change Research 8(3), 185198. https://doi.org/10.1016/j.accre.2017.08.003.CrossRefGoogle Scholar
Schoof, JT and Pryor, S (2001) Downscaling temperature and precipitation: A comparison of regression-based methods and artificial neural networks. International Journal of Climatology: A Journal of the Royal Meteorological Society 21(7), 773790. https://doi.org/10.1002/joc.655.Google Scholar
Shrestha, AB, Wagle, N and Rajbhandari, R (2019) Chapter 6 – A review on the projected changes in climate over the Indus Basin. In Khan, SI and Adams, TE (eds), Indus River Basin. Elsevier, pp. 145158. https://doi.org/10.1016/B978-0-12-812782-7.00007-2.Google Scholar
Shukla, AK, Ojha, CSP, Singh, RP, Pal, L and Fu, D (2019) Evaluation of TRMM precipitation dataset over Himalayan catchment: The upper ganga basin, India. Water 11(3), 613. https://doi.org/10.3390/w11030613.CrossRefGoogle Scholar
Smolenaars, WJ, Lutz, AF, Biemans, H, Dhaubanjar, S, Immerzeel, WW and Ludwig, F (2021) From narratives to numbers: Spatial downscaling and quantification of future water, food & energy security requirements in the Indus basin. Futures 133, 102831. https://doi.org/10.1016/j.futures.2021.102831.CrossRefGoogle Scholar
Snelson, E, Ghahramani, Z and Rasmussen, C (2003) Warped Gaussian processes. Advances in Neural Information Processing Systems 16. https://proceedings.neurips.cc/paper_files/paper/2003/file/6b5754d737784b51ec5075c0dc437bf0-Paper.pdf (accessed 11 May 2025).Google Scholar
Srivastava, P, Yang, R, Kerrigan, G, Dresdner, G, McGibbon, J, Bretherton, C and Mandt, S (2024) Precipitation downscaling with spatiotemporal video diffusion. Advances in Neural Information Processing Systems 37, 5637456400. https://doi.org/10.48550/arXiv.2309.15214.Google Scholar
Stephens, GL, L’Ecuyer, T, Forbes, R, Gettelmen, A, Golaz, J-C, Bodas-Salcedo, A, Suzuki, K, Gabriel, P and Haynes, J (2010) Dreary state of precipitation in global models. Journal of Geophysical Research: Atmospheres 115(D24). https://doi.org/10.1029/2010JD014532.CrossRefGoogle Scholar
Sun, AY, Jiang, P, Yang, Z-L, Xie, Y and Chen, X (2022) A graph neural network (GNN) approach to basin-scale river network learning: The role of physics-based connectivity and data fusion. Hydrology and Earth System Sciences 26(19), 51635184. https://doi.org/10.5194/hess-26-5163-2022.CrossRefGoogle Scholar
Tapiador, FJ, Roca, R, Del Genio, A, Dewitte, B, Petersen, W and Zhang, F (2019) Is precipitation a good metric for model performance? Bulletin of the American Meteorological Society 100(2), 223233. https://doi.org/10.1175/BAMS-D-17-0218.1.CrossRefGoogle ScholarPubMed
Tazi, K (2024) Code for ‘Precipitation Prediction Over the Upper Indus Basinfrom Large-Scale Circulation Patterns Using Gaussian Processes’ [Code]. https://doi.org/10.5281/zenodo.14170951.CrossRefGoogle Scholar
Tazi, K, Lin, JA, Viljoen, R, Gardner, A, John, S, Ge, H and Turner, RE (2023). Beyond intuition, a framework for applying GPs to real-world data. In ICML Workshop on Structured Probabilistic Inference and Generative Modeling. https://doi.org/10.48550/arXiv.2307.03093.CrossRefGoogle Scholar
Tazi, K, Orr, A, Hernandez-González, J, Hosking, S and Turner, RE (2024) Downscaling precipitation over high-mountain Asia using multi-fidelity Gaussian processes: Improved estimates from ERA5. Hydrology and Earth System Sciences 28(22), 49034925. https://doi.org/10.5194/hess-28-4903-2024.CrossRefGoogle Scholar
TROPOMI (2019) GMTED2010 Elevation Data at Different Resolutions [Data Set]. (accessed 11 Nov 2024) https://www.temis.nl/data/gmted2010/index.phpGoogle Scholar
Umirbekov, A, Peña-Guerrero, MD and Müller, D (2022) Regionalization of climate teleconnections across central Asian mountains improves the predictability of seasonal precipitation. Environmental Research Letters 17(5), 055002. https://doi.org/10.1088/1748-9326/ac6229.CrossRefGoogle Scholar
Vaughan, A, Tebbutt, W, Hosking, JS and Turner, RE (2022) Convolutional conditional neural processes for local climate downscaling. Geoscientific Model Development 15(1), 251268. https://doi.org/10.5194/gmd-15-251-2022.CrossRefGoogle Scholar
Wang, F, Tian, D, Lowe, L, Kalin, L and Lehrter, J (2021) Deep learning for daily precipitation and temperature downscaling. Water Resources Research 57(4), e2020WR029308. https://doi.org/10.1029/2020WR029308.CrossRefGoogle Scholar
Weare, BC and Nasstrom, JS (1982) Examples of extended empirical orthogonal function analyses. Monthly Weather Review 110(6), 481485. https://doi.org/10.1175/1520-0493(1982)110%3C0481:EOEEOF%3E2.0.CO;2.2.0.CO;2>CrossRefGoogle Scholar
Wolvin, S, Strong, C, Rupper, S and Steenburgh, W (2024) Climatology of orographic precipitation gradients over High Mountain Asia derived from dynamical downscaling. Journal of Geophysical Research: Atmospheres 129(20), e2024JD041010. https://doi.org/10.1029/2024JD041010.CrossRefGoogle Scholar
Yin, Z-Y, Zhang, X, Liu, X, Colella, M and Chen, X (2008) An assessment of the biases of satellite rainfall estimates over the Tibetan plateau and correction methods based on topographic analysis. Journal of Hydrometeorology 9(3), 301326. https://doi.org/10.1175/2007JHM903.1.CrossRefGoogle Scholar
Zhong, X, Du, F, Chen, L, Wang, Z and Li, H (2024) Investigating transformer-based models for spatial downscaling and correcting biases of near-surface temperature and wind-speed forecasts. Quarterly Journal of the Royal Meteorological Society 150(758), 275289. https://doi.org/10.1002/qj.4596.CrossRefGoogle Scholar
Figure 0

Figure 1. Map of Indus Basin showing elevation (TROPOMI, 2019), the Indus River (Kelso and Patterson, 2010, in blue) and the UIB (NASA JPL, 2013, watershed boundary shown with dashed line). The inset shows the watershed’s location with respect to the Asian continent.

Figure 1

Figure 2. ERA5 monthly precipitation over the UIB between 1970 and 2004. The average annual and seasonal precipitation over the UIB (left) and time series for three locations in the UIB (right) are shown. The time series shows the typical precipitation regimes in the basin (Figure 3) and are fit with a linear regression model (dotted green lines) to identify trends over the studied time period.

Figure 2

Figure 3. UIB precipitation regimes. Clusters are generated by applying soft $ k $-means to ERA5 monthly precipitation data from 1970 to 2004. The figure shows the clusters for a weighting threshold of 0.6, following the approach of Lalchand et al. (2022).

Figure 3

Table 1. Considered model features with dimensions and abbreviations. These monthly variables are or are derived from ERA5 (Hersbach et al., 2020), with the exception of the atmospheric indices, which are calculated by NOAA (2011) using historical observations

Figure 4

Table 2. Performance of the single-location(SLM), cluster (Khyber, Gilgit, Ngari), and whole-basin models (WBM). The RMSE is given in $ \mathrm{mm}\;{\mathrm{d}}^{-1} $

Figure 5

Figure 4. Test $ {R}^2 $ (left), RMSE (centre), and MLL (right) for the full single-location model across the UIB evaluated on all precipitation data between 2005 and 2019.

Figure 6

Figure 5. Learnt non-stationary kernel lengthscales over the UIB for longitude, latitude and their magnitude.

Figure 7

Figure B1. Model MLL as a function of selected input features for the single-location GPs. The plot illustrates the greedy feature selection process, where the best previous kernel design is used to test the next most likely feature. Negative changes imply the feature improves model predictive skill. Features associated with positive changes decrease model predictive skill and are not included in subsequent covariance functions.

Figure 8

Figure B2. As Figure B1 for the cluster GPs. Although EOF500C2 reduces the validation MLL of the Ngari GP, we exclude this feature in the final model as it sets the variance of previously selected features to be very low.

Figure 9

Figure B3. As Figure B1 for the stationary whole-basin GP.

Figure 10

Table C1. Feature variance contribution for the cluster (Khyber, Gilgit, Ngari) and stationary whole-basin (UIB) models. The variances are normalized with respect to the training set distribution, with higher values representing larger contributions towards predicting precipitation. Empty entries correspond to features that were not chosen during the feature selection experiments

Author comment: Precipitation prediction over the upper Indus Basin from large-scale circulation patterns using Gaussian processes — R0/PR1

Comments

No accompanying comment.

Review: Precipitation prediction over the upper Indus Basin from large-scale circulation patterns using Gaussian processes — R0/PR2

Conflict of interest statement

None

Comments

Review report: Precipitation prediction over the upper Indus Basin from large-scale circulation patterns using Gaussian processes

This paper presents a study that uses Gaussian process to predict precipitation over the upper Indus Basin (UIB) region. Several model configurations including single-location, clustered and whole basin models are evaluated. The work serves well as a novel proof-of-concept for applying probabilistic AI models with uncertainty quantification for future precipitation prediction in the UIB region, which is a region highly vulnerable to climate change.

The paper is well-written with clear methodology, appropriate data and impactful applications. It demonstrated the feasibility to use GCM variables to predict precipitation over complex mountain terrain in UIB with probabilistic machine learning model (Gaussian Process). This machine learning based approach represents a novel alternative to the traditional dynamical down-scaling approaches. As an application paper is suitable for publication for Environmental Data Science (EDS).

The methodology is sound and well executed. The choice of data (ERA5) is justified. The design of experiments is appropriate with proper performance metrics. Different experimental setup and the trade-offs between different configurations are explored and discussed. The paper clearly presents the work and is in an appropriate length with clear figures and tables that effectively support the narrative. There are right amount of description of the background and methodology that making the paper easy to read for people from either the machine learning community or the environmental science community.

To further improve the paper, first, there are some well-known AI down-scaling studies available in literature using models like GAN and diffusion models achieving notable results. I think it is valid that the author argues that probabilistic machine learning models like GP can address uncertainties, however it is still beneficial to review some existing AI down-scaling literature in additional to reviews on dynamical down-scaling. Secondly, if possible, to have a discussion or comments on the feature selection outcomes in Appendix B will be interesting, especially to compare with human knowledge to understand if the machine learning algorithm are more sensitive to features that human expert thinks are most important. I had some concerns with regards to the computational scalability of GP, but it has been well addressed in the limitations and future work sections. I recommend the paper to be accepted with minor addition of literature review and discussions.

Review: Precipitation prediction over the upper Indus Basin from large-scale circulation patterns using Gaussian processes — R0/PR3

Conflict of interest statement

No.

Comments

This work is an extension of the earlier submitted work to the Climate Informatics conference. The paper aims to explore the machine learning approach to predict precipitation over the upper Indus Basin using large-scale climate circulation.

The current version of the work improved upon the previous work but I have a few suggestions for the authors to consider.

1. Feature selection is important for the readers to understand the model performance. If authors want to keep the feature selection details in the Appendix, at least authors should consider including what major variables are included in each model in the main text.

2. Many input variables have strong correlations (for those non-EOF variables). Will there be benefit to perform a PCA for the input features before feeding into the GP model?

3. The GP model also can generate uncertainty of the model output which is one of the advantages of the GP model. This can also be useful for the assessment of the model estimation (e.g., understanding whether the difference between the estimation and target is within the range of the model uncertainty). It will be good to show some analysis of the GP model output uncertainty.

4. There is recent development of more efficient GP models that may address the issues that the authors mentioned about the computational cost. Authors may want to explore these new GP implementations as well.

Recommendation: Precipitation prediction over the upper Indus Basin from large-scale circulation patterns using Gaussian processes — R0/PR4

Comments

No accompanying comment.

Decision: Precipitation prediction over the upper Indus Basin from large-scale circulation patterns using Gaussian processes — R0/PR5

Comments

No accompanying comment.

Author comment: Precipitation prediction over the upper Indus Basin from large-scale circulation patterns using Gaussian processes — R1/PR6

Comments

NA

Review: Precipitation prediction over the upper Indus Basin from large-scale circulation patterns using Gaussian processes — R1/PR7

Conflict of interest statement

Reviewer declares none.

Comments

The reply to the earlier review comments are satisfactory, I recommend the publication of this manuscript without further comments.

Recommendation: Precipitation prediction over the upper Indus Basin from large-scale circulation patterns using Gaussian processes — R1/PR8

Comments

No accompanying comment.

Decision: Precipitation prediction over the upper Indus Basin from large-scale circulation patterns using Gaussian processes — R1/PR9

Comments

No accompanying comment.