Fact-checked by Grok 2 weeks ago

Hydrological model

A hydrological model is a simplified mathematical representation of the hydrologic system that simulates the movement, storage, and transformation of within a or catchment, aiding in the prediction of runoff, , and components. These models integrate physical processes such as , , infiltration, and to forecast responses to environmental stresses like variability or land-use changes. Developed since the mid-19th century with early empirical methods like the rational method (developed by T. Mulvany in 1851), hydrological modeling evolved significantly in the 1960s with the advent of digital computers and conceptual models such as the Stanford Watershed Model, leading to modern process-based and data-driven approaches. Hydrological models are classified into several types, including lumped conceptual models that aggregate processes at the catchment scale (e.g., HBV model), distributed physically-based models that resolve spatial variations (e.g., MIKE SHE), and semi-distributed models like for agricultural watersheds. They are widely applied in water resource planning, , assessment, and evaluating the impacts of or on ecosystems. For instance, hydrological models (GHMs) extend these simulations to or planetary scales to analyze water availability under future scenarios, incorporating human influences like reservoirs and . Ongoing advancements focus on improving model accuracy through integration with data and , addressing uncertainties in parameter estimation and calibration.

Introduction

Definition and purpose

A hydrological model is a simplified mathematical or conceptual representation of the terrestrial hydrological system, designed to simulate the movement, storage, and transformation of through various environmental compartments. These models employ variables and equations to approximate fluxes of water across system boundaries, capturing essential processes such as infiltration, , and . By abstracting complex real-world dynamics into computable forms, they enable the prediction and analysis of water-related phenomena under diverse conditions. The primary purposes of hydrological models include forecasting and risks, evaluating water resource availability, informing water management decisions, and assessing the environmental consequences of land-use changes or variability. For instance, they support engineering applications in and planning, while also aiding policymakers in sustainable and impact . These objectives are achieved by integrating observed to replicate historical events and project future scenarios, thereby enhancing resilience to hydrological uncertainties. Central to these models are key concepts such as inputs—typically , rates, and properties—and outputs like runoff volumes and . A foundational principle is the , which expresses the in a hydrological :
P = Q + E + \Delta S
where P denotes , Q represents runoff, E is , and \Delta S indicates change in storage (e.g., in or aquifers). This underpins model formulations by ensuring that inflows equal outflows plus storage variations over a defined period.
Hydrological models encompass both surface water processes (e.g., overland flow and channel routing) and subsurface dynamics (e.g., vadose zone transport and aquifer flow), applicable across scales from individual small catchments to continental or global domains. This versatility allows them to address localized flood events as well as broader issues like basin-wide water scarcity.

Historical development

The development of hydrological modeling began in the early 19th century with empirical approaches to estimate runoff. In 1851, Thomas Mulvany introduced the concept of time of concentration, which represented the time required for runoff from the most distant point in a catchment to reach the outlet, serving as a foundational element for the rational method of peak flow estimation. This method linked rainfall intensity, catchment area, and runoff coefficient to predict maximum discharge, marking an initial shift toward quantitative hydrological analysis. By the mid-20th century, advancements in empirical techniques expanded modeling capabilities. Leroy K. Sherman proposed the unit hydrograph theory in 1932, defining it as the hydrograph resulting from one unit of excess rainfall over a specified duration, assuming linear and time-invariant response. This approach, refined through to , enabled the synthesis of flood s from observed data and became a cornerstone for event-based simulations. In the , Norman H. Crawford and Ray K. Linsley developed the Stanford Model, one of the first digital conceptual models that simulated continuous catchment processes such as infiltration, , and routing using a series of interconnected storage elements. This model, formalized in 1966, represented a precursor to modern conceptual frameworks by integrating multiple hydrological components into a computational structure. The 1970s and 1980s saw the rise of digital computers, which facilitated more complex simulations and the emergence of physically-based models. These models drew on fundamental physical laws, such as the originally formulated by Lighthill and Whitham in 1955 but increasingly applied in hydrological contexts during this period to approximate overland and channel flow without diffusion effects. A key milestone was the International Hydrological Decade (1965-1974), initiated by and supported by the International Association of Hydrological Sciences (IAHS), which fostered global collaboration, standardized data collection, and promoted the development of consistent modeling practices. From the 1990s onward, hydrological modeling evolved with the integration of geographic information systems (GIS) and , enabling spatially explicit representations and a pronounced shift toward distributed models that account for heterogeneity across watersheds. In the 2000s, models increasingly incorporated scenarios, linking hydrological simulations with global climate projections to assess impacts on , such as altered runoff regimes and flood risks. This era emphasized ensemble approaches and to support adaptive water management.

Model Classifications

Lumped versus distributed

Hydrological models are classified based on their spatial representation of the catchment, primarily into lumped, distributed, and semi-distributed approaches. Lumped models treat the entire catchment as a single, homogeneous unit, averaging parameters such as , properties, and infiltration rates across the without accounting for internal spatial variations. This simplification assumes uniform hydrological responses, making these models suitable for small, relatively uniform catchments where detailed spatial data are unavailable. A key advantage of lumped models is their computational efficiency and low data requirements, as they require fewer inputs and are easier to calibrate compared to more complex alternatives. Examples include the Rational Method for peak flow estimation and simple linear reservoir models, which represent storage and outflow using a single compartment to simulate runoff generation. In contrast, distributed models divide the catchment into discrete elements, such as grid cells or sub-basins, allowing for explicit representation of spatial variability in inputs like rainfall, , types, and . These models simulate hydrological processes at each element and route flows between them, capturing heterogeneities that influence runoff paths and timing. They typically rely on high-resolution data sources, including digital elevation models (DEMs) for terrain analysis and maps for and distribution. Distributed approaches, such as the Distributed Hydrology Soil Vegetation Model (DHSVM), provide greater accuracy in predicting responses in large or heterogeneous terrains, where spatial variations significantly affect overall . Semi-distributed models serve as a , aggregating the catchment into sub-units based on dominant characteristics like and , rather than fully resolving every spatial detail. A prominent example is the Soil and Water Assessment Tool (), which uses hydrologic response units (HRUs)—areas with uniform , , and slope—to simulate processes while routing flows through sub-basins. This approach balances physical realism with practicality, offering improved estimation over purely lumped models by incorporating some spatial structure, yet avoiding the full complexity of distributed systems. The choice between these approaches involves trade-offs in computational demands, accuracy, and applicability. Lumped models are favored for their simplicity and speed, particularly in data-scarce regions or for quick assessments, but they may underestimate peak flows or overlook variability in diverse landscapes. Distributed models enhance accuracy in capturing spatial processes, such as variable infiltration or runoff routing in mountainous areas, but incur higher computational costs and require extensive data for , which can be challenging in practice. Semi-distributed options like mitigate these issues by reducing parameter numbers while maintaining reasonable fidelity, making them suitable for basin-scale applications. Selection criteria often depend on catchment scale, data availability, and study objectives, with lumped models preferred for small, uniform areas and distributed ones for larger, heterogeneous systems.

Event-based versus continuous simulation

Hydrological models are classified temporally into event-based and continuous simulation approaches, distinguishing how they represent responses over time. Event-based models simulate the hydrological response to a single, discrete rainfall event, such as a or , focusing primarily on direct runoff, peak discharges, and short-term dynamics. These models typically require inputs limited to the event's rainfall hyetograph, duration, and basic properties, without explicitly tracking long-term state variables. A representative in event-based modeling is the unit hydrograph technique, which generates a runoff from a unit volume of excess rainfall over a specified duration, enabling with observed rainfall for event-specific predictions. This approach excels in applications like peak estimation and stormwater design, where rapid computation is essential. However, event-based models often assume simplified or fixed initial conditions, ignoring the carryover effects of prior wetness, which can lead to underestimation of runoff in sequences of storms. Continuous simulation models, by contrast, run over extended periods—often years or decades—using time series of meteorological data to replicate the full spectrum of hydrological processes, including wet-dry cycles, soil moisture evolution, and baseflow recession. These models incorporate dynamic state variables, such as antecedent soil moisture, to capture inter-event dependencies and provide outputs like annual water yields or seasonal flow regimes. They are particularly valuable for long-term planning, such as assessing reliability or climate change impacts on . While event-based models offer advantages in simplicity, lower data requirements, and faster execution for isolated event analysis, their limitations include poor representation of cumulative effects and sensitivity to assumed states. Continuous models mitigate these issues by simulating antecedent conditions implicitly but demand extensive historical records, sophisticated , and greater computational resources, which can complicate in data-scarce regions. Selection of the simulation type hinges on the modeling objective and data availability: event-based approaches suit short-term flood risk assessments where peak flows dominate, whereas continuous methods are preferred for integrated water management tasks like reservoir operations or flood frequency analysis, especially when long-term meteorological series are accessible.

Types of Hydrological Models

Statistical and empirical models

Statistical and empirical models in rely on observed patterns and historical correlations to estimate runoff and without simulating underlying physical processes. These approaches are particularly valuable for providing rapid assessments in regions with limited , where they derive relationships between inputs like rainfall and outputs like using statistical techniques. Unlike conceptual models that incorporate simplified representations of hydrological processes, statistical and empirical methods prioritize data-driven correlations for and . Empirical models, such as the Rational Method, offer simple formulas based on long-established correlations from field observations. The Rational Method estimates peak discharge Q for small areas using the equation Q = C \cdot I \cdot A where Q is the peak discharge in cubic feet per second, C is the dimensionless runoff coefficient reflecting and (typically ranging from 0.10 to 0.97), I is the rainfall in inches per hour over the , and A is the area in acres; a unit conversion factor of 1.008 is often applied for consistency. Originating from early observations in the late and formalized by Kuichling in 1889, this method assumes uniform rainfall distribution and that the storm duration equals the basin's , making it suitable for basins up to 200 acres but prone to inaccuracies in larger or heterogeneous areas due to unaccounted storage effects. Statistical approaches extend these empirical foundations by employing and to model rainfall-runoff relationships. , including multiple linear regression (MLR), relates streamflow statistics to basin characteristics like area, , and , often using logarithmic transformations to linearize and improve fit; for instance, MLR has been applied to predict runoff signatures such as annual maximum flows with coefficients of determination around 0.7–0.9 in various catchments. (ARIMA) models capture temporal dependencies in through autoregressive (AR), differencing for stationarity (I), and (MA) components, following the Box-Jenkins for identification, , and diagnostics; seasonal like SARIMA account for periodicity in monthly , enabling short-term forecasting in river basins. Parameter in these models frequently involves of statistical moments (e.g., mean and variance) and functions to ensure model adequacy and preserve hydrologic sequence properties. These models find primary applications in data-scarce or ungaged basins for quick peak flow estimates and low-flow predictions, such as in urban drainage design or regional water resource planning, where they provide unbiased estimates within predictor variable ranges but exhibit limitations in extrapolating beyond data due to unmodeled nonlinearities and changing conditions. A prominent example is the U.S. Geological Survey's (USGS) regional equations, which use multilinear with to estimate natural statistics (e.g., 7-day minimum flows or annual means) from basin and climatic predictors across defined hydrologic regions; applied in states like and , these equations achieve prediction errors of 20–50% for ungaged sites matching regional characteristics. Overall, while effective for preliminary assessments, their reliance on historical correlations underscores the need for validation against observed data to mitigate uncertainties in non-stationary environments.

Conceptual models

Conceptual hydrological models represent catchment processes through simplified compartments that capture key storages and fluxes, providing a balance between computational feasibility and realistic simulation of dynamics. These models typically divide the hydrological system into interconnected storages, such as (for canopy ), surface storage (for overland ), (for unsaturated retention), and (for deeper aquifers), with fluxes including input, infiltration into , percolation to deeper layers, losses, and various runoff components like surface, interflow, and . This structure relies on empirical or semi-empirical transfer functions to approximate physical processes, avoiding the full complexity of equations used in more detailed approaches. A seminal example is the HBV model, developed in the 1970s by Swedish hydrologists at the Swedish Meteorological and Hydrological Institute. The HBV model employs storages for snow accumulation and melt, accounting, and two linear reservoirs for upper and lower zone routing, with fluxes governed by threshold-based infiltration, a non-linear soil recharge function, and recession-based runoff generation. Another widely adopted model is the Sacramento Accounting (SAC-SMA), introduced by the in the early 1970s. SAC-SMA features a two-layer with tension and free water storages in both upper (shallow, interception-influenced) and lower (deeper, baseflow-contributing) zones, where fluxes include tension-limited , thresholds, and partitioned runoff via interflow and . These models offer advantages in requiring moderate data inputs—primarily , , and —and low computational demands, making them suitable for operational forecasting in data-scarce regions. Their parameters, such as storage capacities and recession coefficients, are interpretable as approximations of physical catchment properties like depth and , facilitating insights into hydrological behavior. However, limitations arise from the lumped parameterization, which assumes spatial uniformity and thus overlooks heterogeneity in larger or varied catchments. Additionally, parameters are typically not directly measurable and must be inferred through against observed data, leading to equifinality issues where multiple parameter sets yield similar simulations.

Physically-based models

Physically-based hydrological models are constructed using fundamental physical principles, primarily the equations for , , and , to simulate and related processes across catchments. These models explicitly represent hydrological phenomena through partial differential equations derived from , allowing for detailed, process-oriented simulations without relying on empirical simplifications. A key example is the Richards equation, which governs variably saturated flow in porous media: \frac{\partial \theta}{\partial t} = \nabla \cdot [K(\theta) \nabla (h)] - S where \theta is the volumetric moisture content, t is time, K(\theta) is the hydraulic conductivity, h is the pressure head, and S is a sink term accounting for water uptake. This equation, originally formulated for capillary conduction in soils, forms the basis for subsurface flow components in many such models. Prominent examples of physically-based models include MIKE SHE and ParFlow, both of which integrate surface and subsurface processes in a spatially distributed framework. MIKE SHE, developed as a comprehensive system for simulating overland flow, unsaturated zone dynamics, groundwater flow, and channel routing, solves the Saint-Venant equations for surface flow alongside the Richards equation for vadose zone transport, enabling coupled assessments of evapotranspiration, infiltration, and recharge. ParFlow, on the other hand, focuses on three-dimensional variably saturated subsurface flow using a finite-difference discretization of the Richards equation, integrated with kinematic wave approximations for overland flow and the Community Land Model for surface energy balance, making it suitable for large-scale simulations of groundwater-surface water interactions. These models offer significant advantages in predictive capability, particularly in data-scarce regions where empirical is limited, as their reliance on physical laws allows beyond observed conditions. They also effectively capture in properties, , and land use, providing insights into process interactions at fine resolutions that lumped approaches cannot achieve. However, physically-based models face challenges related to extensive parameterization requirements, as they demand detailed inputs for hydraulic properties, soil textures, and boundary conditions, often leading to equifinality issues where multiple parameter sets yield similar outputs. Additionally, their computational demands are high, especially for three-dimensional, high-resolution applications over large domains, necessitating and simplified assumptions to manage runtime.

Data-driven models

Data-driven hydrological models leverage and techniques to infer complex relationships between inputs and outputs directly from observational data, bypassing the need for explicit physical equations or conceptual structures. These models treat hydrological systems as black-box processes, focusing on from historical records such as , , and meteorological variables. Unlike process-based approaches, they excel in capturing nonlinear dynamics inherent in hydrological phenomena without requiring detailed knowledge of underlying mechanisms. Key approaches include artificial neural networks (ANNs), support vector machines (SVMs), and recurrent neural networks like (LSTM) units, which are particularly suited for time-series forecasting. ANNs, often implemented as multilayer perceptrons (MLPs), process inputs through interconnected layers to model nonlinear mappings, while SVMs use kernel functions to map into higher-dimensional spaces for tasks. LSTMs address the challenges of long-term dependencies in sequential by incorporating memory cells and gates to selectively retain or forget information over extended periods. Additionally, ensemble methods such as random forests aggregate multiple decision trees to estimate model parameters or predict outputs, enhancing robustness through bagging and feature randomization. Prominent examples demonstrate their versatility: ANNs have been widely applied to rainfall-runoff modeling, where they predict from and inputs, achieving high accuracy in diverse catchments. SVMs support and forecasting by regressing hydrological variables like river discharge against climatic drivers. LSTMs enable real-time prediction using global datasets, such as ERA5 reanalysis and HydroMT attributes, performing effectively across ungauged basins in regions like the and . Random forests facilitate parameter estimation for lumped models like GR4H, regionalizing values based on catchment descriptors to simulate hourly runoff in urban areas. These models offer significant advantages, including their ability to handle nonlinearities and large volumes of heterogeneous data, such as satellite-derived observations from sources like CHIRPS or GRACE, which traditional models struggle to integrate efficiently. In the 2020s, advances in deep learning, including convolutional LSTMs and transformer-based architectures, have improved real-time forecasting by processing spatiotemporal data at scale, often outperforming conceptual models in accuracy for tasks like precipitation nowcasting and evapotranspiration estimation. Their flexibility allows adaptation to data-scarce environments through transfer learning, reducing the need for site-specific calibration. However, data-driven models face notable limitations, primarily their black-box nature, which hinders interpretability and makes it difficult to discern physical insights or validate against hydrological principles. They often exhibit poor generalization beyond training data distributions, leading to unreliable extrapolations in changing climates or ungauged sites, and are prone to without sufficient regularization. Additionally, their performance degrades with noisy or incomplete datasets, common in , and they lack inherent guarantees for mass or physical . Efforts to address these through explainable techniques, like SHAP values, are emerging but remain computationally intensive.

Hybrid models

Hybrid models in hydrology integrate multiple modeling paradigms, such as conceptual, physically-based, or data-driven approaches, to leverage their complementary strengths and mitigate individual limitations. These models typically combine process-based representations of hydrological dynamics with (ML) techniques to enhance predictive performance, particularly in scenarios involving complex nonlinear interactions or data scarcity. For instance, hybrid frameworks often employ conceptual models for structured simulation of components, augmented by ML for refining outputs or optimizing parameters. One common type involves coupling conceptual models with for post-processing, such as error correction, where algorithms learn and adjust biases in model simulations to improve accuracy. The Modeling Error Learning based Post-Processor (MELPF) framework exemplifies this by using neural networks to post-process outputs from conceptual models like HBV, reducing systematic errors in predictions by up to 20-30% in tested basins. Similarly, physically-based models are hybridized with data-driven methods for parameter optimization, where surrogates accelerate of computationally intensive parameters, as seen in frameworks integrating the and Water Assessment Tool () with artificial neural networks (ANN) to optimize soil and parameters for better representation of subsurface flows. A notable example is the -ANN coupled approach for prediction, which uses ANN to correct nitrate load estimates from , achieving Nash-Sutcliffe efficiency improvements of 0.15-0.25 in forested watersheds compared to standalone . Recent advances from 2023-2025 emphasize ensemble hybrid models for climate projections, integrating physically-based hydrological models with -enhanced statistical to generate robust projections of water availability under changing climates. For example, multi-model ensembles combining global climate models (GCMs) with hybrid numerical models (HNMs) like those in the ISIMIP3b dataset have improved groundwater-inclusive projections by incorporating for bias correction. These developments build on data-driven components like networks for in historical data. The primary benefits of hybrid models include enhanced interpretability from physical components alongside the high accuracy and adaptability of , enabling better handling of uncertainties in non-stationary conditions like . By addressing limitations such as the computational demands of full physically-based models or the lack of mechanistic insight in pure data-driven approaches, hybrids improve overall robustness for applications like . However, challenges persist in the increased complexity of model coupling, which requires sophisticated interfaces and can complicate validation, often leading to higher risks of or propagation of errors across components. Rigorous cross-validation and sensitivity analyses are essential to ensure reliable integration.

Model Development

Key components and inputs

Hydrological models rely on several key components to represent the physical processes governing water movement within a . These include meteorological forcings, which provide the primary drivers of hydrological responses, such as and that influence , infiltration, and runoff generation. Topographic data, encompassing elevation and slope, are essential for delineating catchment boundaries, flow directions, and routing pathways, often derived from digital elevation models (DEMs) to capture terrain variability. Land surface properties, including , vegetation cover, and , determine infiltration capacities, rates, and characteristics, with models like the Soil and Water Assessment Tool () incorporating these via hydrologic response units that aggregate soil and vegetation attributes. Inputs to hydrological models typically consist of time-series data for meteorological forcings, sourced from ground-based rain gauges for precise local measurements or satellite observations for broader coverage, such as the (GPM) mission's Integrated Multi-satellitE Retrievals (IMERG) product, which provides half-hourly estimates at 0.1° resolution to support real-time modeling in data-sparse regions. Initial conditions, particularly content at the start of simulations, are critical for initializing storage states and are often estimated from prior model runs, satellite-derived products like those from the (SMAP) mission, or field measurements to ensure accurate representation of antecedent wetness. Model outputs generally include hydrographs depicting temporal variations at outlets and spatial maps of water fluxes such as , infiltration, and overland flow across the domain, enabling assessment of hydrological responses at multiple scales. Parameters play a pivotal role in tuning these processes; for instance, Manning's roughness coefficient () quantifies channel and overland in modules, with typical values ranging from 0.01 for smooth channels to 0.15 for dense , influencing the and translation of flood waves in models like HEC-HMS. Data preparation is a foundational step to ensure input quality, involving the handling of missing values in time-series through imputation techniques such as or methods like k-nearest neighbors (kNN) to maintain continuity in gauge or satellite records without introducing significant bias. For distributed models, scaling procedures are applied to align data resolutions, such as resampling high-resolution topographic grids to match coarser meteorological inputs or aggregating land surface properties into representative grid cells, which mitigates inconsistencies and enhances computational efficiency.

Governing equations

Hydrological models rely on fundamental governing equations derived from physical principles to represent water movement and storage in the hydrologic cycle. The continuity equation, also known as the mass balance equation, forms the cornerstone of these models by enforcing conservation of mass. In its general three-dimensional form for porous media or fluid flow, it states that the divergence of the flux vector \mathbf{q} equals the negative rate of change of water depth or storage h plus any sources or sinks, expressed as \nabla \cdot \mathbf{q} = -\frac{\partial h}{\partial t} + S, where S represents net inflows or outflows such as precipitation or extraction. This equation is derived from the principle of mass conservation applied to a control volume: the net mass flux across the boundaries must equal the rate of change of mass within the volume, assuming incompressible flow and no chemical reactions altering water mass. Assumptions include constant fluid density and neglect of minor diffusive terms unless explicitly included in advanced models. In one-dimensional channel routing for hydrological applications, such as kinematic-wave models, it simplifies to \frac{\partial A}{\partial t} + \frac{\partial Q}{\partial x} = q_l, where A is the cross-sectional flow area, Q is discharge, x is distance along the channel, t is time, and q_l is lateral inflow per unit length; this form arises by integrating the general equation over the cross-section and assuming hydrostatic pressure and prismatic geometry. For surface flow routing in rivers and overland areas, the Saint-Venant equations provide the standard framework, consisting of coupled continuity and momentum equations for unsteady, one-dimensional open-channel flow. The continuity equation is \frac{\partial A}{\partial t} + \frac{\partial Q}{\partial x} = 0 (assuming no lateral inflow), derived as above but tailored to varying cross-sections. The momentum equation is \frac{\partial Q}{\partial t} + \frac{\partial}{\partial x} \left( \frac{Q^2}{A} \right) + gA \left( \frac{\partial h}{\partial x} + S_f \right) = 0, where g is gravitational acceleration, h is water depth, and S_f is the friction slope (often from Manning's equation, S_f = \frac{n^2 Q^2}{A^2 R^{4/3}}, with n as Manning's roughness and R as hydraulic radius). This momentum form originates from Newton's second law applied to a channel reach: the rate of change of momentum plus net pressure and friction forces equals inertial forces, with hydrostatic pressure distribution assumed (pressure increases linearly with depth) and neglecting viscous shear and Coriolis effects for typical hydrological scales. Common assumptions include gradually varied flow (wavelength much larger than depth), small bed slopes, and quasi-steady friction; full dynamic wave solutions require both equations, while kinematic approximations drop the \frac{\partial h}{\partial x} term for steep slopes where friction dominates inertia. These equations, originally formulated by Barré de Saint-Venant in 1871 for free-surface flows, enable simulation of flood waves and routing in hydrological models. Subsurface flow in soils and aquifers is governed by Darcy's law combined with the continuity equation to form the Richards equation for variably saturated conditions. Darcy's law describes laminar flow through saturated porous media as \mathbf{q} = -K \nabla h, where \mathbf{q} is the specific discharge (Darcy flux), K is hydraulic conductivity, and \nabla h is the hydraulic head gradient (including pressure and gravity potentials); it was empirically derived by Henry Darcy in 1856 from sand column experiments showing flow rate proportional to head difference and cross-sectional area, inversely proportional to length, under steady-state, homogeneous, isotropic, and saturated conditions with low Reynolds numbers to ensure no turbulence. For unsaturated flow, Richards extended this in 1931 by incorporating capillary effects, where K becomes moisture-dependent K(\theta) (with \theta as volumetric water content) and head h includes matric potential \psi (negative in unsaturated zones), yielding \mathbf{q} = -K(h) \left( \nabla h + \mathbf{k} \right), with \mathbf{k} the unit gravity vector. The full Richards equation emerges by substituting Darcy's law into the continuity equation \frac{\partial \theta}{\partial t} + \nabla \cdot \mathbf{q} = S, resulting in the mixed form \frac{\partial \theta}{\partial t} = \nabla \cdot \left[ K(h) \left( \nabla h + \mathbf{k} \right) \right] + S, or in head-based form \frac{\partial \theta}{\partial t} = \frac{\partial}{\partial z} \left[ K(h) \frac{\partial h}{\partial z} + K(h) \right] + S for vertical one-dimensional flow ( z upward). Derivation assumes isothermal conditions, no air-water interactions, local thermodynamic equilibrium, and a soil water retention curve \theta(h) linking content and potential; hysteresis in retention is often neglected for simplicity, though it can be included in advanced formulations. This equation captures infiltration, drainage, and vadose zone dynamics central to hydrological modeling. Evapotranspiration, a key sink in the water balance, is quantified using the Penman-Monteith equation, which physically combines radiation-driven energy balance with aerodynamic mass transfer. The equation for latent heat flux \lambda E is \lambda E = \frac{\Delta (R_n - G) + \rho_a c_p \frac{(e_s - e_a)}{r_a}}{\Delta + \gamma \left(1 + \frac{r_s}{r_a}\right)}, where \Delta is the slope of the saturation vapor pressure curve, R_n net radiation, G soil heat flux, \rho_a air density, c_p specific heat of air, e_s - e_a vapor pressure deficit, \gamma psychrometric constant, and r_a, r_s aerodynamic and surface resistances, respectively. Derived from the Penman (1948) combination method, it balances available energy \Delta (R_n - G) with drying power \rho_a c_p (e_s - e_a)/r_a, weighted by physiological controls via r_s; Monteith (1965) added r_s for vegetation effects. Assumptions include steady-state canopy conditions, uniform wind over a well-watered reference surface (e.g., grass), and negligible advection; the FAO-56 standardization fixes r_s = 70 s m^{-1} for clipped grass and computes r_a = 208 / u_2 (with wind speed u_2 at 2 m), enabling reference evapotranspiration ET_o in mm day^{-1} as ET_o = \frac{0.408 \Delta (R_n - G) + \gamma \frac{900}{T+273} u_2 (e_s - e_a)}{\Delta + \gamma (1 + 0.34 u_2)}, where T is air temperature. This form is widely adopted for its physical basis and minimal parameterization in hydrological models.

Solution methods

Hydrological models often employ analytical methods to derive closed-form solutions for simplified scenarios where governing equations can be solved explicitly, providing insights into fundamental processes without extensive computation. One prominent example is the kinematic wave approximation, which neglects inertial and terms in the Saint-Venant equations to model overland and channel flow as a wave propagating at celerity c = \frac{dx}{dt}, where x is distance and t is time; this approach yields analytical hydrographs for uniform rainfall on plane surfaces, as detailed in foundational kinematic wave theory. Such methods are particularly useful for preliminary assessments in ungauged basins or educational purposes, though they assume steady-state conditions and neglect diffusion, limiting applicability to steep slopes with minimal backwater effects. For more complex hydrological systems involving nonlinear partial differential equations, numerical methods are essential to approximate solutions on discrete grids. The discretizes spatial and temporal derivatives on structured grids, using explicit schemes for simple advection-dominated flows or implicit schemes to handle diffusive terms in unsaturated flow equations, enhancing stability for larger time steps. Finite volume methods, conversely, integrate conservation laws over control volumes to preserve and , making them ideal for simulating and routing where discontinuities like shocks may arise. Finite element methods excel in irregular domains, such as heterogeneous watersheds, by approximating solutions with basis functions over unstructured meshes, allowing flexible representation of and variations. These approaches are often combined in hybrid formulations to balance accuracy and computational demands in distributed models. To address the nonlinearities inherent in hydrological equations, such as those describing variably saturated flow, iterative algorithms linearize and solve the system repeatedly until convergence. The Picard iteration scheme, for instance, updates or moisture content in a successive substitution manner for the Richards equation, with modifications like the mixed-form approach improving mass conservation by incorporating both head- and moisture-based formulations within each iteration. Time-stepping algorithms advance the solution temporally, with explicit Runge-Kutta methods of second or higher order providing efficient integration for hyperbolic systems like kinematic waves, offering adaptive error control while maintaining stability under restrictive conditions. These solvers are typically embedded within or element frameworks, iterating until residuals fall below a tolerance threshold, often on the order of $10^{-4} to $10^{-6} for practical simulations. In software implementations of distributed hydrological models, architectures accelerate simulations over large spatial domains by partitioning the into sub-basins processed concurrently on multi-core processors or clusters, reducing runtime from days to hours for high-resolution grids. Stability criteria, such as the Courant-Friedrichs-Lewy (CFL) condition, guide time-step selection to prevent numerical instabilities, requiring \Delta t \leq \frac{\Delta x}{c} where c is the characteristic speed, typically limiting explicit schemes to CFL numbers below 1 in diffusive or wave-propagating contexts like overland flow. These considerations ensure robust performance in operational forecasting tools, balancing computational efficiency with physical fidelity.

Implementation and Evaluation

Calibration techniques

Calibration techniques in hydrological modeling aim to adjust model parameters so that simulated hydrological responses, such as , closely align with observed data from gauged sites. This process is essential for enhancing model reliability in reproducing catchment behavior under varying conditions. Two main approaches are employed: manual trial-and-error , in which modelers iteratively tweak based on qualitative and quantitative comparisons of simulated and observed hydrographs, and automated optimization methods that employ algorithms to systematically minimize discrepancies. Manual methods rely on expert intuition and are labor-intensive but useful for understanding model sensitivities in preliminary stages. In contrast, automated techniques, such as the shuffled complex algorithm (SCE-UA), offer efficiency and reproducibility by performing global searches in high-dimensional parameter spaces. SCE-UA, specifically designed for hydrological applications, combines elements of and competitive to converge on optimal parameter sets. The primary objective of calibration is to minimize error metrics that quantify the difference between observed (O) and simulated (E) values. Widely used metrics include the error (RMSE), which measures the average magnitude of errors in a set of predictions without considering their direction, and the Nash-Sutcliffe efficiency (NSE), which compares model performance to the baseline of using the of observations. NSE is calculated as: NSE = 1 - \frac{\sum_{i=1}^{n} (O_i - E_i)^2}{\sum_{i=1}^{n} (O_i - \bar{O})^2} where n is the number of observations, O_i are individual observed values, E_i are corresponding simulated values, and \bar{O} is the mean of the observed values; NSE values range from -\infty to 1, with 1 indicating a . These metrics guide the optimization process, often through least-squares minimization. Calibration typically proceeds in steps, starting with to pinpoint parameters that most strongly influence model outputs, thereby reducing the number of variables to optimize and mitigating issues like overparameterization. Local sensitivity methods perturb individual parameters, while global approaches, such as variance-based decomposition, evaluate interactions across the full parameter range. Following this, regionalization techniques extend calibrated parameters to ungaged catchments by relating them to physiographic and climatic attributes, as demonstrated in initiatives like the Model Parameter Estimation Experiment (MOPEX), which benchmarks transfer methods across diverse basins. Practical tools facilitate these processes, with the Parameter (PEST) software being a for automated inverse modeling in ; it employs to estimate parameters while incorporating regularization to handle ill-posed problems. A key challenge addressed by such tools is equifinality, where multiple parameter combinations yield comparable fits to data, potentially leading to non-unique solutions; this is managed through multi-objective frameworks that explore trade-offs and constrain plausible parameter sets based on physical bounds. Recent advances incorporate techniques for more efficient , such as knowledge-informed methods that use a few hundred realizations to estimate parameters while incorporating physical constraints, reducing computational demands compared to traditional approaches.

Validation and performance metrics

Validation of hydrological models involves testing to assess their predictive capability beyond the data used for . A common approach is the split-sample test, where the available is divided into and validation periods, typically using data from different hydrological regimes to evaluate temporal transferability. Proxy-basin methods extend this by calibrating the model on one catchment and validating it on a similar but basin, testing spatial transposability. , a variant of leave-one-out cross-validation, involves iteratively excluding subsets of basins or data points to assess robustness across multiple configurations. Key performance metrics quantify model accuracy through statistical comparisons between simulated and observed streamflows. The Nash-Sutcliffe Efficiency (NSE) measures the relative predictive skill of the model against the mean of observed data, defined as \text{NSE} = 1 - \frac{\sum_{t=1}^T (Q_{o,t} - Q_{m,t})^2}{\sum_{t=1}^T (Q_{o,t} - \overline{Q_o})^2}, where Q_{o,t} and Q_{m,t} are observed and modeled flows at time t, \overline{Q_o} is the mean observed flow, and T is the number of time steps; NSE values range from -\infty to 1, with 1 indicating perfect agreement. The Kling-Gupta Efficiency (KGE) decomposes error into correlation (r), bias ratio (\beta), and variability ratio (\alpha), computed as \text{KGE} = 1 - \sqrt{(r-1)^2 + (\alpha-1)^2 + (\beta-1)^2}, offering a more diagnostic assessment than NSE alone by highlighting specific error components. Volume error (VE) evaluates closure as \text{VE} = \frac{\sum_{t=1}^T Q_{m,t} - \sum_{t=1}^T Q_{o,t}}{\sum_{t=1}^T Q_{o,t}} \times 100\%, with values near 0% indicating unbiased long-term . Acceptable thresholds, such as NSE > 0.5 or |VE| < 10%, vary by application but guide overall model adequacy. Graphical tools complement quantitative metrics by visualizing model behavior. Hydrographs overlay simulated and observed to inspect timing and peak flow reproduction during events. Scatter plots of simulated versus observed flows reveal and patterns, with points near the 1:1 line indicating strong performance. Flow duration curves compare the cumulative distribution of flows, assessing the model's ability to capture flow variability from high to low regimes. Best practices emphasize multi-objective evaluation to balance metrics like NSE, KGE, and , ensuring comprehensive assessment across flow regimes and reducing trade-offs in model fit. Addressing is critical, achieved through validation on held-out to confirm rather than mere memorization during .

Uncertainty analysis

Uncertainty in hydrological models arises from multiple sources, including values, input such as measurements, model structure inadequacies, and observational errors in calibration or validation . uncertainty stems from the inherent equifinality in hydrological systems, where multiple sets can produce acceptable simulations of observed . Input uncertainty is often dominated by errors in forcing like rainfall, which can propagate significantly through the model due to nonlinear responses. Structural uncertainty reflects limitations in the model's conceptualization of physical processes, such as simplified representations of subsurface flow or land-atmosphere interactions. Observational uncertainty arises from measurement errors in or used for model evaluation. These sources interact, making comprehensive quantification challenging, as highlighted in reviews of hydrological modeling practices. To quantify parameter uncertainty, Monte Carlo simulations are widely employed, involving random sampling from prior distributions to generate ensembles of model realizations and assess the variability in outputs. This approach allows estimation of prediction intervals by propagating distributions through the model, though it can be computationally intensive for complex models requiring optimizations like to focus on likely sets. The Generalized Likelihood Uncertainty Estimation (GLUE) framework, introduced by Beven and Binley, extends this by using a behavioral/non-behavioral approach: yielding simulations with likelihoods above a are retained to form posterior distributions, from which uncertainty bounds are derived without assuming a single optimal set. GLUE has been applied extensively in to handle equifinality and provide likelihood-based confidence intervals for predictions. Uncertainty propagation is addressed through predictions that generate probabilistic outputs, such as intervals around deterministic forecasts, enabling in water management. Sensitivity analysis methods like the Sobol' indices decompose output variance into contributions from individual parameters or inputs, identifying key drivers of uncertainty; for instance, soil often emerges as highly influential in runoff simulations. These indices, based on , facilitate targeted model improvements by prioritizing parameters with high first-order or total . Recent advances emphasize Bayesian approaches for , integrating prior knowledge with data to update parameter posteriors and produce full predictive distributions, particularly post-2020 with integrations. Bayesian hierarchical models, for example, have been used to post-process ensemble streamflow forecasts, accounting for multiple uncertainty sources in a coherent probabilistic framework and improving reliability over traditional methods. These techniques enhance in real-time applications, such as , by providing calibrated predictive densities.

Applications

Water resources management

Hydrological models play a pivotal role in water resources management by simulating water availability and flows to inform planning and allocation strategies for agricultural, urban, and ecosystem needs. These models enable decision-makers to optimize resource distribution, ensuring sustainable use amid competing demands such as irrigation, domestic supply, and environmental flows. By integrating spatial and temporal data on precipitation, evapotranspiration, and soil moisture, they facilitate long-term strategies that balance human and ecological requirements, particularly in regions facing variability in water supply. In reservoir operation optimization, hydrological models simulate storage dynamics and release schedules to maximize benefits like generation, , and reliability. For instance, physically-based models like SHETRAN-Reservoir couple catchment with operational rules to predict inflows and outflows, allowing operators to adjust releases based on forecasted demands and constraints. Such approaches have been applied to operating rules for large-scale water supply systems, calibrating parameters to mimic real-world behaviors under varying conditions. Irrigation scheduling relies on hydrological models that compute s to determine water requirements and timing of applications, minimizing and enhancing efficiency. water balance models, often based on the FAO-56 framework, track root-zone moisture deficits through inputs like rainfall and outputs like , recommending amounts to maintain optimal conditions. Real-time variants incorporate forecasts and field measurements to adapt schedules dynamically, supporting in water-limited areas. The Soil and Water Assessment Tool () exemplifies model use in , simulating hydrological processes at basin scales to evaluate impacts on quantity and quality. SWAT has been employed to assess best management practices for reducing runoff and nutrient loads, aiding in the planning of agricultural water allocations across diverse watersheds. In scarcity scenarios, models like multi-agent frameworks simulate adaptive allocation among users, prioritizing equitable distribution during droughts by incorporating hydrological forecasts and rules. Integration with economic models enhances cost-benefit analysis in water management, linking hydrological simulations to monetary valuations of uses. Hydroeconomic frameworks, such as those combining reduced-form with optimization, quantify trade-offs between agricultural profits, urban supplies, and services, often revealing net benefits from sustainable policies like limits. For example, in California's Kings Basin, such models estimated $249 million in present-value benefits from managed pumping, including savings and reserves, outweighing revenue losses. Scenario testing with hydrological models evaluates land-use changes, projecting alterations in runoff and recharge to guide urban expansion or agricultural shifts. Tools like simulate multiple futures, such as increased impervious surfaces reducing infiltration, to inform policies that preserve water yields for downstream users. These assessments highlight how converting forests to cropland can decrease by up to 20%, prompting adaptive allocation plans. In the 2020s, hydrological models have supported Sustainable Development Goal (SDG) 6 for through case studies addressing scarcity and equity. In Spain's Basin, a hydroeconomic model integrated benefits via Weighted Usable Area metrics, showing that environmental water markets could boost social welfare by balancing agricultural gains (up to €2,346 million) with river habitat preservation under , informing EU policies. Globally, ensemble models like CESM2-LE projected risks for 35% of vulnerable regions by 2030, exposing 753 million people and guiding in and the Mediterranean. In , models underscored mega-drought impacts, advocating integrated management to close rural- gaps in access, aligning with SDG targets for universal sanitation by 2030.

Flood and drought forecasting

Hydrological models play a critical role in flood forecasting by simulating river routing and inundation in real time, enabling timely warnings and mitigation efforts. Real-time routing models, such as the Hydrologic Engineering Center's River Analysis System (HEC-RAS), are widely applied to predict flood wave propagation along river channels using unsteady flow computations driven by observed and forecasted inflows. HEC-RAS has been operational for real-time spring flood forecasting on the Columbia River, integrating hydraulic simulations with observed data to assess flood extents and depths. Ensemble forecasting enhances reliability by incorporating probabilistic weather inputs, where multiple meteorological scenarios from numerical weather prediction models are fed into hydrological simulations to generate probabilistic streamflow predictions. For instance, the European Flood Awareness System (EFAS) employs the LISFLOOD distributed hydrological model, forced by ensemble forecasts from the European Centre for Medium-Range Weather Forecasts (ECMWF), to provide pan-European flood predictions up to 10 days ahead at 1 km resolution. LISFLOOD simulates rainfall-runoff processes and kinematic wave routing to forecast river discharges, supporting flood hazard mapping across the continent. Similarly, the U.S. National Weather Service (NWS) River Forecast System (NWSRFS) uses conceptual models like the Sacramento Soil Moisture Accounting (SAC-SMA) model within an ensemble framework to produce operational river stage forecasts, aiding flood warnings through 13 River Forecast Centers. Event-based hydrological models are particularly suited for simulating discrete flood events by focusing on rainfall excess and hydrograph generation. Data-driven approaches complement traditional models for rapid, short-lead forecasts in data-rich environments. For drought forecasting, hydrological models generate simulated s to compute indices that quantify deficits in water availability over extended periods. The Standardized Index (SSI) is a key metric derived from modeled or observed time series, standardizing deviations from the mean to identify hydrological drought severity at various timescales, such as 1 to 48 months. By fitting data to a probability distribution (e.g., gamma) and transforming to a standard , SSI facilitates comparison across regions and integration into early warning systems. These models support drought early warning by projecting future anomalies based on forecasts and states, enabling proactive measures like water allocation adjustments. In operational contexts, systems like the U.S. National Integrated Information System incorporate hydrological model outputs to monitor SSI alongside other indices for nationwide alerts. Recent advances in flood forecasting include coupling hydrological models with nowcasting techniques for improved short-lead predictions, particularly for flash floods. Nowcasting integrates radar-based estimates with numerical weather models to provide high-resolution inputs (e.g., 0-6 hours ahead) to models like HEC-HMS, enhancing accuracy in convective storm scenarios as demonstrated in 2024 studies on adaptive atmospheric-hydrologic systems.

Challenges and Advances

Current limitations

One persistent challenge in hydrological modeling is the of high-quality observational , particularly in remote or ungauged basins where ground-based measurements are sparse or absent. This paucity limits model and validation, often leading to uncertainties in input variables like and that can propagate through simulations, with rainfall estimation errors alone contributing up to 30% uncertainty in predictions. Satellite-derived , while increasingly available, introduces additional errors due to limitations, atmospheric interference, and retrieval biases, exacerbating inaccuracies in regions with complex or vegetation cover. Scale-related issues further complicate hydrological modeling, as processes observed at plot or hillslope scales do not translate straightforwardly to or regional levels, requiring upscaling techniques that often oversimplify . For instance, aggregating fine-scale data to coarser grids can distort representations of runoff and infiltration, leading to biased simulations in heterogeneous landscapes. introduces non-stationarity, where historical relationships between variables like temperature and runoff break down, rendering traditional models inadequate for projecting future hydrological responses under shifting regimes. Computationally intensive models, such as fully distributed three-dimensional physically-based simulations, demand substantial resources for high-resolution runs, often making applications or uncertainty analyses infeasible on standard hardware. Parameter equifinality compounds this, where multiple sets yield similar outputs, obscuring and increasing predictive without additional constraints like multi-objective . Incorporating human influences, such as , remains difficult due to challenges in parameterizing dynamic land-use changes and effects on runoff and infiltration, which are often represented simplistically in models. Global hydrological models, calibrated at coarse scales, exhibit biases when applied locally, exhibiting notable biases in simulations in data-scarce regions due to unaccounted sub-grid variability and forcing mismatches.

Recent innovations and future directions

Recent innovations in hydrological modeling have increasingly incorporated -physical approaches to enhance the of data, improving the and accuracy of projections for local systems. These methods combine the mechanistic understanding of physical processes with the of algorithms, such as generative adversarial networks, to bridge gaps between coarse global models and fine-scale hydrological simulations. For instance, a 2024 framework uses statistical-physical adversarial learning to dynamically data while preserving physical consistency, demonstrating improved performance in simulating and runoff patterns across diverse basins. Similarly, post-processors applied to process-based models have demonstrated improved prediction accuracy, with methods outperforming traditional statistical approaches, particularly for extreme events, for operational in 2025 studies. Digital twins represent another key advancement, enabling simulation and management of hydrological basins through virtual replicas that integrate sensor data, hydrodynamic models, and . Developed in 2024, these s provide continuous updates for and water allocation, with applications in river basins allowing for and in near . A 2025 blueprint for digital twins emphasizes their role in optimizing operational efficiency and filling knowledge gaps in complex watersheds, particularly by assimilating live data streams into coupled models. Additionally, large-scale models like mizuRoute have evolved in 2024 to support global simulations, incorporating flexible lake and water modules that enhance model coupling and improve estimates at continental scales. Artificial intelligence techniques are advancing uncertainty reduction in hydrological models by correcting biases in physics-based forecasts and quantifying prediction errors probabilistically. Hybrid physics-AI frameworks, such as those integrating with national water models, have significantly improved forecast reliability and accuracy in predictions, boosting accuracy by 4-6 times for predictions, as shown in 2023 and 2025 evaluations. Looking to the future, integrating data—through crowdsourced observations of stream stages and quality—promises to enrich model inputs, particularly in data-sparse regions, with ongoing projects demonstrating improved via assimilated volunteer measurements. emerges as a prospective tool for tackling complex simulations, with 2025 developments in quantum-enhanced for prediction and across scales potentially accelerating hyper-resolution modeling beyond classical limits. Emphasis on in modeling for developing regions is growing, with 2024 frameworks advocating for principles in simulations to ensure inclusive outcomes that address disparities in access and vulnerability. Emerging trends point toward explainable to foster trust and interpretability in hydrological predictions, with techniques like SHAP enabling spatiotemporal insights into dynamics and discharge in 2025 applications. Coupling hydrological models with socio-economic frameworks is also gaining traction to support , as seen in 2025 integrated models that simulate human decision-making alongside water cycles for sustainable . These directions address prior limitations in model and holistic , paving the way for more robust, equitable hydrological tools.

References

  1. [1]
    A Review on Hydrological Models - ScienceDirect.com
    In this paper, we are going to discuss briefly about variable infiltration capacity model (VIC), TOPMODEL, HBV, MIKESHE and soil and water assessment tool (SWAT) ...
  2. [2]
    Hydrologic Modeling | U.S. Geological Survey - USGS.gov
    USGS hydrologic models predict responses of hydrologic systems to stresses and the fate and movement of solutes and contaminants in water.
  3. [3]
    A Brief History of Hydrology* \ - AMS Journals
    Crawford and Linsley's (1966) work on the Stanford. Watershed Model and Harley et al.'s (1970) MIT. Catchment Model showed that digital computers offered ...
  4. [4]
    State-of-the-art hydrological models and application of the HEC ...
    Feb 2, 2023 · The hydrologic model is a simplified representation of an existing hydrologic system that helps water resources comprehension, forecasting, ...
  5. [5]
    Global hydrological models: a review - Taylor & Francis Online
    The paper reviews and compares 12 known global modelling efforts since 1989, the year the first GHM was published.
  6. [6]
    A short history of philosophies of hydrological model evaluation and ...
    Oct 3, 2024 · This historical review addresses the issues of the evaluation and testing of hydrological models, with a focus on rainfall–runoff models.PHILOSOPHIES OF... · DISCURSION 1: MODEL... · DISCURSION 6: MODEL...
  7. [7]
  8. [8]
    What is Hydrology? | U.S. Geological Survey - USGS.gov
    Hydrologic and water quality mathematical models are developed and used by hydrologists for planning and management and predicting water quality effects of ...Introduction · What is Hydrology? · What Hydrologists Do? · Surface Water
  9. [9]
    Hydrology Models - an overview | ScienceDirect Topics
    A hydrology model is defined as a simplified representation of the terrestrial environmental system that describes the movement and storage of water, ...
  10. [10]
    [PDF] applied-hydrology-ven-te-chow.pdf - we civil engineers
    Applied hydrology. (McGraw-Hill series in water resources and environmental engineering) ... equation) from fluid mechanics is used to apply the physical ...
  11. [11]
    [PDF] Hydrological catchment modelling: past, present and future - HESS
    The Rational Method proposed by Mulvany (1850) is a clear exposition of the concept of time of concentration and its relation to the maximum runoff; it ...
  12. [12]
    The relation of hydrographs of runoff to size and character of ...
    The writer defined a unit graph as the hydrograph of runoff from a given area, due to a one-inch runoff-depth applied in one day or in any other convenient ...
  13. [13]
    Hydrologic modeling: progress and future directions
    May 28, 2018 · Abstract. Briefly tracing the history of hydrologic modeling, this paper discusses the progress that has been achieved in hydrologic modeling ...
  14. [14]
    history of the stanford watershed model - ResearchGate
    The computer software that formed the Stanford Wa- tershed Model evolved over a period from about 1959 through 1966.
  15. [15]
    Kinematic wave modelling in water resources: a historical perspective
    Mar 13, 2001 · The history of the kinematic wave theory and its applications in water resources are traced. It is shown that the theory has found its niche ...
  16. [16]
    [PDF] Chapter 13: Tools for analysis - USDA Forest Service
    Lumped vs. distributed model. The “lumped” watershed on the left has a single value to describe soil, vegetation, and climate conditions in the entire watershed ...
  17. [17]
    [PDF] A review of rainfall-runoff modeling for stormwater management ...
    The hydrologic parameters used in the rainfall-runoff models can be represented in either a lumped or distributed manner. The lumping method averages the total ...
  18. [18]
    8 Hydrological-Hydraulic Modeling
    This book chapter will guide you through the modeling process with the example of the Nauvalisoy catchment. 8.3 The Simplest Model - The Linear Reservoir Model.
  19. [19]
    [PDF] HYDROGEIOS: a semi-distributed GIS-based hydrological model for ...
    Jul 28, 2008 · Recent research revealed the advantages of conceptual semi-distributed models for streamflow estimation, in com- parison to lumped ones (Boyle ...
  20. [20]
    Calibration of SWAT and Two Data-Driven Models for a Data-Scarce ...
    Jan 16, 2019 · In this study, a semi-distributed soil and water assessment tool (SWAT) model is used to simulate streamflow at the headwater of Çarşamba River, ...
  21. [21]
  22. [22]
    Event-Based vs. Continuous Hydrological Modeling with HEC-HMS
    Event-based modeling, employing methods such as the SCS curve number (CN) and SCS unit hydrograph, demonstrates exceptional performance in simulating short-term ...
  23. [23]
    Why continuous simulation? The role of antecedent moisture in ...
    Jun 27, 2012 · The advantage of continuous simulation approaches is that the catchment moisture state prior to the flood-producing rainfall event is implicitly ...Missing: disadvantages | Show results with:disadvantages
  24. [24]
    [PDF] Continuous simulation or event-based modelling to estimate flood ...
    The fundamental difference is whether statistic is applied on the input or output of the model.
  25. [25]
    Quality analysis of statistical and data-driven rainfall-runoff models ...
    This paper provides a qualitative analysis of various rainfall-runoff models. The models which were used in this paper are based on statistical and data-driven ...
  26. [26]
    [PDF] Comparison of Peak Discharge and Runoff Characteristic Estimates ...
    The Rational Method is an empirical relation between rainfall intensity and peak flow that is widely accepted by hydraulic engineers; however, the origin of ...
  27. [27]
    Predicting Runoff Signatures Using Regression and Hydrological ...
    Oct 1, 2018 · This paper explores and compares the prediction of runoff signatures using four approaches: (1) two linear regression approaches (MLR and MLLR) ...
  28. [28]
    None
    ### Summary of ARIMA Models in Hydrologic Sequences
  29. [29]
    Regional regression equations for estimation of natural streamflow ...
    The equations define the statistical relations between streamflow statistics (response variables) and basin and climatic characteristics (predictor variables).
  30. [30]
    [PDF] Hydrological Modeling – Current Status and Future Directions
    Jul 17, 2017 · Mulvany, in year 1850, proposed the Rational Method which is based on the concept of time of concentration and its relation to the maximum ...<|control11|><|separator|>
  31. [31]
    [PDF] THE HBVMODEL - SMHI
    The report describes the structure of the most widely used version of the HBV modeland summarizes national and international experiences from its application.
  32. [32]
    None
    ### Summary of SAC-SMA Model Structure, Advantages, Limitations, and Key Citations
  33. [33]
    Physically based modeling in catchment hydrology at 50: Survey ...
    Jul 30, 2015 · This paper first gives an historical overview (past 50 years) of some of the key developments in physically based hydrological modeling.Introduction · Progress Over Five Decades · Current Challenges · Conclusions
  34. [34]
    MIKE SHE | Integrated Hydrological Modelling Software - DHI Group
    Delve into overland flow, unsaturated flow, groundwater flow, and dynamic channel flow with MIKE SHE's integrated physics-based models.
  35. [35]
    [PDF] EVALUATION OF THE MIKE SHE MODEL FOR APPLICATION IN ...
    MIKE SHE is a first generation of spatially dis- tributed and physically based hydrologic model. (Abbott et al., 1986a,b). MIKE SHE simulates the terrestrial ...
  36. [36]
    ParFlow hydrologic model
    ParFlow is a parallel, integrated hydrology model that simulates spatially distributed surface and subsurface flow, as well as land surface processes.
  37. [37]
    [PDF] Simulating coupled surface–subsurface flows with ParFlow v3.5.0
    The core functionality of the ParFlow model is the so- lution of three-dimensional variably saturated groundwa- ter flow in heterogeneous porous media ranging ...
  38. [38]
    Hydrological forecasting at impact scale: the integrated ParFlow ...
    May 29, 2023 · An important advantage of physics-based IHMs as opposed to lumped hydrological models is that they do not require extensive calibration ...
  39. [39]
    (PDF) A Review on Physically Based Hydrologic Modeling
    Feb 12, 2015 · Physically-based models (Islam 2011 ) are built on fundamental equations of fluid mechanics (mass and momentum balance) and aim to simulate ...
  40. [40]
    Revolutionizing the future of hydrological science - ScienceDirect.com
    ML and DL models yield a significant advantage by directly learning from the input data and identifying hidden relationships that conventional models cannot ...
  41. [41]
    Neural network modeling of hydrological systems: A review of ...
    Sep 9, 2018 · This paper presents a review of procedural steps and implementation techniques used in the development of artificial intelligence models, ...
  42. [42]
  43. [43]
    Predicting streamflow with LSTM networks using global datasets
    Jun 4, 2023 · Recent research has shown that deep learning methods based on Long Short-Term Memory (LSTM) cells outperform process-based hydrological models ...
  44. [44]
    Random Forest Ability in Regionalizing Hourly Hydrological Model ...
    Jul 25, 2019 · This study investigated the potential of random forest (RF) algorithms for regionalizing the parameters of an hourly hydrological model.
  45. [45]
    Hydrological modelling using artificial neural networks - Sage Journals
    This review considers the application of artificial neural networks (ANNs) to rainfall-runoff modelling and flood forecasting.
  46. [46]
    Advancing Hydrology through Machine Learning - MDPI
    This review explores the current state of ML applications in hydrology, emphasizing the utilization of extensive datasets such as CAMELS, Caravan, GRDC, CHIRPS ...
  47. [47]
    Machine learning assisted hybrid models can improve streamflow ...
    Sep 21, 2020 · Hybrid models, integrating PB models with LSTM, improve streamflow simulation, especially in catchments where PB models fail, and show improved ...Introduction · Data and methodology · Results · Discussion and conclusion<|control11|><|separator|>
  48. [48]
    MELPF version 1: Modeling Error Learning based Post-Processor ...
    Sep 23, 2019 · This paper studies how to improve the accuracy of hydrologic models using machine-learning models as post-processors and presents possibilities to reduce the ...
  49. [49]
    Enhancing physically-based hydrological modeling with an ...
    This study proposes a hybrid hydrological modeling framework, which combines a process-based hydrological model with a machine-learning-based reservoir ...
  50. [50]
    Water quality prediction using SWAT-ANN coupled approach
    Coupling SWAT and ANN improved the nitrate prediction accuracy, especially in forested watersheds. •. The hybrid model developed for monthly nitrate load ...
  51. [51]
    Producing Hydrological Projections Under Climate Change: A ...
    Aug 24, 2025 · A range of HNMs, with various levels of complexity, can be used to produce groundwater-inclusive hydrological projections under a changing ...
  52. [52]
    Towards hybrid modeling of the global hydrological cycle - HESS
    Mar 23, 2022 · In this study, we exemplify a hybrid approach to global hydrological modeling that exploits the data adaptivity of neural networks for representing uncertain ...
  53. [53]
    A review of hybrid deep learning applications for streamflow ...
    The main advantage of hybrid deep learning models over standalone models is that they can combine the strengths of different modeling techniques for streamflow ...
  54. [54]
  55. [55]
    A national topographic dataset for hydrological modeling over the ...
    Jul 8, 2021 · Elevations and slopes are the primary inputs for the hydrologic modeling, but additional datasets defining the drainage networks and watersheds ...
  56. [56]
    IMERG: Integrated Multi-satellitE Retrievals for GPM | NASA Global ...
    Feb 26, 2020 · IMERG is available in near real-time with estimates of Earth's precipitation updated every half-hour, enabling a wide range of applications.<|control11|><|separator|>
  57. [57]
    Selecting a Reach Routing Method - Hydrologic Engineering Center
    The Manning's n roughness coefficient should be the average value for the whole reach. This value can be estimated from pictures of streams with known roughness ...
  58. [58]
    Imputing Missing Data in Hydrology using Machine Learning Models
    Jan 10, 2021 · Handling missing values is a critical component of the data processing in hydrological modeling. The key objective of this research is to ...
  59. [59]
    Scale effect challenges in urban hydrology highlighted with ... - HESS
    Jan 15, 2018 · In this paper we propose a two-step investigation framework to illustrate the extent of scale effects in urban hydrology.
  60. [60]
    [PDF] Basic Concepts of Kinematic-Wave Models
    The development of kinematic-wave theory occurred late in the history of the development of the theory of open-channel flow. It is based on the early develop-.
  61. [61]
    1D Saint-Venant Equations - Hydrologic Engineering Center
    1D Saint-Venant Equations Most clear water hydraulic models compute the boundary friction force with a quasi-empirical formula that accounts for channel ...
  62. [62]
    Correspondence between de Saint-Venant and Boussinesq. 1
    The 1D equations have been proposed in 1871 by Adhémar Barré de Saint-Venant (dSV) thereby allowing for the first computations of unsteady free-surface flows; ...
  63. [63]
    4.1 Darcy's Law – Hydrogeologic Properties of Earth Materials and ...
    Darcy's law is the fundamental equation used to describe the flow of fluid through porous media, including groundwater.
  64. [64]
    [PDF] Generalized Richards' equation to simulate water transport in ...
    Abstract. Simulations of water transport in soil are ubiquitous, and the Richards' equation introduced in 1931 is the main tool for that purpose.
  65. [65]
    Chapter 2 - FAO Penman-Monteith equation
    The FAO Penman-Monteith equation is a close, simple representation of the physical and physiological factors governing the evapotranspiration process. By using ...
  66. [66]
    Analytical solution of a kinematic wave approximation for channel ...
    May 29, 2013 · The kinematic wave approach is often used in hydrological models to describe channel and overland flow. The kinematic wave is suitable for ...
  67. [67]
    [PDF] Numerical Methods in the Hydrological Sciences
    This book provides tools for using numerical methods in hydrological sciences, enabling students to solve complex problems and explore literature using  ...
  68. [68]
    Finite volume multiscale finite element method for solving the ...
    Oct 25, 2005 · In this paper we present a numerical method, the finite volume multiscale finite element method (FVMSFEM), for solving the groundwater flow problems in ...<|control11|><|separator|>
  69. [69]
    Comparison of finite difference and finite element solutions to the ...
    Jan 10, 2003 · Numerical solutions to the equation governing variably saturated flow are usually obtained using either the finite difference (FD) method or ...
  70. [70]
    Numerical Solution of Richards' Equation: A Review of Advances ...
    Oct 19, 2017 · The one-dimensional solution methodology developed by 28, which uses modified Picard iterations to improve mass conservation, has become the ...
  71. [71]
    Numerical daemons of hydrological models are summoned ... - HESS
    Oct 12, 2021 · Higher order Runge–Kutta methods (credited to Carl Runge and Wilhelm Kutta) ... time step because it is commonly used with hydrologic models.
  72. [72]
    A two-level parallelization method for distributed hydrological models
    This paper proposes a scalable two-level parallelization method for distributed hydrological models that can use parallelizability at both the sub-basin level ...Missing: Courant | Show results with:Courant
  73. [73]
    Computational Time Step - Hydrologic Engineering Center
    The best way to estimate a computational time step for HEC-RAS is to use the Courant Condition. This is especially important for dam break flood studies.
  74. [74]
    Catchment scale hydrological modelling: A review of model types ...
    In this review, we summarise the different classifications of hydrological model types, and discuss relative advantages and disadvantages of each type of model.
  75. [75]
    Hydrologic Model Calibration in the National Weather Service.
    Two basic approaches exist for the calibration of model parameters. The first is a manual trial-and-error method [Anderson, 2002], and the second involves the ...<|control11|><|separator|>
  76. [76]
  77. [77]
  78. [78]
    PEST: Home
    PEST is a software package and utility programs for environmental modeling, automating calibration and uncertainty analysis of numerical models.Software · Downloads · PEST: The Book · PEST Roadmaps
  79. [79]
  80. [80]
    Time to Update the Split‐Sample Approach in Hydrological Model ...
    Feb 22, 2022 · Hydrological model calibration is a critical model building process that infers key model parameter values from observed system response data.
  81. [81]
    Hierarchical testing of three rainfall–runoff models in small forested ...
    The statistical analysis for level 4 (proxy-basin differential split-sample test) indicates that TOPMODEL performed significantly better than the lumped models ...
  82. [82]
    Implementing a Proxy-Basin Strategy to Assess the Transposability ...
    Our study implemented a Proxy-Catchment Differential Split-Sample (PBDSS) strategy to assess the transposability of the conceptual hydrological model J2000 in ...Missing: jackknife | Show results with:jackknife
  83. [83]
    [PDF] watershed model calibration and validation: issues and procedures
    • Daily flow timeseries (cfs). – observed and simulated daily flow. – scatter plots. • Flow frequency (flow duration) curves (cfs). • Storm hydrographs, hourly ...
  84. [84]
    (PDF) Hydrologic and Water Quality Models: Performance Measures ...
    Aug 6, 2025 · Summary of graphical performance measures for H/WQ model calibration and validation. Purpose Advantages/Disadvantages. Direct comparison.
  85. [85]
    Full article: Regional models of flow-duration curves of perennial ...
    Regional models of flow-duration curves of perennial and intermittent streams and their use for calibrating the parameters of a rainfall–runoff model.
  86. [86]
    Full article: One decade of multi-objective calibration approaches in ...
    Here we review the recent history of multi-objective hydrological calibration and its usefulness towards establishing more faithful and consistent models. The ...
  87. [87]
    (PDF) The hazards of split-sample validation in hydrological model ...
    This paper investigates the issues related to the use of validation in hydrological model calibration. Traditionally, models are calibrated and then assessed on ...
  88. [88]
    Review: Sources of Hydrological Model Uncertainties and Advances ...
    Thus, model building involves trade-offs between the two sources. Consequently, Clark et al. [159] states how the boundary between structure and parameter is ...
  89. [89]
    Understanding predictive uncertainty in hydrologic modeling: The ...
    May 20, 2010 · Generally speaking, hydrologic modeling is affected by four main sources of uncertainty: (1) input uncertainty, e.g., sampling and measurement ...Bayesian Inference Framework · Experiment B: Estimating Input... · Discussion
  90. [90]
    Monte Carlo assessment of parameter uncertainty in conceptual ...
    Two Monte Carlo-based approaches for assessing parameter uncertainty in complex hydrologic models are considered. The first, known as importance sampling ...
  91. [91]
    The future of distributed models: Model calibration and uncertainty ...
    This paper describes a methodology for calibration and uncertainty estimation of distributed models based on generalized likelihood measures. the GLUE ...
  92. [92]
    Technical Note: Method of Morris effectively reduces the ... - HESS
    Jul 24, 2013 · Sobol' sensitivity analysis has proven to be a valuable tool for diagnostic analyses of hydrologic models.
  93. [93]
    [PDF] Sobol', I. M. (1990) Sensitivity estimates for nonlinear mathematical ...
    Here the expansion theorem from [4] is generalized, the Monte Carlo and quasi-Monte-. Carlo algorithms for estimating sensitivity are considered, and the ...Missing: paper | Show results with:paper
  94. [94]
    Multi-Step Ahead Probabilistic Forecasting of Daily Streamflow ...
    An effective way to perform probability forecasting in the field of hydrology is to apply the Bayesian approach due to the benefits of uncertainty ...
  95. [95]
    Bayesian Joint Probability Approach for Post‐Processing Monthly ...
    Jul 31, 2024 · In this paper, we employ the Bayesian Joint Probability (BJP) modeling approach to generate monthly probabilistic forecasts for LN, N, and MN events in ...
  96. [96]
    SWAT | Soil & Water Assessment Tool
    The Soil & Water Assessment Tool is a small watershed to river basin-scale model used to simulate the quality and quantity of surface and ground water.SWAT+ · Software · Docs · ArcSWAT
  97. [97]
    A method to include reservoir operations in catchment hydrological ...
    We present SHETRAN-Reservoir, a physically-based spatially distributed modelling tool to simulate catchment hydrology, including reservoir operations.
  98. [98]
    Developing water supply reservoir operating rules for large-scale ...
    Sep 12, 2024 · Here we design a set of simple reservoir operating rules (with only two calibrated parameters) focused on simulating small water supply ...
  99. [99]
    Soil water balance models for determining crop water and irrigation ...
    Nov 1, 2020 · This study reviews soil water balance (SWB) model approaches to determine crop irrigation requirements and scheduling irrigation adopting the FAO56 method.
  100. [100]
    Real‐Time Irrigation Scheduling Based on Weather Forecasts, Field ...
    Dec 15, 2023 · Irrigation scheduling is mostly based on an agro-hydrological simulation model to evaluate the impacts of different irrigation application plans ...
  101. [101]
    Introduction to Soil and Water Assessment Tool (SWAT) | US EPA
    Apr 20, 2017 · SWAT is a small watershed to river basin-scale model to simulate the quality and quantity of surface and ground water and predict the environmental impact.
  102. [102]
    An adaptive multi-agent model for water allocation under scarcity
    Jul 9, 2025 · This paper presents a novel multi-agent framework for adaptive, fair water allocation under varying hydrological conditions (severe drought, ...
  103. [103]
    Hydroeconomic modeling of sustainable groundwater management
    Mar 6, 2017 · We quantify the costs and benefits of sustainable groundwater management, including energy pumping savings, drought reserve values, and avoided capital costs.
  104. [104]
    [PDF] Hydro-Economic Modeling of Water Resources Management ...
    We conclude that the field of HEM would benefit from developing more operational models and enhancing the integration of commonly neglected variables, such as ...
  105. [105]
    Hydrologic Response to Land Use and Land Cover Change Scenarios
    Apr 7, 2023 · The study intends to evaluate scenarios with an absence of land use conflict and verify if it can contribute to reductions in surface runoff.
  106. [106]
    Hydrologic futures: using scenario analysis to evaluate impacts of ...
    Jul 31, 2012 · Hydrologic modeling coupled with scenario analysis can (1) elucidate hydrologic responses to anticipated changes in land use and (2) improve ...Missing: testing | Show results with:testing
  107. [107]
    Integrating ecosystem benefits for sustainable water allocation in ...
    May 5, 2022 · A hydroeconomic model is developed to analyze water management policies for adaptation to reduced water availability in the Ebro Basin of Spain.
  108. [108]
    The first emergence of unprecedented global water scarcity ... - Nature
    Sep 23, 2025 · To address this, we use data from the CNRM model under two scenarios, SSP2-4.5 and SSP3-7.0. Both scenarios project that 51% of DZD-affected ...
  109. [109]
    Toward Water Security (SDG 6) amidst the Climate and Water Crisis
    Aug 24, 2023 · Achieving water security through the accomplishment of SDG 6 (clean water and sanitation) is a key point of the 2030 Agenda proposed by the UN for sustainable ...Missing: 2020s | Show results with:2020s
  110. [110]
    [PDF] HEC Fact Sheet, River Analysis System, HEC-RAS
    HEC-RAS is currently being used for real-time forecasting of the Spring flood runoff on the Columbia. River. □ Mississippi River - one-dimensional unsteady flow ...
  111. [111]
    Major upgrade of the European Flood Awareness System | ECMWF
    This article introduces the new 6-hourly calibration of the LISFLOOD hydrological model and provides a summary of its performance.
  112. [112]
    LISFLOOD hydrological model
    LISFLOOD is a model that simulates the full water cycle from rainfall to water in rivers, lakes and groundwater. The model simulates large areas such as ...
  113. [113]
    Catchment modeling and initial parameter estimation for the ...
    Description: The system (NWSRFS) of conceptual hydrologic models and other procedures, used in the operational river forecasting program of the United States ...
  114. [114]
    [PDF] A Brief review of flood forecasting techniques and their applications
    Jan 22, 2018 · Deterministic rainfall- runoff (RR) models include components for the various hydrological and related processes, such as precipitation,.
  115. [115]
    Research papers A hybrid model coupling process-driven and data ...
    This paper proposes a hybrid model for real-time flood forecasting that couples process-driven hydrological models (HMs) with data-driven models (DDMs).
  116. [116]
    [PDF] AN OVERVIEW OF DROUGHT MONITORING AND PREDICTION ...
    Drought monitoring and prediction systems (DMAPS) are reviewed for research and decision-making, providing early warnings at regional and global scales.
  117. [117]
    Coupling WRF with HEC-HMS and WRF-Hydro for flood forecasting ...
    Sep 23, 2024 · The results of this study can help establish an adaptive atmospheric–hydrologic coupling system to improve flood forecasting for different watersheds and ...
  118. [118]
    Scale issues in hydrological modelling: A review - Wiley Online Library
    It is argued that upscaling typically consists of two steps: distributing and aggregating. Conversely, downscaling involves disaggregation and singling out.
  119. [119]
    Understanding non-stationarity of hydroclimatic extremes and ...
    Aug 2, 2023 · Non-stationarities introduced by climatic and anthropogenic changes could lead to substantial errors in estimating return levels of extreme.
  120. [120]
    Equifinality and Flux Mapping: A New Approach to Model Evaluation ...
    Jun 26, 2019 · The operational definition of equifinality in hydrological literature is in fact parameter equifinality. Equifinal parameters are uncertain.Facets of Model Equifinality... · Flux Mapping: An Approach to... · Discussion
  121. [121]
    A critical review of existing hydrologic and water quality models
    Apr 1, 2023 · We conducted a review of 21 different models commonly used for modeling hydrology (8), water quality (6) or both (7) at the watershed scale.<|control11|><|separator|>
  122. [122]
    Global hydrological parameter estimates to local applications
    Mar 11, 2023 · HIGHLIGHTS. Global parameter estimates with global forcing can produce acceptable local discharge simulations in our study area.Hydrological Model And Setup · Results · Discussion
  123. [123]
    Statistical‐Physical Adversarial Learning From Data and Models for ...
    Jun 10, 2024 · Here, we develop a dynamic data-driven downscaling (super-resolution) method that incorporates physics and statistics in a generative framework ...
  124. [124]
    Hybrid approaches enhance hydrological model usability for local ...
    Apr 30, 2025 · Hydrological models are essential for predicting water flux dynamics, including extremes, and managing water resources, yet traditional process ...Results · Hydrological Regime As A Key... · Methods
  125. [125]
    Digital twinning of river basins towards full-scale, sustainable and ...
    Dec 13, 2024 · This article outlines a framework for building widely applicable digital-twin basins and addressing the main obstacles.
  126. [126]
    Blueprint conceptualization for a river basin's digital twin
    Feb 13, 2025 · The core objectives of the digital twin include optimizing operational efficiency, identifying uncertainties, bridging crucial knowledge gaps, ...ABSTRACT · INTRODUCTION · METHODOLOGY · CONCLUSION
  127. [127]
    A Flexible Framework for Simulating the Water Balance of Lakes ...
    May 6, 2024 · In this work, we present the implementation of different water balance models of lakes and reservoirs into mizuRoute, a vector-based routing ...
  128. [128]
    A Hybrid Physics–AI Model to Improve Hydrological Forecasts in
    The hybrid physics-AI model uses a National Water Model (NWM) and a deep learning model to estimate streamflow forecast errors, improving forecast reliability.
  129. [129]
    AI Improves the Accuracy, Reliability, and Economic Value of ...
    Jun 19, 2025 · Here we present a novel hybrid framework that integrates AI-based machinery termed Errorcastnet (ECN) with the National Water Model (NWM)
  130. [130]
    [PDF] Integrating Citizen Science Data into Hydrological Research and ...
    Citizen science projects can use satellite imagery to complement ground-based observations, offering insights into large-scale trends and anomalies.
  131. [131]
    A New Quantum Computing Python Package for Irrigation Demand ...
    Oct 8, 2025 · In this study, we developed QuantumIrrigation—a novel Python package that integrates quantum computing with deep learning to predict ETo and ...
  132. [132]
    Flow Across Scales with a Quantum Computing Boost | Feature
    Jan 17, 2025 · A potential quantum advantage for fluid dynamics simulations from molecular to atmospheric scales with a new formula for success.
  133. [133]
    Integrating equity and justice principles in water resources modeling ...
    Sep 20, 2024 · Integrating equity and justice principles in water resources modeling and management. Seleshi Yalew, Pooja Prasad, Marloes Mul and Pieter van der Zaag.
  134. [134]
    Explainable AI for Interpreting Spatiotemporal Groundwater ...
    Oct 13, 2025 · Machine learning with SHAP enhances insight into large-scale hydrological systems through detailed spatiotemporal interpretations ...
  135. [135]
    Integrating human decision-making into a hydrological model to ...
    May 25, 2025 · This study emphasizes the need for socio-economic integration in environmental modeling to develop more effective and sustainable watershed ...