21,120
Views
196
CrossRef citations to date
0
Altmetric
Original Articles

Global hydrological models: a review

Revue des modèles hydrologiques globaux

&
Pages 549-565 | Received 20 Jun 2013, Accepted 30 Apr 2014, Published online: 10 Mar 2015

Abstract

Global hydrological models (GHMs) have effectively become a separate research field in the last two decades. The paper reviews and compares 12 known global modelling efforts since 1989, the year the first GHM was published. Structure, strengths and weaknesses of individual models are examined, and the objectives of model development and their initial applications are documented. Issues such as model uncertainty, data scarcity, integration with remote sensing data and spatial resolution are discussed.

Editor D. Koutsoyiannis

Résumé

Les modèles hydrologiques globaux (MHG) sont en fait devenus un domaine de recherche distinct au cours des deux dernières décennies. Cet article examine et compare 12 démarches de modélisation globale connues depuis 1989, année de publication du premier MHG. La structure, les forces et faiblesses de chaque modèle ont été examinées et les objectifs de développement des modèles et leurs premières applications ont été décrits. Des questions telles que l’incertitude du modèle, la rareté des données, l’intégration des données de télédétection et la résolution spatiale ont été discutées.

INTRODUCTION

The continuing debate on climate change (e.g. Kundzewicz and Stakhiv Citation2010, Huard Citation2011) and other global drivers of change highlights the interdependence of various earth systems and the need for integration of those systems into global simulation models (Wilby Citation2010). There is an impact on regional and, hence, global climate, from changes in soil moisture and terrestrial evapotranspiration (Munro et al. Citation1998, Koster et al. Citation2003, Koster et al. Citation2004, Seneviratne et al. Citation2010, Dirmeyer Citation2011), and hence river discharge which in turn impacts on sea characteristics (Milly et al. Citation2010). Land-use changes upstream affect the hydrology and water quality thousands of miles downstream (Freeman et al. Citation2007). Global hydrology is closely linked with the nutrient cycle (cause of eutrophication of coastal zones) (Foley et al. Citation2005, Rabalais et al. Citation2009, Fekete et al. Citation2010) and the carbon cycle (impacting on the climate). The changes in these global cycles eventually have social and economic implications. Due to globalization, virtual water trade, for example, has become an important factor of both the global water cycle and food security (Islam et al. Citation2006). While the scientific community has long been aware of, and has dealt with, impacts of climate on hydrology, the feedback has only recently started to receive attention. All this leads to the notion of the ‘global water system’ (Alcamo et al. Citation2008, Alcamo Citation2009), in which the global water flow is connected to other systems through physical relationships, economics and institutions. This system is further complicated by the interference from humans through water storage and withdrawals (Rost et al. Citation2008b).

Models that attempt to simulate global hydrology and associated processes are similar to numerous stand-alone hydrological models and to hydrological components of the general circulation models (GCMs). However, they differ in the detail of description of processes, parameter estimation approaches, time scales, and spatial resolution of input data and simulations (Haddeland et al. Citation2011). The stand-alone models are usually applied at the basin scale, or a smaller catchment scale, and have many parameters that need to be calibrated or estimated regionally. Examples of such models include the Soil and Water Assessment Tool (SWAT; Neitsch et al. 2002), the Hydrological Simulation Program-Fortran (HSPF; Bicknell et al. 1997) and the Hydrologiska Byråns Vattenbalansavdelning (HBV; Lindström et al. 1997). Some of these models can, in principle, be applied at the global scale, but, due to severe information constraints, this never happens in practice. The hydrological models that are a part of GCMs are usually land surface schemes (LSSs) that simulate the energy balance at soil, atmosphere and vegetation interfaces at finer time scales (often hours), and do not have a flow routing component. Examples of such models include the Biosphere–Atmosphere Transfer Scheme (BATS; Dickinson et al. 1986), the Simple Biosphere Model (SiB; Sellers et al. 1986) and the Joint UK Land Environment Simulator (JULES; http://www.jchmr.org/jules/index.html).

The actual global hydrological models (GHMs) have few calibratable parameters and are calibrated either at eco-region, climatic-region or large river basin scales (Vörösmarty et al. Citation1989, Döll et al. Citation2003, Widén-Nilsson et al. Citation2007). Some models, such as the Water Balance Model – Water Transport Model (WBM-WTM; Vörösmarty et al. 1998), are not calibrated per se but have an adjustment factor to tune them. The model Water – Global Analysis and Prognosis (WaterGAP; Alcamo et al. Citation2003, Döll et al. Citation2003) is a combination of calibration and tuning. It is first calibrated with a single parameter against streamflow. The basins that underestimate or overestimate flows are then tuned by two adjustment factors, runoff and discharge correction (Döll et al. Citation2003). The spatial resolution of GHMs is defined by the resolution of available global climate input data. GHMs are relatively new and have emerged in the last two decades. Lately, there has been increasing activity in this field and a concerted effort is evolving (Lawford et al. Citation2004, Döll et al. Citation2008). Also, GHMs are becoming more complex and resolute as more functionality is added to them and finer global spatial datasets are becoming available. Issues of the sensitivity of models to varying spatial and temporal scales of input and output data, measuring modelling uncertainty and coupling GHMs with other models have become prominent (Döll et al. Citation2008, Voß et al. Citation2008).

The explosion of global data availability from satellites in the last two decades (Tang et al. Citation2009) has had an influence on the development of GHMs. Although some GHMs have been applied at a range of scales, they are, as a rule, built for global-scale studies. They would not be the preferred choice in basin-scale applications, due to the coarser resolution of GHMs at present, and the fact that there is a large family of hydrological models that have been designed for this purpose. However, GHMs may provide valuable spatial and temporal estimates of global water resources, and help to analyse possible projections/scenarios of changes of those estimates; GHMs have been built effectively for this purpose. Global estimates obtained through GHMs could be an improvement over those simply based on the statistical analysis of ground-based observed data, which, at a global scale, remain limited and, hence, contain a lot of uncertainty (Rodda Citation1995). An even greater use of GHMs is revealed when they are linked with other models, i.e. describing global economy, ecology, trade, biodiversity, energy balance, land-use change, climate change, crop growth and other development issues/components related to water (Vörösmarty et al. Citation2000, Islam et al. Citation2006, Alcamo Citation2009). The requirements from the GHMs depend upon the demands of such associated models.

To the best of the authors’ knowledge, a single source/compilation of GHMs does not exist. Recently, Trambauer et al. (Citation2013) examined GHMs in the context of drought forecasting in Africa. This paper extends the discussion to the broader context of developing simulation tools for global food trade, agriculture and economy, and examines constraints and development trends of GHMs, including those of uncertainties, input data quality and integration with remote sensing data.

EXISTING GLOBAL HYDROLOGICAL MODELS—DESCRIPTION

Over the last few decades, many isolated efforts were made to simulate the global hydrological cycle. A review of the existing literature suggests that there are at least 12 GHMs (or tools that can be interpreted as such) at present. lists these GHMs in chronological order of their development and includes details of the main developers, their objectives for developing the model and some example applications. summarizes the technical details of these models.

Table 1 Origin and known applications of existing global hydrological models.

Table 2 Technical features of existing global hydrological models.

An initial effort in the development of GHMs was made by Vörösmarty et al. (Citation1989), who developed the Water Balance Model (WBM) and linked it with the Water Transport Model (WTM). The WBM predicts soil moisture (SM), evapotranspiration (ET) and surface runoff at grid level, and the WTM routes the discharge generated in the grid cell through the channel using a linear reservoir model (Vörösmarty et al. Citation1989). The field capacity (FC) and, hence, water availability are linked to the vegetation type and soil at grid-level resolution. The WBM-WTM model relies on a narrow range of parameters based on those published in the literature and then iterates annually until a dynamic steady state is achieved (i.e. soil moisture, ET, runoff and snowmelt come within an acceptable level of difference for successive runs). It has been modified over the years to include irrigation and reservoirs, accounting for the irrigated and non-irrigated fraction, reservoir operations (WBMplus class), effect of permafrost (Pan-Arctic Water Balance Model (PWBM) and HydroDynamic Model (HDTM 1.0), along with the multi-layer soil moisture routine (HDTM 1.0; http://www.wsag.unh.edu/wbm.html). The parameter values in the WBM are assigned a priori and are not calibrated. In the Macro Probability Distribution Model (Macro-PDM; Arnell Citation1999), developed in 1998, streamflow is simulated as direct runoff and delayed runoff. A similar vegetation classification as in WBM is used, but the soil moisture FC is statistically distributed within a grid to account for sub-grid variations. The ratio between ET and potential evapotranspiration (PET) is described by a linear function of average soil moisture in a grid cell. In this model, the values of parameters are set from a literature review or previous model applications. Six out of 13 parameters are globally uniform. Around the same time, the Max-Planck Institute (MPI) in Germany developed a model that coupled the European Centre Hamburg Model (ECHAM) GCM with a hydrological discharge (HD) model (Hagemann and Dümenil Citation1997). This model is known as the Max Planck Institute-Hydrology Model (MPI-HM). In this model, vegetation in each grid cell is partitioned based on vegetation maps, as developed by Matthews (Citation1983), and the fraction of such vegetation dependent on the available water content in the root zone (decreases below 40% of FC). The streamflow and surface runoff are simulated as a cascade of linear reservoirs, whereas the baseflow is described by a single reservoir. The Global Water Availability Assessment model (GWAVA; Meigh et al. Citation1999), developed to examine issues of water scarcity, also used a probability distribution to simulate soil moisture storage. This model includes water demand based on population, livestock, irrigation schemes and industrial water use. It uses four simplified land cover categories (trees, bushes, grass and bare soil). Groundwater is first explicitly calculated by using potential yield from a well and estimating borehole density in each grid. It is then compared to groundwater recharge calculated through the surface flow model and a conservative value of the two, based on aquifer classification, is used.

All the above models treat soil as a single layer. The Variable Infiltration Capacity model (VIC; Nijssen et al. Citation2001b), developed in 2001, considers two layers of soil and includes sub-grid variability in land surface vegetation, soil moisture storage capacity and precipitation. The VIC model can run either as an energy and water balance model or just as a water balance model. Multiple land covers (13) including bare soil are considered in the VIC model. It also accounts for the top vegetation cover. The second layer is modelled as a nonlinear storage. Snow is simulated by a single layer energy and mass balance model. A separate river routing model is used. The Land Dynamics model (LaD; Milly and Shmakin Citation2002) was developed to improve an older water and energy balance model by Manabe (Citation1969) by adding sensible heat storage on land, groundwater storage and stomata resistance. Grid cells are treated as either glaciated or non-glaciated. Water storage is considered in snow, glacier ice, root zone and groundwater. Runoff is generated after the soil reaches its water-holding capacity, and is then passed through a groundwater reservoir (with some residence time) before becoming streamflow. The PCRaster GLOBal Water Balance model (PCR-GLOBWB; van Beek and Bierkens Citation2008), developed in 2004, broke up the soil layers into three storage ‘buckets’, each contributing to the streamflow in the form of surface runoff, interflow and baseflow, respectively. The surface vegetation is simplified into three categories—natural, rainfed and irrigated—which are further classified as short (retrieving moisture from the top soil layer) or tall vegetation (retrieving moisture from the bottom layer). Evapotranspiration is broken up into transpiration by vegetation and evaporation from bare soil.

The Water – Global Analysis and Prognosis model (WaterGAP; Alcamo et al. 2003, Döll et al. Citation2003) and the Lund-Potsdam-Jena managed Land model (LPJmL; Bondeau et al. Citation2007) are two GHMs that were developed in Germany during the past decade. While the WaterGAP model puts more emphasis, at least at present, on water demand simulation, the LPJmL model is more detailed in vegetation and crop simulations. The WaterGAP model is made up of two modules—the GHM per se and the global water use model. The latter incorporates human water consumption (in the form of domestic, manufacturing, thermal power production, livestock and irrigation use). The GHM model considers soil as a single layer and calculates the vertical and lateral water balance within each grid. Surface and subsurface runoff are transported as a fast response, and the baseflow (i.e. contribution from groundwater) is transported as a slow response. The lateral movement of runoff is described through multiple storages representing local and global lakes, reservoirs, wetlands and streams. The LPJmL model was developed by adding a hydrology component to an existing dynamic global vegetation model (DGVM). The vegetation routine of the model simulates photosynthesis, evapotranspiration, plant respiration and the carbon cycle. It allows for nine natural plant types and 12 crop types (both rainfed and irrigated). There can be one stand of natural vegetation and multiple stands of crop types within each grid. Some of the inputs for the vegetation are dynamically generated within the model. The hydrological routine calculates water balance for each stand of the grid and each grid flows into another via a river routing module (Rost et al. Citation2008a). The evapotranspiration is broken up into productive water consumption (plant transpiration) and non-productive consumption (interception loss and evaporation). Soil storage is represented by two ‘buckets’.

The Water And Snow balance MODeling system (WASMOD-M; Widén-Nilsson et al. 2007), originating from WASMOD (Xu Citation2002), uses a minimum number (4–6) of calibrating parameters. WASMOD-M runs at monthly time steps and separates runoff into fast and slow components for each grid. It considers rain, snowmelt, snow accumulation and evapotranspiration for each grid within its routine. The parameters in the model are regionalized to simplify model application. It has six parameters, of which five can be calibrated and one is fixed (Widén-Nilsson Citation2007). The H08 model (Hanasaki et al. Citation2008a, Citation2008b) is a more detailed and integrated system, made up of six modules—land surface hydrology, river routing, crop growth, reservoir operation, environmental flow requirements and anthropogenic water withdrawal. The land-surface hydrological module calculates both the hydrology balance and energy balance. It uses the ‘leaky-bucket’ concept to model soil moisture (Hanasaki et al. Citation2008a). The river routing module uses the Total Runoff Integrating Pathways model (TRIP; (Oki and Sud Citation1998). The TRIP model is a network of integrated pathways providing information on lateral water movement over land. The crop routine uses heat unit theory (Barnard Citation1948, Phillips Citation1950) and is based on the Soil and Water Integrated Model (SWIM; Krysanova et al. 1998). The basis of the heat unit theory is that each plant has its own range of temperature within which it grows. SWIM is a river basin model developed by the Potsdam Institute for Climate Impact Research (PIK), Germany, to simulate hydrology, sediments, nutrients and carbon movement, along with plant growth and yield. Nineteen crop types are used and large reservoirs (with a storage capacity greater than 109 m3) are incorporated.

The ISBA-TRIP model (Alkama et al. Citation2010) was developed by combining two models—the modified land surface model by Noilhan and Planton (Citation1989) and the surface runoff pathway network from TRIP, by Oki and Sud (Citation1998). The heterogeneity within a grid is described by using a tile approach that divides a grid into a series of sub-grid patches based on land use and soil properties; partial grid saturation is also accounted for. The model calculates the surface energy and water budget for each grid with the soil represented by three layers. TRIP routes the daily discharge simulated by ISBA into streamflow.

COMPARISION OF MODEL FEATURES

All GHMs run in a grid format. In most GHMs, the spatial resolution is 0.5 degrees (just over 3100 km2 per grid cell at the Equator). The grid format and resolution are dictated by the availability of global meteorological data and computational resources. The gridded format makes it convenient to manipulate other available input data, which are also usually available in grid format (e.g. originating from remote sensing (RS) data). With faster computational devices and availability of finer resolution input data, higher resolution GHMs are being developed, such as a 5-min version of WaterGAP (http://www.massentransporte.de/index.php?id=303), 6-min version of PCR-GLOBWB (called PCR-GLOBWB 2.0: http://www.globalhydrology.nl/models/pcr-globwb-2-0/) and the latest version of WBMplus that can run at a range of resolutions dictated by the underlying gridded network (45″, 90″, 3′, 5′, 6′, 15′ – B. Fekete, City University of New York (CUNY), pers. comm.).

Most of the GHMs have a temporal resolution of 1 day despite the fact that the available input data normally have a monthly time step. Historical data covering the full 20th century, for example, are only available at a monthly temporal resolution, which necessitates statistical downscaling into a daily temporal scale. With improvement of meteorological reanalysis data, the models have started using more of these as inputs. The fourth generation of ECMWF reanalysis (ERA) daily datasets (from 1979 to date) are being developed; ERA data have already been used in some GHMs (Wada et al. Citation2010) and in model intercomparison studies (Weedon et al. Citation2011).

Only five models—WBMplus, WaterGAP, PCR-GLOBWB, LPJmL and H08—explicitly consider reservoir storage. Out of these, PCR-GLOBWB does not simulate irrigation. Some models consider soil layers as a single storage unit (Macro-PDM, MPI-HM, WASMOD-M, H08 and WaterGAP), whereas others consider multiple storage layers (PCR-GLOBWB: two layers, VIC: two layers, LPJmL: two layers, ISBA-TRIP: three layers, and LaD: five layers).

Only H08 and LPJmL have an explicit crop growth model. Although all the models can handle changes in meteorological data due to climate change (such as precipitation and ET due to changes in temperature), only the models that have crop growth and/or vegetation models can deal with the impact on plant physiology due to changes in carbon dioxide concentration and increase in temperature. From that perspective, LPJmL and H08 may be well suited to evaluate the impact of climate change scenarios on crop yield (and hence on hydrology). The LPJmL modle simulates natural vegetation (along with agriculture), and the proportion of different vegetation classes in each grid is based on relative advantage in terms of light, water, space and other environmental factors (Gerten Citation2013).

A few models, such as VIC, LaD, WaterGAP and LPJmL, have a detailed land-use classification. Other models have used a simplified classification, for example, WBM-WTM uses only three categories (i.e. forest, grassland and shrubland); and the newer version, WBMplus, accounts for multiple categories (e.g. Wisser et al. Citation2010 handle four categories). The model can simulate up to 32 of the vegetation classes of the Terrestrial Ecosystem model (Melillo et al. Citation1993). The current version of WBM can read the parameters related to the vegetation directly as input from an external source and, hence, the model can simulate as many classes as formed by the different parameter combinations (B. Fekete, pers. comm.). Macro-PDM uses only forest and grassland categories for land use, whereas PCR-GLOBWB uses three categories: natural vegetation, rainfed crops and irrigated crops, which are further divided into short and tall vegetation (to simulate ET from different layers of the soil profile).

Models also differ in terms of how they account for energy balance, sub-grid effects and soil stratification. About half of the models reviewed (MPI-HM, PCR-GLOBWB, VIC, LaD, H08 and ISBA-TRIP) have their own energy balance module. Some models treat a grid as homogenous with respect to infiltration, climate data and vegetation, while others account for sub-grid effects to describe heterogeneity within a grid cell (VIC, ISBA-TRIP). Some consider root-zone depth as the height of one of the soil layers (such as ISBA-TRIP), whereas others have a fixed depth of the first layer (PCR-GLOBWB: 0.3 m, LPJmL: 0.5 m).

The main output of all the models is streamflow, which is derived from partitioning precipitation into evapotranspiration, soil moisture, fast and slow flow (although the terms used may be different, e.g. surface runoff, lateral flow and baseflow). The VIC, WaterGAP, LPJmL and H08 models also look at crop water requirements. These are the models that also have a detailed land-use classification (). All models provide output at grid level (mostly at 0.5° resolution at present). These grid-level outputs are aggregated to river basin scale, based on the number of grids that fall within a river basin boundary. None of the models uses an incremental drainage sub-basin discretization directly.

Global hydrological models are often developed to support simulation of certain phenomena which are not always hydrology per se (). Consequently, they have certain areas of preference. For example, LPJmL has a vegetation model along with the hydrological model. The LaD, ISBA-TRIP and VIC models include hydrology, but also focus on energy balance, and the H08 model focuses on crop growth and energy balance. WaterGAP, WBMplus (irrigation only) and GWAVA also consider water use. Although the use of any PET equation in a hydrological model may be seen as an energy balance, in this study, unless the model has an energy balance built into its framework, it is not considered. None of the models reviewed incorporates land-use change models, i.e. that simulate changes in land use due to development (for a review of land-use change models, see Agarwal et al. Citation2002). Also, the GHMs do not consider the impact of market forces or economics (in terms of human interventions) on hydrology.

ISSUES AND TRENDS IN GHM DEVELOPMENT

Attempts are being made to coordinate the development of GHMs. The Global Water System Project (GWSP), along with the Integrated Project Water and Global Change (WATCH) (http://www.eu-watch.org/templates/dispatcher.asp?page_id=25222705), invited teams developing GHMs to participate in a model intercomparison project (http://www.eu-watch.org/nl/25222736-Global_Modelling.html). Seven different GHMs were forced with the same set of climatic inputs, and their results were compared at both a continental scale and the scale of major river basins. The study concluded that there was limited agreement between models on global water projections. For example, the range of predicted global runoff from different models was about 45% of the mean simulated runoff (Haddeland Citation2011). Similarly, Gudmundsson et al. (Citation2012) applied an ensemble of nine large-scale hydrological models in Europe and found a large spread in model accuracy including high variation in the low runoff percentile.

Another 4-year initiative, the Inter-Sectoral Impact Model Intercomparison Project (ISI-MIP; http://www.pik-potsdam.de/research/climate-impacts-and-vulnerabilities/research/rd2-cross-cutting-activities/isi-mip) is currently underway. Several GHMs, including VIC, H08, WaterGAP, MacPDM, WBM, MPI-HM and PCR-GLOBWB, along with other global impact models, are included in this project, where new climatic and socio-economic scenarios developed as part of the Fifth Assessment Report of the IPCC (Intergovernmental Panel on Climate Change) are used as forcing, and their possible impact on different sectors (such as agriculture, water, ecosystems, infrastructure and health) are being evaluated.

Focusing on uncertainties

In the WATCH study, although, at the global level, the deviation between the models of the simulated total annual flow was not very large—a coefficient of variation (CV) of about 0.09 (Voß et al. Citation2008), there were larger deviations in colder (polar, without Antarctica: CV ≈ 1.0) and drier regions (Africa and Australia: CV ≈ 0.41). Also, despite the fact that there was a closer match in the total flow, there was a very wide variation in the soil moisture (CV ≈ 0.33) and evaporation products (CV ≈ 0.27) suggesting the importance of looking into structural differences in the models. While some models simulated an increase in total global freshwater flow for the period 2071–2100, others showed a reduction (Voß et al. Citation2008), the range being from −40 to +30%. The study showed high uncertainty in the intermediate processes. Wada et al. (Citation2013) used seven GHMs to assess the impact of climate change on future irrigation water demand. They concluded that high uncertainty, until at least mid-century, is due to GHMs rather than to GCMs.

Uncertainties in GHMs have still not been sufficiently analysed or emphasized. Most of the time the validity of model outputs of GHMs is asserted by comparing the results with the output of other GHMs. Sometimes the assessment is made based on anecdotal terms, for example, “... on the whole the selected parameters produce a reasonably good representation of monthly hydrological regimes across Europe.” (Arnell Citation1999); or “… we see that WBM runoff is ‘well behaved’ in relation to the precipitation data used ...” (Fekete et al. Citation2002). There are four types of uncertainties in hydrological modelling: those associated with model inputs, model structure, parameter values and observed data. Palmer et al. (Citation2008) analysed the uncertainties associated with WaterGAP. Based on this and other analyses (such as Kaspar Citation2004), it is concluded that the uncertainties due to inputs are higher than the parameter uncertainties. Using ‘goodness’ of the model calibration as an indicator, they found total uncertainty in streamflow to be medium in 43% of large river basins, lower in 32% and higher in 25%. Widén-Nilsson (Citation2007) compared WASMOD-M and five other global models and found a high volume error (i.e. error in the river runoff) in all the models, although the regions of high error were not consistent. The volume error mostly ranged from −65 to 50%. Biemans et al. (Citation2009) found a multiplier effect of precipitation uncertainty in predicting streamflow at the global scale. An average of 30% uncertainty in the precipitation data, when used as an input with LPJmL, resulted in an average 90% uncertainty in streamflow.

One of the reasons for high uncertainty in model inputs is data scarcity, or inconsistency in the input data (Arnell Citation1999, Döll and Siebert Citation2002, McMillan et al. Citation2012, Kauffeldt et al. Citation2013, Mulligan Citation2013, Müller Schmied et al. Citation2014). Data scarcity is greater in developing regions (Schuol and Abbaspour Citation2006), and is due to well-known problems of weak observational infrastructure, gaps due to missing data, and unwillingness to share data. Also, low quality ‘disinformative data’ (Kauffeldt et al. Citation2013) often provide a false sense of data richness, effectively being responsible for a large part of model uncertainty.

Integrating with remote sensing

Continuously progressing remote sensing (RS) technologies can help reduce uncertainties associated with inputs and observations. The future of GHMs is to a large degree associated with the development of RS technologies and public availability of RS data. Most of the GHMs are already using some satellite products, such as solar radiation, digital elevation models (DEMs), land use, etc., as their input data. Although (at least at present) ground-based measurements are assumed or perceived to be more accurate than satellite-based measurements, it is obviously impossible to cover the entire world with the former. All GHMs (other than WaterGAP) are tuned by first converting flow from the grids into streamflow and then validating against the observed data from, generally, large gauged rivers. For example, Macro-PDM uses 31 large river basins for validation, MPI-HM uses 35, and PCR-GLOBWB uses annual streamflow time series from 99 large river basins. VIC, LaD, WaterGAP (which is also calibrated), GWAVA, WASMOD-M, H08 and ISBA-TRIP use 26, 82, 724, 96, 663, 37 and 33 river basins, respectively. However, the number of basins used for validation/calibration for the same model may vary based on the study, e.g. Döll et al. (Citation2003) use 724 calibration stations for WaterGAP, whereas Hunger and Döll (Citation2008) use 1235 stations for the same model. Clearly, such limited tuning/calibration introduces uncertainty in model outputs. There are also inaccuracies inherent in the land-based streamflow measurements. Also, as most of the large rivers have been highly modified (Döll et al. Citation2009), it is imperative for future GHM modelling activities to include river alterations and water withdrawals (van Beek et al. Citation2011). In addition, there has been a significant decrease in the number of river gauging stations worldwide since the early 1990s, and the existing gauging network monitors streamflow from only approximately 50% of the land mass (Fekete and Vörösmarty Citation2007).

The future of hydrological modelling also rests in using satellite data (directly or derived) for calibration. Significant progress has been made in satellite-based data acquisition in the last decade. With the launch of the European Space Agency’s (ESA) Soil Moisture and Ocean Salinity (SMOS) satellite in November 2009, the first global soil moisture maps have become available. The satellite uses large microwaves (L-band) to measure ‘brightness temperature’ through which surface soil moisture is calculated every three days. The data are being validated by a network of ground-based measurements (International Soil Moisture Network), coordinated by the Global Energy and Water Cycle Experiment (GEWEX) in cooperation with the Group on Earth Observations (GEO) and the Committee on Earth Observation Satellites (CEOS) and ESA (http://www.esa.int/esaLP/LPsmos.html). There is a lot of promise in composite systems comprising active and passive microwave technologies (Das et al. Citation2011). The US National Aeronautics and Space Administration (NASA) launched a combined radiometer and synthetic aperture radar operating at L-band (1.20–1.41 GHz) in January 2015 (SMAP; http://smap.jpl.nasa.gov/mission/). Although there are still limitations in terms of measuring soil moisture beyond a certain depth from the surface, RS soil moisture can still be used as a boundary condition for equations of soil moisture dynamics within a model. Satellite-derived evapotranspiration products are also becoming easily available (such as the MODerate-resolution Imaging Spectro-radiometer (MODIS) product, MOD 16 – (http://modis.gsfc.nasa.gov/data/dataprod/dataproducts.php?MOD_NUMBER=16). An algorithm based on a modified Penman-Monteith equation was used to develop MOD16 (Mu et al. Citation2007). There are known attempts to use satellite-based data to derive ET and calibrate a model against it (although not a GHM) for a 46 000-km2catchment in southern India (Immerzeel and Droogers Citation2008). The authors used the Surface Energy Balance Algorithm for Land (SEBAL) to calculate ET from satellite data. Such satellite-derived ET estimates are in themselves quite complex and parameter-intensive, and may hardly be regarded as observations. However, such attempts point to possible alternatives for calibration of GHMs in the future, when and if satellite-based ET estimates become more accurate and direct. Similarly, satellite-based leaf area index (LAI) products are now available. They include MOD15 from MODIS, ECOCLIMAP (http://www.cnrm.meteo.fr/gmme/PROJETS/ECOCLIMAP/page_ecoclimap.htm), GLOBCARBON (http://www.fao.org/gtos/tcopjs4.html) and CYCLOPES (http://postel.obs-mip.fr/?CYCLOPES-Project,82; Garrigues et al. Citation2008).

Since the launch of the GRACE satellite, research has focused on measuring terrestrial storage (Strassberg et al. Citation2009, Syed et al. Citation2009, Alkama et al. Citation2010, Tang et al. Citation2010, Döll et al. Citation2014) and to incorporate it into hydrology (Werth et al. Citation2009). GRACE, however, has a very coarse spatial and temporal resolution combined with the uncertainties in partitioning the GRACE signal. GRACE observes the combined mass changes in the atmosphere, surface waters (rivers, lakes, reservoirs, floodplains and wetlands), soil and the groundwater. Hence, the only way to measure any of the above independently requires some estimates of all the other components. Also, GRACE only provides infrequent (approx. 10 days) snapshots of changes in continental water storage, where the actual storage conditions have to be known at the time of the satellite’s pass over. Its spatial resolution is approx. 200 km (i.e. 2 × 2° at the Equator). Overall, at present, the uncertainties associated with the GRACE data are on a par with, or even higher than, the uncertainties in GHMs for small to medium size river basins, which GRACE measurement can aim to reduce (Alkama et al. Citation2010). However, it may be possible that further development of GRACE or related technologies would reduce such uncertainties. The handicap of coarser resolution is being resolved, to some extent, with the Gravity field and steady-state Ocean Circulation Explorer (GOCE) mission (2009–2013), which provides data related to Earth’s gravity at 100-km resolution (http://www.esa.int/SPECIALS/GOCE/index.html).

Improving precipitation inputs

The performance of any hydrological model depends on the quality and scale of available input data, especially precipitation. Traditional (raingauge) precipitation data remain limited worldwide. Radar technology, although an improvement over raingauge data, still has limitations due to obstructions and range. Satellites can provide uninterrupted, global meteorological data collection and hence help, in principle, to improve the primary input to hydrological models. Both the geostationary satellites, observing visible (VIS) and infrared (IR) radiation, and the low Earth Orbiting (LEO) satellites using microwave (MW) technology (along with VIS and IR), provide precipitation information. They have their distinct strengths and weaknesses (Kidd et al. Citation2009). While VIS and IR data are available at higher resolution, the precipitation derived from it is dependent on the relationship between cloud top temperature and rainfall, i.e. ‘cold-cloud duration’ methods (Kidd et al. Citation2009), but the precipitation estimates are indirect measurements and prone to errors. Passive MW technology addresses these issues, but their existing spatial and temporal resolution is still a hindrance to effective use in hydrological modelling. Precipitation data derived from passive MW techniques still need to be calibrated. Various efforts are being made to combine the strengths of the LEO and geostationary satellites to develop more reliable precipitation datasets (Kuligowski Citation2002). The use of satellite-based precipitation data in hydrological modelling is overall still in its early stages and has not yet led to significant improvements in the outputs of GHMs as compared to radar- or ground-based precipitation data (Tobin and Bennett Citation2009). Meteorological reanalysis data can help fill the gap for missing long-term historical rainfall data. Li et al. (Citation2013) compared hydrological output in southern African river basins by forcing a regional hydrological model with rainfall derived from two different sources—satellite-based Tropical Rainfall Measuring Mission (TRMM) and reanalysis of ERA-40. According to their analysis, streamflow simulations were better when the model was forced with rainfall (after bias correction) from reanalysis compared to satellite-based rainfall.

Improving spatial resolution

Wood et al. (Citation2011) argue that the uncertainties associated with existing GHMs are too high and their current usual spatial resolution at the scale of 50-km grids (considering 0.5° grids) is too coarse to simulate “water cycle science questions” effectively (Wood et al. Citation2011). The authors make a case for developing future GHMs at “hyper-resolution” scale, i.e. at the scale of 1 km globally and 100 m at the continental level. They believe that the computational constraints that prevented such studies earlier no longer exist due to easy access to parallel-clustered computing resources. Gosling et al. (Citation2010) used Campus Grid to run multiple GCMs with climate change data. They were able to reduce the computation time from 750 h to 9 h on a single-processer personal computer. Also, high resolution RS data are becoming common. Such high-resolution models would capture heterogeneity and would better represent surface–subsurface and land–atmosphere interactions. Although the idea of increasing the resolution of GHMs is feasible with increasing computer power, it is important to also consider that many local issues may be better addressed by using basin-specific models rather than GHMs. Increasing complexity and resolution of GHMs may require more spatially diverse observational data at the ground level (which is already a big problem). One question that can be raised is that of the scope of application of GHMs: with increasing spatial resolution, is it the ultimate intention to address local water resources problems by means of GHMs?

Modelling at a finer spatial resolution opens yet another possibility for running GHMs at the small basin scale (instead of grids). Some countries have a very detailed system of basin delineation. For example, South Africa evaluates and monitors its water resources on a so-called ‘quaternary’ sub-basin scale. Quaternary sub-basins are incremental drainage subdivisions, ‘flowing’ one into another and covering the entire country. There are almost 2000 of these sub-basins with an area ranging from some 60 km2 (in more humid regions of the country) to some 2000 km2 (in arid areas). The recent trend is to address water management problems at the scale of even finer subdivisions, the ‘quineries’. Countries like India, have not yet established such detailed drainage subdivisions and for many large-scale assessments only 19 major drainage areas are used. These drainage areas range from 22 000 km2 (Sabarmati River basin) to 860 000 km2 (Ganga River basin) (Amarasinghe et al. Citation2005). In the United States, watershed delineation is carried out in detail, with the entire country delineated to six levels. This is publicly available as the Watershed Boundary Dataset (http://www.ncgc.nrcs.usda.gov/products/datasets/watershed/index.html). The six levels are classified as regions (level 0), sub-regions (level 1), basins (level 2), sub-basins (level 3), watersheds (level 4) and sub-watersheds (level 5). The sixth-level drainage delineation has a drainage area of up to 160 km2. Such fine-resolution river basin delineations are still restricted to a few countries, but the recent trend is to address water management problems at the scale of even finer subdivisions. The advantage of this could be closer integration of such models with the country-specific national drainage divisions.

CONCLUSIONS

The last two decades have seen substantial effort in the development of GHMs. The models are based on either energy balance or hydrological balance, or both. The complexities of the models vary based on their objectives and applications. Structural issues, such as the number of soil layers, number of land-use classifications, crop growth models, irrigation water applications and reservoir storage, are dealt with differently in the various models. A comparative study of seven models has shown structural differences in the models and little agreement in the projections made by them. The structural differences also relate to inclusion of anthropogenic disturbances, inclusion of vegetation growth models and the type of bucket model used (leaky bucket versus non-leaky bucket).

Most of the models use satellite-based or Climatic Research Unit (CRU) weather input data, available in a 0.5-degree gridded format. This also defines the spatial resolution of the models. Most of the GHMs run on a daily temporal scale. However, for the long-term assessment of the impacts of climate change and impact on global water and food trade, a monthly temporal scale is sufficient and may simplify the calibration process.

The models also need to move to finer spatial resolution to address some of the global societal issues, but they should not aim to replace more locally focused modelling efforts. Progress in GHMs is also dependent on improved access to observed hydrometeorological data, particularly in developing countries. Access to more observed data could expand the ‘calibrated area’ of the globe, thus also improving the overall accuracy of global hydrological modelling efforts and providing more accurate, although simulated, time series of various components of the hydrological cycle. This may be seen as a possible interim (or even long-term) solution for ground-based ‘data-poor’ areas, where ground observations will probably never materialize to required levels, while RS data still require decades to ensure that time series are of a reasonable length. Hence, simulations will always be needed in such areas. Therefore, improving the accuracy of GHMs by all possible means, may be effectively seen as a form of international cooperation and economic/humanitarian assistance. However, countries have to be prepared to cooperate by allowing better access to national archives of observed hydro-meteorological data.

It is suggested that, to improve the GHMs, there should be a closer integration of satellite-based products and models that would help reduce the internal uncertainties within the models.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Acknowledgements

The authors gratefully acknowledge comments provided by Robyn Johnston (IWMI, Colombo) in reviewing the manuscript. Comments provided by two anonymous reviewers are also acknowledged.

Additional information

Funding

This study was funded by the CGIAR Research Program on Water, Land and Ecosystems (WLE).

REFERENCES

  • Agarwal, C., et al., 2002. A review and assessment of land-use change models: dynamics of space, time, and human choice. General technical report, NE-297. Newton Square, PA: US Department of Agriculture, Forest Service, Northeastern Research Station.
  • Alcamo, J., 2009. Managing the global water system. In: S. Levin, et al., eds. Princeton guide of ecology. Princeton, NJ: Princeton University Press.
  • Alcamo, J., et al., 2003. Development and testing of the WaterGAP 2 global model of water use and availability. Hydrological Sciences Journal, 48 (3), 317–337. doi:10.1623/hysj.48.3.317.45290.
  • Alcamo, J., et al., 2008. A grand challenge for freshwater research: understanding the global water system. Environmental Research Letters, 3, 1–6. doi:10.1088/1748-9326/3/1/010202.
  • Alcamo, J. and Henrichs, T., 2002. Critical regions: a model-based estimation of world water resources sensitive to global changes. Aquatic Sciences, 64 (4), 352–362. doi:10.1007/PL00012591.
  • Alkama, R., et al., 2010. Global evaluation of the ISBA-TRIP continental hydrological system. Part I: comparison to GRACE terrestrial water storage estimates and in situ river discharges. Journal of Hydrometeorology, 11 (3), 583–600. doi:10.1175/2010JHM1211.1.
  • Amarasinghe, U.A., et al., 2005. Spatial variation in water supply and demand across river basins of India. Research report 83. International Water Management Institute, Colombo, Sri Lanka.
  • Arnell, N., 1999. A simple water balance model for the simulation of streamflow over a large geographic domain. Journal of Hydrology, 217 (3–4), 314–335. doi:10.1016/S0022-1694(99)00023-2.
  • Arnell, N.W. and Gosling, S.N., 2013. The impacts of climate change on river flow regimes at the global scale. Journal of Hydrology, 486, 351–364. doi:10.1016/j.jhydrol.2013.02.010.
  • Barnard, J.D., 1948. Heat units as a measure of canning crop maturity. The Canner, 106, 28.
  • Bicknell, B.R., et al., 1997. Hydrological simulation program—Fortran. User’s manual for version 11. Athens, GA: US Environmental Protection Agency, National Exposure Research Laboratory, EPA/600/R-97/080.
  • Biemans, H., et al., 2009. Effects of precipitation uncertainty on discharge calculations for main river basins. Journal of Hydrometeorology, 10, 1011–1025. doi:10.1175/2008JHM1067.1.
  • Bierkens, M.F.P. and van Beek, L.P.H., 2009. Seasonal predictability of European discharge: NAO and hydrological response time. Journal of Hydrometeorology, 10, 953–968. doi:10.1175/2009JHM1034.1.
  • Bondeau, A., et al., 2007. Modelling the role of agriculture for the 20th century global terrestrial carbon balance. Global Change Biology, 13, 679–706. doi:10.1111/j.1365-2486.2006.01305.x.
  • Brakenridge, G.R., et al., 2012. Calibration of satellite measurements of river discharge using a global hydrology model. Journal of Hydrology, 475, 123–136. doi:10.1016/j.jhydrol.2012.09.035.
  • Das, N., Entekhabi, D., and Njoku, E., 2011. An algorithm for merging SMAP radiometer and radar data for high resolution soil moisture retrieval. IEEE-Transactions on Geoscience and Remote Sensing, 49 (5), 1504–1512. doi:10.1109/TGRS.2010.2089526.
  • de Graaf, I.E.M., et al., 2014. Dynamic attribution of global water demand to surface water and groundwater resources: effects of abstractions and return flows on river discharges. Advances in Water Resources, 64, 21–33. doi:10.1016/j.advwatres.2013.12.002.
  • Dickinson, R.E., et al., 1986. Biosphere Atmosphere Transfer Scheme (BATS) for the NCAR community climate model. NCAR Technical Note, NCAR/TN-275 + STR.
  • Dirmeyer, P.A., 2011. The terrestrial segment of soil moisture–climate coupling. Geophysical Research Letters, 38, L16702. doi:10.1029/2011GL048268.
  • Döll, P. and Siebert, S., 2002. Global modeling of irrigation water requirements. Water Resources Research, 38 (4), 8–10. doi:10.1029/2001WR000355.
  • Döll, P., et al., 2008. Advances and visions in large-scale hydrological modelling: findings from the 11th Workshop on Large-Scale Hydrological Modelling. Advances in Geosciences, 18, 51–61. doi:10.5194/adgeo-18-51-2008.
  • Döll, P., et al., 2012. Impact of water withdrawals from groundwater and surface water on continental water storage variations. Journal of Geodynamics, 59–60, 143–156. doi:10.1016/j.jog.2011.05.001.
  • Döll, P., et al., 2014. Seasonal water storage variations as impacted by water abstractions: comparing the output of a global hydrological model with GRACE and GPS observations. Surveys in Geophysics, doi:10.1007/s10712-014-9282–2.
  • Döll, P., Fiedler, K., and Zhang, J., 2009. Global-scale analysis of river flow alterations due to water withdrawals and reservoirs. Hydrology and Earth System Sciences, 13, 2413–2432. doi:10.5194/hess-13-2413-2009.
  • Döll, P., Kaspar, F., and Lehner, B., 2003. A global hydrological model for deriving water availability indicators: model tuning and validation. Journal of Hydrology, 270, 105–134. doi:10.1016/S0022-1694(02)00283-4.
  • Fekete, B.M. and Vörösmarty, C.J., 2007. The current status of global river discharge monitoring and potential new technologies complementing traditional discharge measurements [online]. In: Predictions in ungauged basins: PUB kick-off (Proceedings of the PUB kick-off meeting held in Brasilia, 20–22 November 2002), IAHS Publication 309, Brasilia, 129–136. Available from: http://iahs.info/uploads/dms/309015.pdf [Accessed 12 February 2015].
  • Fekete, B.M., et al., 2010. Millennium Ecosystem Assessment scenario drivers (1970–2050): climate and hydrological alterations. Global Biogeochemical Cycles, 24 (4). doi:10.1029/2009GB003593.
  • Fekete, B.M., Vörösmarty, C.J., and Grabs, W., 2002. High-resolution fields of global runoff combining observed river discharge and simulated water balances. Global Biogeochemical Cycles, 16 (3), 15–10. doi:10.1029/1999GB001254.
  • Fekete, M., et al., 2006. Application of isotope tracers in continental scale hydrological modeling. Journal of Hydrology, 330, 444–456. doi:10.1016/j.jhydrol.2006.04.029.
  • Foley, J.A., et al., 2005. Global consequences of land use. Science, 309, 570–574. doi:10.1126/science.1111772.
  • Freeman, M.C., Pringle, C.M., and Jackson, C.R., 2007. Hydrologic connectivity and the contribution of stream headwaters to ecological integrity at regional scales. JAWRA Journal of the American Water Resources Association, 43 (1), 5–14. doi:10.1111/j.1752-1688.2007.00002.x.
  • Fritsche, M., Döll, P., and Dietrich, R., 2012. Global-scale validation of model-based load deformation of the Earth’s crust from continental water mass and atmospheric pressure variations using GPS. Journal of Geodynamics, 59–60, 133–142. doi:10.1016/j.jog.2011.04.001.
  • Garrigues, S., et al., 2008. Validation and intercomparison of global Leaf Area Index products derived from remote sensing data. Journal of Geophysical Research, 113, G02028. doi:10.1029/2007JG000635.
  • Gerten, D., 2013. A vital link: water and vegetation in the Anthropocene. Hydrology and Earth System Sciences, 17 (10), 3841–3852. doi:10.5194/hess-17-3841-2013.
  • Gerten, D., et al., 2005. Contemporary ‘green’ water flows: simulations with a dynamic global vegetation and water balance model. Physics and Chemistry of the Earth, Parts A/B/C, 30 (6–7), 334–338. doi:10.1016/j.pce.2005.06.002.
  • Gong, L., et al., 2009. Large-scale runoff routing with an aggregated network-response function. Journal of Hydrology, 368 (1–4), 237–250. doi:10.1016/j.jhydrol.2009.02.007.
  • Gosling, S.N., et al., 2010. Global hydrology modelling and uncertainty: running multiple ensembles with a campus grid. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 368 (1926), 4005–4021. doi:10.1098/rsta.2010.0164.
  • Gudmundsson, L., et al., 2012. Comparing large-scale hydrological model simulations to observed runoff percentiles in Europe. Journal of Hydrometeorology, 13 (2), 604–620. doi:10.1175/JHM-D-11-083.1.
  • Haddeland, I., et al., 2011. Multi-model estimate of the global terrestrial water balance: setup and first results. Journal of Hydrometeorology, 12 (5), 869–884. doi:10.1175/2011JHM1324.1.
  • Hagemann, S. and Dümenil, L., 1997. A parametrization of the lateral waterflow for the global scale. Climate Dynamics, 14 (1), 17–31. doi:10.1007/s003820050205.
  • Hagemann, S. and Dümenil, L., 2001. Validation of the hydrological cycle of ECMWF and NCEP reanalyses using the MPI hydrological discharge model. Journal of Geophysical Research, 106, 1503–1510. doi:10.1029/2000JD900568.
  • Hanasaki, N., et al., 2008a. An integrated model for the assessment of global water resources—part 1: model description and input meteorological forcing. Hydrology and Earth System Sciences, 12, 1007–1025. doi:10.5194/hess-12-1007-2008.
  • Hanasaki, N., et al., 2008b. An integrated model for the assessment of global water resources—Part 2: applications and assessments. Hydrology and Earth System Sciences, 12, 1027–1037. doi:10.5194/hess-12-1027-2008.
  • Hanasaki, N., et al., 2010. An estimation of global virtual water flow and sources of water withdrawal for major crops and livestock products using a global hydrological model. Journal of Hydrology, 384 (3–4), 232–244. doi:10.1016/j.jhydrol.2009.09.028.
  • Hanasaki, N., et al., 2013. A global water scarcity assessment under Shared Socio-economic Pathways—Part 2: water availability and scarcity. Hydrology and Earth System Sciences, 17 (7), 2393–2413. doi:10.5194/hess-17-2393-2013.
  • Huard, D., 2011. A black eye for the Hydrological Sciences Journal. Discussion of ‘A comparison of local and aggregated climate model outputs with observed data’ (by G.G. Anagnostopoulos et al., 2010, Hydrological Sciences Journal 55 (7), 1094–1110). Hydrological Science Journal 56 (7), 1330–1333.
  • Hunger, M. and Döll, P., 2008. Value of river discharge data for global-scale hydrological modeling. Hydrology and Earth System Sciences, 12, 841–861. doi:10.5194/hess-12-841-2008.
  • Immerzeel, W.W. and Droogers, P., 2008. Calibration of a distributed hydrological model based on satellite evapotranspiration. Journal of Hydrology, 349, 411–424. doi:10.1016/j.jhydrol.2007.11.017.
  • Irannejad, P. and Henderson-Sellers, A., 2007. Evaluation of AMIP II global climate model simulations of the land surface water budget and its components over the GEWEX-CEOP regions. Journal of Hydrometeorology, 8 (3), 304–326. doi:10.1175/JHM579.1.
  • Islam, M.D., et al., 2006. A grid-based assessment of global water scarcity including virtual water trading. Water Resources Management, 21 (1), 19–33. doi:10.1007/s11269-006-9038-y.
  • Jin, S. and Feng, G., 2013. Large-scale variations of global groundwater from satellite gravimetry and hydrological models, 2002–2012. Global and Planetary Change, 106, 20–30. doi:10.1016/j.gloplacha.2013.02.008.
  • Kaspar, F., 2004. Development and uncertainty analysis of a global hydrological model (PhD dissertation in German). University of Kassel, Kassel, Germany.
  • Kauffeldt, A., et al., 2013. Disinformative data in large-scale hydrological modelling. Hydrology and Earth System Sciences, 17, 2845–2857. doi:10.5194/hess-17-2845-2013.
  • Kidd, C., et al., 2009. Satellite precipitation measurements for water resource monitoring. JAWRA Journal of the American Water Resources Association, 45 (3), 567–579. doi:10.1111/j.1752-1688.2009.00326.x.
  • Korzoun, V.I., et al., 1977. Atlas of world water balance. Paris: UNESCO.
  • Koster, R.D., et al., 2003. Observational evidence that soil moisture variations affect precipitation. Geophysical Research Letters, 30 (5), 1241. doi:10.1029/2002GL016571.
  • Koster, R.D., et al., 2004. Realistic initialization of land surface states: impacts on subseasonal forecast skill. Journal of Hydrometeorology, 5, 1049–1063. doi:10.1175/JHM-387.1.
  • Krysanova, V., Müller-Wohlfeil, D.I., and Becker, A., 1998. Development and test of a spatially distributed hydrological/water quality model for mesoscale water-sheds. Ecological Modelling, 106, 261–289. doi:10.1016/S0304-3800(97)00204-4.
  • Kuligowski, R.J., 2002. A self-calibrating real-time GOES rainfall algorithm for short-term rainfall estimates. Journal of Hydrometeorology, 3, 112–130. doi:10.1175/1525-7541(2002)003<0112:ASCRTG>2.0.CO;2.
  • Kundzewicz, Z.W. and Stakhiv, E.Z., 2010. Are climate models “ready for prime time” in water resources management applications, or is more research needed? Editorial. Hydrological Sciences Journal, 55 (7), 1085–1089. doi:10.1080/02626667.2010.513211.
  • Lawford, R.G., et al., 2004. Advancing global-and continental-scale hydrometeorology: contributions of GEWEX hydrometeorology panel. Bulletin of the American Meteorological Society, 85 (12), 1917–1930. doi:10.1175/BAMS-85-12-1917.
  • Li, L., et al., 2013. Comparison of the global TRMM and WFD precipitation datasets in driving a large-scale hydrological model in southern Africa. Hydrology Research, 44 (5), 770–788. doi:10.2166/nh.2012.175.
  • Liang, X., et al., 1994. A simple hydrologically based model of land surface water and energy fluxes for general circulation models. Journal of Geophysical Research, 99 (D7), 14415–428. doi:10.1029/94JD00483.
  • Lindström, G., et al., 1997. Development and test of the distributed HBV-96 hydrological model. Journal of Hydrology, 201, 272–288. doi:10.1016/S0022-1694(97)00041-3.
  • Manabe, S., 1969. Climate and the ocean circulation. The atmospheric circulation and the hydrology of the earth’s surface. Monthly Weather Review, 97, 739–774. doi:10.1175/1520-0493(1969)097<0739:CATOC>2.3.CO;2.
  • Matthews, E., 1983. Global vegetation and land use: new high-resolution data bases for climate studies. Journal of Climate and Applied Meteorology, 22, 474–487. doi:10.1175/1520-0450(1983)022<0474:GVALUN>2.0.CO;2.
  • McMillan, H., Krueger, T., and Freer, J., 2012. Benchmarking observational uncertainties for hydrology: rainfall, river discharge and water quality. Hydrological Processes, 26, 4078–4111. doi:10.1002/hyp.9384.
  • Meigh, J.R., McKenzie, A.A., and Sene, K.J., 1999. A grid-based approach to water scarcity. Estimates for eastern and southern Africa. Water Resources Management, 13, 85–115. doi:10.1023/A:1008025703712.
  • Meigh, J.R. and Tate, E.L., 2002. The GWAVA model—development of a global-scale methodology to assess the combined impact of climate and land use changes [online]. EGS XXVII General Assembly. Nice, abstract #1276. Available from: http://adsabs.harvard.edu/abs/2002EGSGA..27.1276M [Accessed 24 February 2015].
  • Melillo, J.M., et al., 1993. Global climate change and terrestrial net primary production. Nature, 363, 234–240. doi:10.1038/363234a0.
  • Milly, P.C.D. and Shmakin, A.B., 2002. Global modeling of land water and energy balances. Part I: the Land Dynamics (LaD) Model. Journal of Hydrometeorology, 3 (3), 283–299. doi:10.1175/1525-7541(2002)003<0283:GMOLWA>2.0.CO;2.
  • Milly, P.C.D., et al., 2010. Terrestrial water-storage contributions to sea-level rise and variability. In: J.A. Church, et al. eds. Understanding sea-level rise and variability. Oxford: Wiley-Blackwell.
  • Mu, Q., et al., 2007. Development of a global evapotranspiration algorithm based on MODIS and global meteorology data. Remote Sensing of Environment, 111, 519–536. doi:10.1016/j.rse.2007.04.015.
  • Müller, C. and Lucht, W., 2007. Robustness of terrestrial carbon and water cycle simulations against variations in spatial resolution. Journal of Geophysical Research, 112 (D6), 1–7. doi:10.1029/2006JD007875.
  • Müller Schmied, H., et al., 2014. Sensitivity of simulated global-scale freshwater fluxes and storages to input data, hydrological model structure, human water use and calibration. Hydrology and Earth System Sciences Discussions, 11, 1583–1649. doi:10.5194/hessd-11-1583-2014.
  • Mulligan, M., 2013. WaterWorld: a self-parameterising, physically-based model for application in data-poor but problem-rich environments globally. Hydrology Research, doi:10.2166/nh.2012.217. in press.
  • Munro, R.K., et al., 1998. Modelling land surface–atmosphere interactions over the Australian continent with an emphasis on the role of soil moisture. Environmental Modelling & Software, 13 (3–4), 333–339. doi:10.1016/S1364-8152(98)00038-3.
  • Neitsch, S.L., et al., 2002. Soil and Water Assessment Tool Theoretical Documentation, Version 2000. Grassland, Soil and Water Research Laboratory, Temple, Texas and Blackland Research Center, Temple, Texas. Texas Water Resources Institute, College Station, Texas.
  • Nijssen, B., et al., 2001a. Predicting the discharge of global rivers. Journal of Climate, 14, 3307–3323. doi:10.1175/1520-0442(2001)014<3307:PTDOGR>2.0.CO;2.
  • Nijssen, B., Schnur, R., and Lettenmaier, D.P., 2001b. Global retrospective estimation of soil moisture using the variable infiltration capacity land surface model, 1980–93. Journal of Climate, 14, 1790–1808. doi:10.1175/1520-0442(2001)014<1790:GREOSM>2.0.CO;2.
  • Noilhan, J. and Planton, S., 1989. A simple parameterization of land surface processes for meteorological models. Monthly Weather Review, 117, 536–549. doi:10.1175/1520-0493(1989)117<0536:ASPOLS>2.0.CO;2.
  • Oki, T. and Sud, Y.C., 1998. Design of Total Runoff Integrating Pathways (TRIP) – a global river channel network. Earth Interactions, 2. doi:10.1175/1087-3562(1998)002<0001:DOTRIP>2.3.CO;2.
  • Palmer, M.A., et al., 2008. Climate change and the World’s River Basins: anticipating management options. Frontiers in Ecology and the Environment, 6 (2), 81–89. doi:10.1890/060148.
  • Petrescu, A.M.R., et al., 2010. Modeling regional to global CH4 emissions of boreal and arctic wetlands. Global Biogeochemical Cycles, 24, GB4009. doi:10.1029/2009GB003610.
  • Phillips, E.E., 1950. Heat summation theory as applied to canning crops. The Canner, 27, 13–15.
  • Rabalais, N.N., et al., 2009. Global change and eutrophication of coastal waters. ICES Journal of Marine Science, 66 (7), 1528–1537. doi:10.1093/icesjms/fsp047.
  • Rodda, J.C., 1995. Guessing or assessing the World’s water resources? Water and Environment Journal, 9 (4), 360–368. doi:10.1111/j.1747-6593.1995.tb00953.x.
  • Rost, S., et al., 2008a. Agricultural green and blue water consumption and its influence on the global water system. Water Resources Research, 44 (9), 1–17. doi:10.1029/2007WR006331.
  • Rost, S., Gerten, D., and Heyder, U., 2008b. Human alterations of the terrestrial water cycle through land management. Advances in Geosciences, 18, 43–50. doi:10.5194/adgeo-18-43-2008.
  • Schuol, J. and Abbaspour, K.C., 2006. Calibration and uncertainty issues of a hydrological model (SWAT) applied to West Africa. Advances in Geosciences, 9, 137–143. doi:10.5194/adgeo-9-137-2006.
  • Sellers, P.J., et al., 1986. A simple biosphere model (SiB) for use within general circulation models. Journal of the Atmospheric Sciences, 43, 505–531. doi:10.1175/1520-0469(1986)043<0505:ASBMFU>2.0.CO;2.
  • Seneviratne, S.I., et al., 2010. Investigating soil moisture-climate interactions in a changing climate: a review. Earth-Science Reviews, 99 (3–4), 125–161. doi:10.1016/j.earscirev.2010.02.004.
  • Shmakin, A.B., Milly, P.C.D., and Dunne, K.A., 2002. Global Modeling of land water and energy balances. Part III: interannual variability. Journal of Hydrometeorology, 3 (3), 311–321. doi:10.1175/1525-7541(2002)003<0311:GMOLWA>2.0.CO;2.
  • Sperna Weiland, F.C., et al., 2011. Extracting information from an ensemble of GCMs to reliably assess future global runoff change. Journal of Hydrology, 412–413, 66–75.
  • Strassberg, G., Scanlon, B.R., and Chambers, D., 2009. Evaluation of groundwater storage monitoring with the GRACE satellite: case study of the High Plains aquifer, Central United States. Water Resources Research, 45 (5), W05410. doi:10.1029/2008WR006892.
  • Syed, T.H., Famiglietti, J.S., and Chambers, D.P., 2009. GRACE-based estimates of terrestrial freshwater discharge from basin to continental scales. Journal of Hydrometeorology, 10 (1), 22–40. doi:10.1175/2008JHM993.1.
  • Tang, Q., et al., 2009. Remote sensing: hydrology. Progress in Physical Geography, 33 (4), 490–509. doi:10.1177/0309133309346650.
  • Tang, Q., et al., 2010. Dynamics of terrestrial water storage change from satellite and surface observations and modeling. Journal of Hydrometeorology, 11 (1), 156–170. doi:10.1175/2009JHM1152.1.
  • Tobin, K.J. and Bennett, M.E., 2009. Using SWAT to model streamflow in two river basins with ground and satellite precipitation data. JAWRA Journal of the American Water Resources Association, 45 (1), 253–271. doi:10.1111/j.1752-1688.2008.00276.x.
  • Trambauer, P., et al., 2013. A review of continental scale hydrological models and their suitability for drought forecasting in (sub-Saharan) Africa. Physics and Chemistry of the Earth, Parts A/B/C, 66, 16–26. doi:10.1016/j.pce.2013.07.003.
  • van Beek, L.P.H. and Bierkens, M.F.P., 2008. The global hydrological model PCR-GLOBWB: conceptualization, parameterization and verification [online]. Report, Department of Physical Geography, Utrecht University, Utrecht, The Netherlands. Available from: http://vanbeek.geo.uu.nl/suppinfo/vanbeekbierkens2009.pdf [Accessed March 2014].
  • van Beek, L.P.H., Wada, Y., and Bierkens, M.F.P., 2011. Global monthly water stress: 1. Water balance and water availability. Water Resources Research, 47, W07517. doi:10.1029/2010WR009791.
  • Vörösmarty, C.J., et al., 1989. Continental scale models of water balance and fluvial transport: an application to South America. Global Biogeochemical Cycles, 3 (3), 241–265. doi:10.1029/GB003i003p00241.
  • Vörösmarty, C.J., et al., 2000. Global water resources: vulnerability from climate change and population growth. Science, 289, 284–288. doi:10.1126/science.289.5477.284.
  • Vörösmarty, C.J., Federer, C.A., and Schloss, A., 1998. Potential evaporation functions compared on US watersheds: possible implications for global-scale water balance and terrestrial ecosystem modeling. Journal of Hydrology, 207, 147–169. doi:10.1016/S0022-1694(98)00109-7.
  • Voß, F., et al., 2008. First results from intercomparison of surface water availability modules. Technical Report No. 1. CESR.
  • Wada, Y., et al., 2010. Global depletion of groundwater resources. Geophysical Research Letters, 37, L20402. doi:10.1029/2010GL044571.
  • Wada, Y., et al., 2011. Global monthly water stress: 2. Water demand and severity of water stress. Water Resources Research, 47, W07518. doi:10.1029/2010WR009792.
  • Wada, Y., et al., 2013. Multimodel projections and uncertainties of irrigation water demand under climate change. Geophysical Research Letters, 40, 4626–4632. doi:10.1002/grl.50686.
  • Wada, Y., Wisser, D., and Bierkens, M.F.P., 2014. Global modeling of withdrawal, allocation and consumptive use of surface water and groundwater resources. Earth System Dynamics, 5, 15–40. doi:10.5194/esd-5-15-2014.
  • Weedon, G.P., et al., 2011. Creation of the WATCH Forcing Data and its use to assess global and regional reference crop evaporation over land during the twentieth century. Journal of Hydrometeorology, 12 (5), 823–848. doi:10.1175/2011JHM1369.1.
  • Werth, S. and Güntner, A., 2010. Calibration analysis for water storage variability of the global hydrological model WGHM. Hydrology and Earth System Sciences, 14 (1), 59–78. doi:10.5194/hess-14-59-2010.
  • Werth, S., et al., 2009. Integration of GRACE mass variations into a global hydrological model. Earth and Planetary Science Letters, 277 (1–2), 166–173. doi:10.1016/j.epsl.2008.10.021.
  • Widén-Nilsson, E., 2007. Global-scale modelling of the land-surface water balance [online]. Development and Analysis of WASMOD-M. PhD. Dissertation. Available from: http://www.dissertations.se/dissertation/380756370b/ [Accessed January 2014].
  • Widén-Nilsson, E., Halldin, S., and Xu, C., 2007. Global water-balance modelling with WASMOD-M: parameter estimation and regionalisation. Journal of Hydrology, 340 (1–2), 105–118. doi:10.1016/j.jhydrol.2007.04.002.
  • Wilby, R.L., 2010. Evaluating climate model outputs for hydrological applications – opinion. Hydrological Sciences Journal, 55 (7), 1090–1093. doi:10.1080/02626667.2010.513212.
  • Wisser, D., et al., 2008. Global irrigation water demand: variability and uncertainties arising from agricultural and climate data sets. Geophysical Research Letters, 35 (24), 1–5. doi:10.1029/2008GL035296.
  • Wisser, D., et al., 2010. The significance of local water resources captured in small reservoirs for crop production – a global-scale analysis. Journal of Hydrology, 384 (3–4), 264–275. doi:10.1016/j.jhydrol.2009.07.032.
  • Wood, E.F., et al., 2011. Hyperresolution global land surface modeling: meeting a grand challenge for monitoring Earth’s terrestrial water. Water Resources Research, 47 (5), 1–10. doi:10.1029/2010WR010090.
  • Xu, C.-Y., 2002. WASMOD – The water and snow balance modeling system – chapter 17. In: V.J. Singh and D.K. Frevert, eds. Mathematical models of small watershed hydrology and applications. Highlands Ranch, CO: Water Resources Publications LLC, 555–590.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.