More stories

  • in

    The energy source is not the deciding factor

    The use of solar power in irrigation raises concerns about sustainable groundwater extraction. However, an empirical study from Bangladesh shows that the energy source is just one piece of a larger puzzle — groundwater sustainability ultimately depends on how solar irrigation systems are managed.

    Access through your institution

    Buy or subscribe

    This is a preview of subscription content, access via your institution

    Access options

    Access through your institution

    Subscribe to this journal

    Receive 12 digital issues and online access to articles

    $119.00 per year
    only $9.92 per issue

    Learn more

    Rent or buy this article
    Prices vary by article type
    from$1.95
    to$39.95

    Learn more

    Prices may be subject to local taxes which are calculated during checkout

    Additional access options:

    Log in

    Learn about institutional subscriptions

    Read our FAQs

    Contact customer support

    Fig. 1: Groundwater extraction by smallholder farmers.

    ReferencesVivanco, F. et al. J. Ind. Ecol. 26, 1543–1563 (2022).Article 

    Google Scholar 
    Balasubramanya, S. et al. Science 383, 256–258 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Alam, M. F. et al. Nat. Water https://doi.org/10.1038/s44221-025-00534-4 (2025).Fisher, J., Goldstein, J. & Ramkumar, A. Dev. Change 56, 229–253 (2025).Article 

    Google Scholar 
    Rahman, M. M. & Connor, J. D. Agric. Food Secur. 11, 35 (2022).Article 

    Google Scholar 
    Kabir, H., Ahmed, S. & Hossain, S. Int. J. Soc. Sci. Human Res. 8, 5309–5318 (2025).Article 

    Google Scholar 
    Minot, N., Hossain, S., Kabir, R., Dorosh, P. A. & Rashid, S. Assessing the Impact of Rice Price Stabilization Policies in Bangladesh: Results From a Stochastic Spatial Equilibrium Model IFPRI Discussion Paper 2252 (International Food Policy Research Institute, 2024).Islam, M. M. & Hossain, M. E. J. Politic. Econ. 31, 4 (2016).
    Google Scholar 
    Waalewijn, P. et al. Governance in Irrigation and Drainage: Concepts, Cases, and Action-Oriented Approaches—A Practitioner’s Resource (World Bank, 2020).PIGD Pulse Podcast. #25 Africa’s Water Challenge – and the Solar-Powered Solution (PIDG, 2025).Tech Trends in Energy Access: Assessing the Solar Water Pump Market (Efficiency for Access, 2023).Download referencesAuthor informationAuthors and AffiliationsDeutsche Gesellschaft für Internationale Zusammenarbeit (GIZ), Eschborn, GermanyLucie PluschkeAuthorsLucie PluschkeView author publicationsSearch author on:PubMed Google ScholarCorresponding authorCorrespondence to
    Lucie Pluschke.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Rights and permissionsReprints and permissionsAbout this articleCite this articlePluschke, L. The energy source is not the deciding factor.
    Nat Water (2025). https://doi.org/10.1038/s44221-025-00557-xDownload citationPublished: 09 December 2025Version of record: 09 December 2025DOI: https://doi.org/10.1038/s44221-025-00557-xShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Increased rainfall-runoff drives flood hazard intensification in Central Himalayan river systems

    AbstractThe Central Himalayan floodplain is one of the most flood-affected regions of the world, but we have limited knowledge of climate-change impacts on fluvial floods. Here, we provide the first large-ensemble, regional climate-change impact assessment of design floods for the Karnali River in Nepal and China, based on high-resolution modelling, considering climatic, hydrological and statistical uncertainties, needed for building local mitigation strategies. Our simulations project increases in the 1% annual exceedance probability flood magnitude of + 40% (P10/90: 33/48%) (medium-emissions) and + 79% (P10/90: 54/82%) (high-emissions) for 2060–2099, with rainfall-runoff contributing ≥ 90% of the additional flood water. We show that the uncertainty in the projections is linked to the relative differences between the magnitudes of the largest and more common events and argue that it is important to include both atmospheric only general circulation models and earth system Models in ensembles to capture the range of event and model uncertainty.

    Similar content being viewed by others

    Dynamics-constrained rainfall projection reveals substantial increase in population exposure to unprecedented floods in the North China Plain

    Article
    Open access
    19 June 2025

    Reconciling disagreement on global river flood changes in a warming climate

    Article

    28 November 2022

    Enhancement of river flooding due to global warming

    Article
    Open access
    30 November 2022

    IntroductionThe Central Himalayan floodplain is one of the most flood-affected regions of the world1 with the September 2024 floods causing damages of 1% of Nepal’s GDP with 236 deaths and 8,400 people displaced2 and by 2050, flood damages are projected to account for 2.2% of annual GPD3. Heavy rainfall and pluvial floods are projected to increase in this region, with changes occurring at faster rates than the global average4,5. Fluvial flood response to climate change is more uncertain, due to the complex interactions between environmental, climatic and hydrological conditions6. Understanding this response and its driving processes is critical for the development of flood risk management strategies targeted at the catchment-scale.The current understanding of the climate-change impact on Central Himalayan river floods is largely based on global-scale projections of Global Circulation Model (GCM) runoff projections7, hydrodynamic routing of GCM runoff projections8,9, or hydrological modelling using GCM climate projections10. While global studies are valuable for understanding regional trends, their validity on the catchment-scale is limited for several reasons: (i) the global application leads to a loss of performance at the catchment-scale (e.g. the observed direction of change is not captured by GCM runoff predictions for 40% of the catchments7; (ii) the projections are averaged over a coarse modelling grid, which restricts change impact assessments to large catchments7; (iii) the small-scale variability of the hydro-climatic system of mountainous catchments is not captured by the coarse modelling grid of GCMs5,11. Regional-scale assessments provide an improved representation of this variability due to finer modelling grids, but the studies investigating the Central Himalayas have used older generations of climate projections12,13, and/or do not capture appropriate design floods used for flood risk management purposes (e.g. 1% Annual Exceedance Probability (AEP)). Furthermore, the studies that use Flood Frequency Analysis (FFA) to determine design flood magnitudes do not estimate the uncertainties introduced by the FFA, although this is also needed to correctly design flood mitigation strategies14.Here, we present the first, comprehensive regional climate change impact assessment of design flood hazards in the Central Himalayan Karnali River in Nepal and China (Fig. 1) derived from high-resolution, large ensemble modelling using the latest generation of climate projections, and capturing climatic, hydrological and statistical uncertainties. The applied framework couples the climate projections with a continuous, fully distributed (500 × 500 m) hydrological model and a FFA. We apply: probabilistic climate projections (12 bias-corrected and downscaled CMIP6 models15 for medium (SSP245) and high emission (SSP585) scenarios for the near-future (2020–2059) and far future (2060–2099) to estimate the climatic uncertainty; a Generalised Likelihood Uncertainty Estimation framework (GLUE)16 with 64 parameter sets to capture the hydrological uncertainty, and utilise the GLUE predictions to capture the measurement uncertainty; and a bootstrapping framework17 to capture the sampling uncertainty in the FFA. We apply the ‘Spatial Processes in Hydrology’ (SPHY) hydrological model18 which represents cryospheric and hydrological processes, including the simulation of glacier retreat19, which improves predictions for long-term climate change assessments. However, the current land cover is maintained throughout the modelling period as no high-resolution projections of potential future land cover exist for the Karnali catchment, and estimating the potential future land cover change would increase the uncertainty in the model predictions. Thus, the flood hazard projections do not account for potential land use changes.Fig. 1Map of the Karnali River catchment in Nepal and China. The topography is derived from the SRTM 90 m V4.167. The station network shows the location of the hydrological and meteorological stations used in this study. The numbers on the station network relate to the gauge’s identification. These stations are maintained by the Department of Hydrology and Meteorology (DHM), Nepal. The map was created with QGIS Version 3.34.3 (https://qgis.org/) and uses the WGS84 UTM Zone 44° N coordinate system.Full size imageOur simulations find, with high confidence, that flood magnitudes for the 1% AEP will increase by 23% and 26% in the near-future (2020–2059), and increase further in the far-future (2060–2099) by 40% and 79% for the medium and high emissions scenarios respectively. We also show that rainfall-runoff is the largest source of flood water (> 90%), but that snowmelt is an important source for individual medium-discharge flood events. We find that the FFA uncertainty is strongly affected by the timing and occurrence of rare precipitation events, which dominate flood magnitude response. We therefore argue that a large ensemble of climate models is needed to capture a large enough sample size of rare events to enable the analysis of flood frequencies under future climates. Both Atmosphere Only General Circulation Models (AOGCMs) and Earth System Models (ESMs) should be included in the ensemble to capture climate model uncertainty, as AOGCMs tend to project a wider range of changes with respect to the baseline, with lower minimum bounds, whilst ESMs project a narrower range of and higher median changes. These results build on existing studies to provide an advance in understanding of fluvial flood uncertainty and drivers, and propose a modelling framework for predicting AEP flood magnitudes under climate change, necessary for flood risk mitigation.Projected changes in flood magnitudesFlood discharge projections obtained from hydrological modelling (Fig. 2a) are used in the FFA to estimate flood magnitudes for probabilistic climate projections. Our results show that in the near-future (2020–2059), the P50 flood magnitudes of events with a 1% AEP are projected to increase from the baseline (1975–2014) scenario by + 23% (P10 + 21%, P90 + 17%) for the medium-emission scenario, SSP245, and + 26% (P10 + 25%, P90 + 27%) for the high-emission scenario, SSP585 (Fig. 2b), with 10 (SSP245) and 11 (SSP585) of the 12 climate models projecting increases. However, the ensemble-median difference between both scenarios is 3%, with higher increases for SSP245 than for SSP585 for 6 of the 12 climate models.Fig. 2The flood hazard simulations of the flood discharge classified as AMAX flows (A), the flood magnitudes predicted by the FFA (B), and the 30-year mean flood discharge (C), for the scenario and two projected climate scenarios predicted for probabilistic climate projections of 12 CMIP6 models. The solid lines indicate the median, and the dotted lines indicate the 10 and 90th percentiles of the ensemble.Full size imageIn the far-future (2060–2099), the P50 1% AEP flood magnitude increases from the baseline by 40% (P10 + 33%, P90 + 48%) for SSP245 and 79% (P10 + 54%, P90 + 82%) for SSP585 (Fig. 2b) with 12 (SSP245) and 11 (SSP585) of the 12 climate models projecting increases. Similarly, the frequency of a given flood magnitude increases, with the P50 1% AEP event projected to occur once every 11 years for SSP245 and every 3 years for SSP585, with 12 (SSP245) and 11 (SSP585) of the 12 climate models projecting increases. An intensification of magnitude occurs with higher emissions, with 11 of the 12 climate models projecting higher increases for SSP585 than for SSP245.The timing when both emission scenarios diverge varies between the percentiles of the ensemble. The P50 30-year-mean flood discharge increases at similar rates for SSP245 (+ 20%) and SSP585 (+ 21%) between the years 2020 and 2059 (Fig. 2c). In the year 2060, the P50 30-year-mean flood discharge of both scenarios starts diverging, and it further increases by 7% (SS245) and 35% (SSP585) between 2060 and 2099. The timing when higher flood discharges start to be predicted for SSP585 varies with the percentile, the P90 starts diverging between 2035 and 2040, the P50 around the year 2060, and the P10 between 2065 and 2070.We do not identify systematic patterns in the projected floods for related climate models20, despite models in the same family usually having similar climate properties; models of the same family can project both small and large changes in 1% AEP flood magnitudes (Fig. 3). Some models (e.g. HadAM, CCM) project smaller spreads, whilst others project larger spreads (e.g. ECMWF). Nevertheless, we find that ESMs generally project larger increases and provide a smaller range of change than AOGCMs, possibly related to their inclusion of the carbon cycle, vegetation, atmospheric chemistry and other processes, although other comparisons accounting for differences between AOGCMs and Earth System Models (ESMs) are currently lacking20, and further testing is necessary to confirm this result. Fig. 3The projected changes of the median 1%-AEP flood magnitudes relative to the Baseline for the 12 CMIP6 members grouped by their climate model family (A); and grouped by the type of climate model into 6 atmospheric only general circulation models (AOGCMs) and 6 earth system models (ESMs) (B). The classification of model family and model type is based on Kuma et al.20.Full size imageThe sources of floodwaterWe analyse the floodwater sources simulated by the hydrological model, divided between direct rainfall-runoff, glacier melt, groundwater-driven baseflow and snowmelt, to understand the drivers of the projected flood intensification. In the baseline period, rainfall-runoff is the most important source, with a mean floodwater contribution of 78 ± 11% (± standard deviation), followed by glacier melt (10 ± 5%), baseflow (8 ± 5%) and snowmelt (3 ± 8%) (Fig. 4). The importance of rainfall-runoff increases with the flood discharge, with rainfall-runoff contributing ≥ 90% to events with discharges ≥ 17,500 m3/s (high flows) (Fig. 5a). Consequently, the importance of the other sources decreases with increasing discharge. However, other sources are important for events with lower (< 10,000 m3/s) and medium (10,000–17,500 m3/s) discharges with contributions ≥ 25%, and 10–25% respectively (Fig. 5a). High rainfall-runoff contributions are predicted without clear seasonal patterns, but event timing is important: higher snowmelt contributions are predicted in the pre- and early-monsoon seasons (typically up to the 27th week of the calendar year); glacier meltwater contributions increase towards the mid-monsoon season (weeks 27–35); and higher baseflow contributions are predicted in the late-monsoon and post-monsoon seasons (week 35 onwards). These results provide a new contribution to the literature.Fig. 4The mean flood discharge composition. The error bars indicate the standard deviation, which is caused by differences in the event characteristics, climate projections, and the hydrological model parameterisation.Full size imageFig. 5Composition of each simulated flood event (n = 12 Climate models x 64 hydrological parameter sets x 40 years). Column A shows the relationship between the runoff contribution and the magnitude of the flood discharge. Column B shows the relationship between the runoff contribution and the timing of the flood event.Full size imageThe drivers of changeIn the near-future and far-future, the projected flood intensification is driven by increasing rainfall-runoff for both scenarios. The mean flood discharge increases by 1,380–4,581 m3/s compared to the baseline, whereas the mean rainfall-runoff increases between 1,202 and 4,255 m3/s, and thus accounts for 91–93% of the projected increase of the mean flood discharge (Fig. 4). Furthermore, increasing rainfall-runoff also causes the intensification of high flood discharges, which remain composed of ≥ 90% rainfall-runoff (Fig. 5a). The baseflow and glacier melt contributions to the mean flood discharge increase by 77–215 m3/s and 82–215 m3/s, respectively (Fig. 4). Their relative importance decreases due to the higher rainfall-runoff increase, and we project no changes in their seasonality. The flood discharge intensification is dominated by increasing rainfall-runoff in both emissions scenarios, and thus, differences in future emissions affect the rate of change rather than their driving processes.Temperature-related changes in melt processes lead to a decrease in snowmelt runoff for most flood events. The mean snowmelt contribution decreases from 3 ± 8% in the baseline to 2 ± 7% in the near-future (SSP245 and SSP585) and 1 ± 6% (SSP245) and 0 ± 2% (SSP585) in the far-future, which is caused by an earlier onset of the melting season (Fig. 5b). However, the occurrence probability of medium-discharge flood events with 10–40% snowmelt contributions is projected to increase from 0.30% in the baseline to 0.65% in the near-future, after which it decreases to 0.17% for SSP585. For SSP245, an increase to 0.45% is delayed to the far-future. The changes in the frequency of these events indicate that rising temperatures cause an initial increase in melt contributions for individual pre-monsoon and early monsoon season flood events. The warming also shifts the melting season earlier, which leads to lower flood discharge contributions as the temperature continues to increase, as indicated by the lower frequency in the far-future of SSP585 (Fig. 5a). While the occurrence probability is below 1% at the mountain outlet, the projected changes may indicate a shift in the flood seasonality, frequency and magnitude in smaller upstream subbasins.Performance of the hydrological modelThe hydrological ensemble is a behavioural representation of the hydrological system of the Karnali River catchment, indicated by the hydrograph inspection, the model efficiencies and the streamflow composition. The ensemble replicates the flow seasonality with high flows during the monsoon and low flows during the winter and times the transitions in the Pre-monsoon and Post-monsoon seasons well (Fig. 6). The seasonality of the runoff composition (Supplementary Figure S4) with snowmelt contributions in the Pre- and early Monsoon season (Mar – Jun), rainfall-runoff and glacier melt during the Monsoon season (Jun – Sep), and baseflow in the Post-monsoon season (Oct – Nov) reflects our understanding of Central Himalayan catchment hydrology11,21,22,23. The good model performance is, furthermore, illustrated by the high ensemble median efficiencies of: (i) the Nash-Sutcliffe-Efficiency (NSE) of 0.85 (range: 0.75–0.87) for the calibration period and 0.82 (range: 0.70–0.85) for the validation period, and (ii) the coefficient of determination (R2) of 0.85 (range: 0.77–0.87) and 0.84 (0.80–0.86) for the calibration and validation periods, respectively. These efficiencies are at the higher level of reported efficiencies in Central Himalayan catchments, ranging between 0.56 and 0.88 (NSE) and 0.63–0.89 (R2)23,24,25,26,27.Fig. 6Hydrographs of the Karnali River for the calibration and validation periods. The discharge observations are obtained by the Department of Hydrology and Meteorology Nepal (DHM). The solid lines represent the ensemble-median (simulations) and the observations. The intervals represent the 2.5th to 97.5th interval of the ensemble predictions, and the discharge uncertainty of the observations. The median Nash-Sutcliffe efficiencies are 0.85 and 0.82 for the calibration and validation periods, respectively.Full size imageThe evapotranspiration module was calibrated and validated against satellite-based annual actual evapotranspiration (ETa) estimates (Fig. 7a). The ensemble-median PBIAS is -11% (-20% to + 10%) in the calibration period and − 13% (-22% to + 7%) in the validation period, indicating a moderate underestimation of the catchment-mean ETa rates. However, the performance of each ensemble member is satisfactory or better28, and the correlation between observed and simulated land-cover-specific ETa rates is very high (R2 > 0.95), indicating a very good representation of the spatial ETa variability (Fig. 7b). However, the model’s ability to reproduce ETa rates varies between the years, and the percentage difference is within − 3% to -15%, but reaches up to -22% to -32% in the years 2005, 2014 and 2015. However, given the large uncertainty in the climate input datasets, these biases are likely caused by precipitation input deficits, as precipitation products are known to underestimate the high-mountain precipitation29,30. If Eta observations are not considered in the model calibration, precipitation input deficits may be compensated by an ETa underestimation31,32. Thus, we argue that it is important to consider ETa observations during the calibration of hydrological models in Himalayan catchments.Fig. 7Comparison of the annual actual evapotranspiration (ETa)simulations with MODIS satellite estimates. (A) shows the catchment-mean ETa rates, whereas the bars show the ensemble-median rates and the error bars the standard deviation. (B) compares the land-cover-specific mean ETa of the ensemble-median and MODIS.Full size imageThe uncertainty in the modelling ensembleWe analyse the uncertainty composition of the ensemble predictions using the 3-way analysis of variance (ANOVA) framework33. However, instead of decomposing the uncertainty contribution of climate models, climate scenarios and the internal variability, we assess the contribution of the ensembles of climate models, hydrological parametersets, and flood frequency curves. It is worth noting that the FFA ensemble estimates the uncertainty associated with the internal variability from a bootstrapping approach17.The hydrological ensemble is the largest uncertainty source in the baseline, contributing 52% to the overall variance of the 1% AEP event (Fig. 8). For the projected scenarios, the hydrological uncertainty varies between the climate scenarios, contributing 53% (SSP245) and 43% (SSP585) to the 1% AEP in the near-future, and 37% (SSP245) and 35% (SSP585) in the far-future. However, the mean standard deviation (σ) of the hydrological ensemble averaged over the climate ensemble remains with 20–22% stable for all climate scenarios and return periods (Supplementary Figure S 5). This indicates that the hydrological parametersets introduce a large, but relatively constant uncertainty, and that the decreasing relative variance contribution reflects an increase in the overall variance.Fig. 8The uncertainty composition of the modelling ensemble estimated from 3-Way ANOVA. The main uncertainty components are the hydrological parametersets (HY), the climate model ensemble (CM) and the flood frequency analysis ensemble (FFA). The p-values of all main components and HY: CM interaction effects are below 0.05. The p-values of the FFA component interactions are > 0.1 with exception of the CM: FFA interactions for SSP245 (2060–2099) (< 0.1) (Supplementary Table S5).Full size imageThe variance introduced by the climate models increases with time and emissions, but this trend is superimposed by an increasing FFA variance for higher frequencies. The climate model ensemble contributes 30% to the overall variance of the 1% AEP in the baseline. For SSP245, this fraction decreases to 26% in the near-future, after which it increases to 36% in the far-future. For SSP585, the contribution is with 41% (near-future) and 49% (far-future) notably higher than for SSP245, which indicates that the flood response becomes increasingly uncertain with time and emissions. The initial decrease in the climate ensemble variance for the near-future of SSP245 is caused by superimposing effects of the FFA variance, which is indicated by consistent climate variance increases with time and emissions for lower frequencies 10% AEP to 2% AEP).The uncertainty associated with the FFA ensemble is the lowest of the three components, but it interacts with the climate model uncertainty. Generally, the FFA uncertainty increases with the frequency, because the FFA is is sensitive to the slope of the flood frequency curve above the 5% AEP, which is fitted to a low sample size and extrapolated beyond the 40-year record length to the 1% AEP34,35. In the baseline, the FFA variance increases with the event frequency from 8% (10% AEP) to 11% (1% AEP). For the projections, the FFA variance patterns vary between the scenarios, whereas the 1% AEP variance decreases to 8–9% for SSP585, and increases to 12% (near-future) and 15% (far-future) for SSP245. Generally, the variance of component interactions is less than 3%, but in the case of SSP245 (far-future), the interaction effect of climate and FFA ensembles is 7% (p-value < 0.1).We attribute this CM-FFA interaction effect, and the higher FFA variance to the flood discharge time series projected for individual climate members. The Flood Percentile (FPn) aggregates the flood events by the nth frequency percentile (see Supplementary Information 1). In the case of the far-future of SSP245, 5 out of 12 models predict > 40% higher FP100 discharge (highest discharge in the record) relative to the FP92.5 discharge (4th highest discharge) (Fig. 9). This large difference between the rarest and more frequent events increases the sampling uncertainty because the rarest events are removed or duplicated in the bootstrapping process, which increases the difference between the slopes of the individual flood frequency curves. The case of the far-future of SSP245 indicates that the FFA uncertainty is affected by projections of rare extreme precipitation events and is thus sensitive towards the selection of climate models and the time frame in which the most extreme events may or may not occur.Fig. 9Comparison of the standard deviation of the 1% AEP event and the FP100 discharge (A). (B) compares the standard deviation of the 1% AEP event with the relative difference between the FP100 and FP92.5 discharges. The red points are the samples of the SSP245 scenario for the far-future. The standard deviation of the 1% AEP is affected by the relative difference between the FP100 and FP92.5 but there is no significant relationship between standard deviation of the 1% AEP and the FP100 discharge.Full size imageDiscussionCurrent climate projections indicate with high confidence that heavy precipitation and pluvial floods increase with increasing Greenhouse Gas (GHG) emissions in the Central Himalayas, but are less certain how these increases affect fluvial floods4,6. Previous studies suggest an increase in fluvial flood discharges, and while direct comparisons are hampered by differences in the research design (e.g. spatial and temporal extent, datasets, models, catchments, event frequencies), our projections fall within the reported range of + 20 to + 108%9,10,12,13. We can attribute these changes to increasing rainfall-runoff, consistently predicted for a large ensemble of climate projections and hydrological model configurations, suggesting that projected increases in heavy precipitation will cause an intensification of river floods in the foreland of Central Himalayan catchments.The uncertainty in predicting the hydrological response to climate change can exceed the uncertainty in predicting climate response to GHG emissions in mountainous catchments in which melt processes are important36. In the Himalayas, the uncertainty in hydrological modelling applications is increased by the coarse resolution of physical and climatic datasets and a lack of hydro-meteorological observations at higher elevations (> 4,000 masl)5,23,29,37. Here, the hydrological modelling is the largest source of uncertainty in the baseline. However, the ‘good’ to ‘very good’ modelling efficiencies28, the agreement in the seasonality of the runoff composition with other studies11,21,22, and the similarities in the projected changes in the hydrological behaviour, such as an earlier onset of the snow-melt season11,38,39, indicate that the hydrological ensemble reflects the catchment hydrology well. The relative uncertainty in the hydrological predictions decrease for future climates, which may be attributed to the dominance of rainfall-runoff in floodwater generation, as the rainfall-to-runoff conversion is less complex than other processes that include freezing, melting, and percolation.This study does not consider projected land use changes in the experimental design, although the land use is an important component of the hydrological cycle and affects flood hazard characteristics40. In the case of Nepal, two drivers of land use change may be of particular relevance: (i) climatic changes shift climate zones and, thus, the natural vegetation cover41,42, and; (ii) the expansion or abandonment of agricultural terraces (e.g. caused by migration), which affect the surface and subsurface runoff generation and the hydrological connectivity and, thus, alter the contribution to the flood discharge43,44.The Himalayan foreland in Nepal is, under both past and current climatic conditions, prone to flooding, with high mortalities and impacted populations, economic damages, and other repercussions, including increasing food insecurity and the outbreak of epidemics1,45,46,47. Future GHG emissions are likely to exacerbate these impacts. For the near-future, the relationship between flood magnitude and emissions is less pronounced due to the small differences between the SSP245 and SSP585 flood magnitudes from the relatively small differences in GHG emissions between the two climate scenarios (also possibly due to internal variability (i.e. natural climatic fluctuations) obscuring responses48,49. For the far-future however, our simulations suggest that GHG emissions cause a sustained flood hazard intensification – even for the SSP245 scenario, where emissions are higher for each year of the near-future than for the far-future, while the 30-year mean flood discharge continuously increases until the year 2099. The flood projections imply that: (i) Central Himalayan floods will intensify with time and emissions; (ii) this intensification is likely to continue for decades after the peak emissions and; (iii) flood magnitudes are likely to remain above current levels until the end of the century.Flood magnitudes in the far-future are characterised by an increased spread of flood magnitude predictions between the individual CMIP6 members, ranging from + 9% to + 69% (SSP245) and − 12% to + 144% (SSP585) for the median 1%-AEP projections from the hydrological model and FFA ensemble. Furthermore, the FFA projections are sensitive to the slope of the frequency curves, and thus to the rarest, most extreme events (i.e., the FP100) and the discharge difference between the rarest and more common events (i.e., the FP92.5 and FP100). Both the climate model differences and the FFA sensitivity to rare events make the projection of design floods (i.e. the 1%-AEP event) susceptible to the unknown true frequency of the rarest projected events. Therefore, it is important to maintain a large ensemble of climate models to guarantee a larger sample size of the rarest events (i.e. the FP100 events of the climate models). If necessary, a new selection criterion should be developed for sub-setting climate models, which account for the frequency of extremes rather than the commonly used temperature and annual precipitation subset (e.g. dry-cold, dry-warm, wet-cold, wet-warm). Furthermore, a subset should include a combination of AOGCMs and ESMs, as these tend to project different rates of change.The near-future flood projections indicate that the Karnali River is already in a phase of flood hazard intensification, so existing flood risk management strategies need to be adapted to maintain current flood protection levels. Flood defences, such as embankments and outlet structures, are typically designed and implemented on decadal time scales and, therefore, new projects should take the projected future changes in flood hazards into consideration50,51. The inherent uncertainties of future emissions and their impact on the climate and flood response create conflicts between designing protection levels for the worst-case scenario versus the economic feasibility of such structures. New, flexible, and anticipatory flood risk management strategies are therefore needed that can account for emerging new knowledge and understanding. Such strategies could include flexible designs for future flood risk infrastructure, which include the potential for later adaptations and upgrades, such as increases in storage space for floodwater, or increased dam walls levels. Additional strategies could include allocating space now for flood water retention areas within urban-zoning development plans, and policies and institutional workflows that reduce the period between planning and operation.ConclusionsThis study presents an environmental modelling framework that uses high-resolution, large ensemble modelling to predict future design floods at the catchment scale for the Central Himalayan Region. We provide design floods (i.e. 1%-AEP) for future climate change scenarios, and the hydrological processes driving these potential changes, thereby providing practical information for climate change adaptation for flood risk management. The framework extends current uncertainty analyses by incorporating sampling uncertainty into the FFA. The uncertainty analysis highlights the sensitivity of the FFA to the flood discharge projections of individual climate models, emphasising the necessity to maintain large climate ensembles for flood impact analyses. These ensembles should contain both AOGCMs and ESMs, as these families differ in their flood projections.The application of this framework projects the intensification of flood hazards over time and with increasing emissions. The hydrological and climate ensemble consistently predicts that the flood hazard intensification is driven by rainfall-runoff increases, which increases the evidence that increases in heavy precipitation will lead to increased river floods in the foreland of Central Himalayan catchments.Online methodsResearch designWe combine hydrological and statistical modelling with probabilistic climate projections to predict the potential changes in future flood hazards in the Karnali River in Nepal and China. In the first stage, we implement the hydrological model Spatial Processes in Hydrology (SPHY) for historical conditions (past and observed hydro-meteorological conditions) and identify an ensemble of behavioural parameter sets. This ensemble is then forced with probabilistic climate projections for three scenarios: the Baseline (1975–2014), the medium-emission scenario SSP245 (2020–2099) and the high-emission scenario SSP585 (2020–2099). In the last stage, a Flood Frequency Analysis (FFA) is conducted with the simulated flood discharges, which are classified by the Annual Maximum flow (AMAX). We compare the differences between the Baseline and the projected scenarios (SSP245 and SSP585) to quantify the changes in flood hazard magnitude and frequency between the past (1975–2014) and the near-future (2020–2059), as well as the far-future (2060–2099). We chose 40-year time frames over the commonly used 30-year time frames to increase the length of the flood record, which is the main source of uncertainty in the FFA34.Data processingThe utilised datasets comprise gridded and point-based datasets that determine the environmental (static) and climatic (dynamic) boundary conditions, plus calibration and validation datasets for the hydrological model (Supplementary Table 1). All gridded datasets are resampled to the modelling resolution (500 × 500 m) and reprojected to the UTM44N coordinate system (EPSG: 32644). The preprocessing of hydro-climatic data includes:

    Precipitation (historical): The network of precipitation gauges is concentrated in the floodplains and valleys, and precipitation at high elevations is poorly observed29. We therefore use monthly gridded satellite-based precipitation estimates (GPM IMERG Final Precipitation L3 1 Month V00652) and disaggregate them to the daily resolution using precipitation measurements of the Department of Hydrology and Meteorology, Nepal (DHM), following the approach of Arias-Hidalgo et al. (2013)53.

    Temperature (historical): We use the WATCH forcing Data methodology applied to ERA-Interim dataset (WFDEI)54,55 to determine the historical temperature boundary conditions. This dataset has a spatial resolution of 0.5 × 0.5° and is downscaled to the modelling resolution using a lapse rate derived from DHM temperature observations.

    Climate projections: We use the temperature and precipitation predictions of an ensemble of 12 CMIP6 Global Circulation Models (GCM), which have been bias-corrected using Empirical Quantile Mapping and downscaled to 0.25 × 0.25º by Mishra et al. (2020)15. The CanESM5 member was removed from the ensemble because it does not depict the monsoon seasonality in the catchment (Supplementary Fig. 3). The temperature data is downscaled to the modelling resolution using a lapse rate, and the precipitation data is downscaled using bilinear interpolation.

    Discharge: We use daily discharge data and stage-discharge observations (DHM) at the mountain outlet for the model calibration. The uncertainty in the discharge is estimated using the BaRatin approach56.

    The hydrological modelThe catchment hydrology is simulated by the Spatial Processes in Hydrology (SPHY) model version 3.018. SPHY is a distributed (raster-based) model that simulates soil processes, snow processes and evapotranspiration on the grid scale, and glacial processes on a sub-grid scale. The model distinguishes four types of runoff; rainfall-runoff QRR, baseflow QBF, snowmelt QSM, and glacier melt QGM which compose the total runoff QTOT of each cell:$$:{Q}_{TOT}=:{Q}_{RR}+:{Q}_{BF}+{Q}_{SM}+{Q}_{GM}:$$
    (1)
    The total runoff is then routed to the downstream cell based on a flow direction raster. The model implements a flow recession coefficient to account for channel friction (Supplementary Table 2).The soil processes are calculated as in the SWAT and SWAP models57. The soil is divided into a rootzone layer SW1, a subzone layer SW2 and the groundwater layer SW3. The upper two layers SW1 and SW2 contribute to rainfall-runoff whereas the model simulates surface runoff (saturation excess runoff and infiltration excess runoff), and slower lateral flow from the soil layers. Percolation to the groundwater storage and baseflow are calculated using a recession coefficient to relate the baseflow to the groundwater recharge.The reference evapotranspiration is estimated from the modified Hargreaves evapotranspiration equation58 and crop coefficients are used to estimate the potential evapotranspiration rates for different land covers59. Snow processes are simulated by a dynamic snow storage model60 using a degree-day-factor approach to quantify the snow melt61. Glacier melt is also calculated using the degree-day factor approach whereas separate factors are defined for clean-ice and debris-covered glaciers. The evolution of glaciers (the retreat or advance) is simulated from the mass balance (snow accumulation – glacier melt) using a mass conserving ice redistribution approach which is applied once per year19.Hydrological modelling setupWe use the SPHY model version 3.0 with the dynamic glaciers, infiltration, groundwater and snow modules. The model is applied at 500 × 500 m grid resolution with daily time steps. It is calibrated and validated for 2002–2006 and 2007–2015, respectively.A stepwise calibration approach is used to identify behavioural parameter sets. In the first stage, a Regional Sensitivity Analysis with 1,500 Latin Hypercube Samples (LHS)62 of 21 parameters is conducted to identify sensitive parameters. In the second stage, the 15 sensitive parameters are calibrated within a GLUE framework using 10,000 parameter sets sampled with LHS. We implemented two scaling factors as calibration parameters: (i) a precipitation correction factor to account for the bias in the precipitation data (similar to Lutz et al. (2014)11, and (ii) a multiplication factor for the crop coefficients which control the evapotranspiration for different land cover classes (similar to Khanal et al. (2021)19. The precipitation correction factor is used for the model calibration but is not transferred to the climate projections because the CMIP6 data is bias-corrected.We focus on the model performance during high flows for selecting the behavioural parameter sets using a multi-criteria approach to increase the robustness of the ensemble predictions63. However, we exclude parameter sets with percentage BIAS (PBIAS) exceeding ± 30% for the annual actual evapotranspiration from MODIS, the 8-day snow extent from MODIS, and the discharge during the winter months to prevent the selection of parameter sets that do not depict critical aspects of the catchment hydrological behaviour. Of the remaining parameter sets, we select the 25 members with the highest Nash-Sutcliffe Efficiency (NSE), PBIAS of high flows (≥ 5,000 m3/s), and a modified version of the extended GLUE (eGLUE)64 to account for the uncertainty in the discharge observations. The eGLUE quantifies the number of days during which the simulated discharge is within the uncertainty interval of the observed discharge. The modified eGLUE is the sum of the simulated discharge for all days that fall within the observed confidence interval divided by the total number of days. This adjustment is made to prevent lower flows, which occur throughout most of the year, from dominating the classification. After the removal of duplicates, 64 parameter sets are maintained.Flood frequency analysisWe apply a Flood Frequency Analysis (FFA) to quantify the changes in flood magnitudes for the projected climate scenarios. The Flood Frequency Analysis (FFA) is a statistical method that utilises an Extreme Value Distribution (EVD) to estimate flood magnitude based on flood frequency (occurrence probability) from a record of flood events65. The FFA is fitted to each AMAX record of the modelling ensemble (12 CMIP6 members X 64 hydrological models) using the L-Moments approach for parameter estimation66. We evaluated the performance of eight EVD because it is generally impossible to determine the most suitable distribution before the application65, and selected the Wakeby distribution for the Flood frequency analysis (Supplementary Table 4). We employ a bootstrapping approach (n = 1,000) based on Burn (2003)17 to estimate the sampling uncertainty, which relates to the difference between the distribution used in the FFA and the true distribution of the flood discharge. The flood discharges are sorted before the bootstrapping, and the same bootstrap samples are taken for all combination of climate models and hydrological parametersets to ensure the comparability of FFA samples in the uncertainty decomposition.

    Data availability

    The code and datasets to reproduce the figures are available on Zenodo at https://doi.org/10.5281/zenodo.16408489.
    ReferencesEM-DAT/CRED EM-DAT: The emergency events database. (2021).The World Bank Group & Asian Development Bank. Climate Risk Country Profile: Nepal https://climateknowledgeportal.worldbank.org/sites/default/files/2021-05/15720-WB_Nepal%20Country%20Profile-WEB.pdf (2021).Ahmed, M. Assessing the Costs of Climate Change and Adaptation in South Asia (Asian Development Bank, 2014).Gulev, S. K. et al. Changing State of the Climate System. Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change 287–422 https://doi.org/10.1017/9781009157896.004 (2023).Scott, C. A. et al. Water in the Hindu Kush himalaya. in The Hindu Kush Himalaya Assessment: Mountains, Climate Change, Sustainability and People (eds (eds Wester, P., Mishra, A., Mukherji, A. & Shrestha, A. B.) 257–299 (Springer International Publishing, Cham, doi:https://doi.org/10.1007/978-3-319-92288-1_8. (2019).Chapter 

    Google Scholar 
    Seneviratne, S. I. et al. Weather and climate extreme events in a changing climate. Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change 1513–1766 https://doi.org/10.1017/9781009157896.013 (2023).Zhang, S. et al. Reconciling disagreement on global river flood changes in a warming climate. Nat. Clim. Change. 12, 1160–1167 (2022).Article 
    ADS 

    Google Scholar 
    Hirabayashi, Y. et al. Global flood risk under climate change. Nat. Clim. Change. 3, 816–821 (2013).Article 
    ADS 

    Google Scholar 
    Hirabayashi, Y., Tanoue, M., Sasaki, O., Zhou, X. & Yamazaki, D. Global exposure to flooding from the new CMIP6 climate model projections. Sci. Rep. 11, 3740 (2021).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Dankers, R. et al. First look at changes in flood hazard in the Inter-Sectoral impact model intercomparison project ensemble. Proc. Natl. Acad. Sci. 111, 3257–3261 (2014).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Lutz, A. F., Immerzeel, W. W., Shrestha, A. B. & Bierkens, M. F. P. Consistent increase in high asia’s runoff due to increasing glacier melt and precipitation. Nat. Clim. Change. 4, 587–592 (2014).Article 
    ADS 

    Google Scholar 
    Pechlivanidis, I. G. et al. Analysis of hydrological extremes at different hydro-climatic regimes under present and future conditions. Clim. Change. 141, 467–481 (2017).Article 
    ADS 

    Google Scholar 
    Wijngaard, R. R. et al. Future changes in hydro-climatic extremes in the upper Indus, Ganges, and Brahmaputra river basins. PLOS One. 12, e0190224 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Pappenberger, F. & Beven, K. J. Ignorance is bliss: or seven reasons not to use uncertainty analysis. Water Resour. Res. 42, W05302 (2006).Article 
    ADS 

    Google Scholar 
    Mishra, V., Bhatia, U. & Tiwari, A. D. Bias-corrected climate projections for South Asia from coupled model intercomparison Project-6. Sci. Data. 7, 338 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Beven, K. & Binley, A. The future of distributed models: model calibration and uncertainty prediction. Hydrol. Process. 6, 279–298 (1992).Article 
    ADS 

    Google Scholar 
    Burn, D. H. The use of resampling for estimating confidence intervals for single site and pooled frequency analysis / Utilisation d’un rééchantillonnage pour l’estimation des intervalles de confiance Lors d’analyses fréquentielles mono et multi-site. Hydrol. Sci. J. 48, 25–38 (2003).Article 

    Google Scholar 
    Terink, W., Lutz, A. F., Simons, G. W. H., Immerzeel, W. W. & Droogers, P. SPHY v2.0: Spatial processes in hydrology. Geosci. Model Dev. 8, 2009–2034 (2015).Article 
    ADS 

    Google Scholar 
    Khanal, S. et al. Variable 21st century climate change response for rivers in high mountain Asia at seasonal to decadal time scales. Water Resour. Res. 57, e2020WR029266 (2021).Article 
    ADS 

    Google Scholar 
    Kuma, P., Bender, F. A. & Jönsson, A. R. Climate model code genealogy and its relation to climate feedbacks and sensitivity. J. Adv. Model. Earth Syst. 15, e2022MS003588 (2023).Bookhagen, B. & Burbank, D. W. Toward a complete Himalayan hydrological budget: Spatiotemporal distribution of snowmelt and rainfall and their impact on river discharge. J. Geophys. Research: Earth Surf. 115, F03019 (2010).Article 
    ADS 

    Google Scholar 
    Andermann, C., Crave, A., Gloaguen, R., Davy, P. & Bonnet, S. Connecting source and transport: suspended sediments in the Nepal Himalayas. Earth Planet. Sci. Lett. 351, 158–170 (2012).Article 
    ADS 

    Google Scholar 
    Nepal, S., Krause, P., Flügel, W. A., Fink, M. & Fischer, C. Understanding the hydrological system dynamics of a glaciated alpine catchment in the Himalayan region using the J2000 hydrological model. Hydrol. Process. 28, 1329–1344 (2014).Article 
    ADS 

    Google Scholar 
    Bhattarai, S., Zhou, Y., Shakya, N. M. & Zhao, C. Hydrological modelling and climate change impact assessment using HBV light model: a case study of Narayani river Basin, Nepal. Nat. Environ. Pollution Technol. 17, 691–702 (2018).
    Google Scholar 
    Dahal, P., Shrestha, M. L., Panthi, J. & Pradhananga, D. Modeling the future impacts of climate change on water availability in the Karnali river basin of Nepal himalaya. Environ. Res. 185, 109430 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Pandey, V. P., Dhaubanjar, S., Bharati, L. & Thapa, B. R. Spatio-temporal distribution of water availability in Karnali-Mohana Basin, Western nepal: hydrological model development using multi-site calibration approach (Part-A). J. Hydrology: Reg. Stud. 29, 100690 (2020).
    Google Scholar 
    Pradhananga, S., Nepal, S., Kamal, S. K. & Hafeez, M. Climate change will exacerbate seasonal flow variability in the Karnali river basin: implications for water, energy, and agriculture sectors. Hydrol. Res. 56, 471–491 (2025).Article 

    Google Scholar 
    Moriasi, D. N., Gitau, M. W., Pai, N. & Daggupati, P. Hydrologic and water quality models: performance measures and evaluation criteria. Trans. ASABE. 58, 1763–1785 (2015).Article 

    Google Scholar 
    Immerzeel, W. W., Wanders, N., Lutz, A. F., Shea, J. M. & Bierkens, M. F. P. Reconciling high-altitude precipitation in the upper indus basin with glacier mass balances and runoff. Hydrol. Earth Syst. Sci. 19, 4673–4687 (2015).Article 
    ADS 

    Google Scholar 
    Immerzeel, W. W., Pellicciotti, F. & Shrestha, A. B. Glaciers as a proxy to quantify the Spatial distribution of precipitation in the Hunza basin. Mt. Res. Dev. 32, 30–38 (2012).Article 

    Google Scholar 
    Dhami, B., Himanshu, S. K., Pandey, A. & Gautam, A. K. Evaluation of the SWAT model for water balance study of a mountainous snowfed river basin of Nepal. Environ. Earth Sci. 77, 21 (2018).Article 
    ADS 

    Google Scholar 
    Shrestha, A. & Thakuri, P. S. Spatial and Temporal variations in water balance components of a Class-II river system in nepal: a SWAT-based analysis. H2Open J. 8, 42–58 (2025).Article 

    Google Scholar 
    Zhang, S., Zhou, Z., Peng, P. & Xu, C. A new framework for estimating and decomposing the uncertainty of climate projections. J. Clim. 37, 365–384 (2024).Article 
    ADS 

    Google Scholar 
    Apel, H., Merz, B. & Thieken, A. H. Quantification of uncertainties in flood risk assessments. Int. J. River Basin Manage. 6, 149–162 (2008).Article 

    Google Scholar 
    Kjeldsen, T. R., Lamb, R. & Blazkova, S. Uncertainty in flood frequency analysis. In Applied Uncertainty Analysis for Flood Risk Management (eds Beven, K. & Hall, J.) 153–197 (Imperial College Press, 2014).Giuntoli, I., Vidal, J. P., Prudhomme, C. & Hannah, D. M. Future hydrological extremes: the uncertainty from multiple global climate and global hydrological models. Earth Sys. Dyn. 6, 267–285 (2015).Article 
    ADS 

    Google Scholar 
    Collier, E. & Immerzeel, W. W. High-resolution modeling of atmospheric dynamics in the Nepalese himalaya. J. Geophys. Research: Atmos. 120, 9882–9896 (2015).Article 
    ADS 

    Google Scholar 
    Caretta, A. M. M. A., Arfanuzzaman, R. B. M., Morgan, S. M. R. & Kumar, M. Water. Climate change 2022: Impacts, adaptation, and vulnerability. In Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. (2022).Hock, R. et al. High mountain areas. In IPCC special report on the ocean and cryosphere in a changing climate 131–202 (eds H. O. Pörtner, DC Roberts, V. Masson-Delmotte, P. Zhai, M. Tignor, E).Rogger, M. et al. Land use change impacts on floods at the catchment scale: challenges and opportunities for future research. Water Resour. Res. 53, 5209–5219 (2017).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Beck, H. E. et al. High-resolution (1 km) Köppen-Geiger maps for 1901–2099 based on constrained CMIP6 projections. Sci. Data. 10, 724 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Malla, R., Neupane, P. R. & Köhl, M. Climate change impacts: vegetation shift of broad-leaved and coniferous forests. Trees Forests People. 14, 100457 (2023).Article 

    Google Scholar 
    Gardner, R. A. M. & Gerrard, A. J. Runoff and soil erosion on cultivated rainfed terraces in the middle hills of Nepal. Appl. Geogr. 23, 23–45 (2003).Article 

    Google Scholar 
    Arnáez, J., Lana-Renault, N., Lasanta, T., Ruiz-Flaño, P. & Castroviejo, J. Effects of farming terraces on hydrological and Geomorphological processes. A review. CATENA 128, 122–134 (2015).Article 

    Google Scholar 
    DWIDP. Annual disaster review 2014. (2015).Hannah, D. M. et al. Water and sanitation for all in a pandemic. Nat. Sustain. 3, 773–775 (2020).Article 

    Google Scholar 
    MoHA. Nepal Disaster Report: The hazardscape and vulnerability. (2009).Hawkins, E. & Sutton, R. The potential to narrow uncertainty in regional climate predictions. Bull. Am. Meteorol. Soc. 90, 1095–1108 (2009).Article 
    ADS 

    Google Scholar 
    Wu, Y. et al. Quantifying the uncertainty sources of future climate projections and narrowing uncertainties with bias correction techniques. Earth’s Future. 10, e2022EF002963 (2022).Article 
    ADS 

    Google Scholar 
    Sinha, R. & Kosi Rising Waters, dynamic channels and human disasters. Economic Political Wkly. 43, 42–46 (2008).
    Google Scholar 
    Dinesh Kumar Mishra. The Bihar flood story. Economic Political Wkly. 32, 2206–2217 (1997).
    Google Scholar 
    Huffman, G. et al. J. NASA global precipitation measurement (GPM) – integrated multi-satellitE retrievals for GPM (IMERG). 38 (2019).Arias-Hidalgo, M., Bhattacharya, B., Mynett, A. & Van Griensven, A. Experiences in using the TMPA-3B42R satellite data to complement rain gauge measurements in the Ecuadorian coastal foothills. Hydrol. Earth Syst. Sci. 17, 2905–2915 (2013).Article 
    ADS 

    Google Scholar 
    Weedon, G. et al. Creation of the WATCH forcing data and its use to assess global and regional reference crop evaporation over land during the twentieth century. J. Hydrometeorol. 12, 823–848 (2011).Article 
    ADS 

    Google Scholar 
    Weedon, G. P. et al. The WFDEI meteorological forcing data set: WATCH forcing data methodology applied to ERA-Interim reanalysis data. Water Resour. Res. 50, 7505–7514 (2014).Article 
    ADS 

    Google Scholar 
    Le Coz, J., Renard, B., Bonnifait, L., Branger, F. & Le Boursicaud, R. Combining hydraulic knowledge and uncertain gaugings in the Estimation of hydrometric rating curves: A bayesian approach. J. Hydrol. 509, 573–587 (2014).Article 

    Google Scholar 
    Neitsch, S. L., Arnold, J. G., Kiniry, J. R. & Williams, J. R. Soil and Water Assessment Tool Theoretical Documentation Version 2009. (2011).Hargreaves, G. H. & Samani, Z. A. Reference crop evapotranspiration from temperature. Appl. Eng. Agric. 1, 96–99 (1985).Article 

    Google Scholar 
    Allen, R. G., Pereira, L. S., Raes, D. & Smith, M. Crop evapotranspiration-Guidelines for computing crop water requirements-FAO irrigation and drainage paper 56. Fao Rome. 300, D05109 (1998).
    Google Scholar 
    Kokkonen, T., Koivusalo, H., Jakeman, T. & Norton, J. Construction of a degree-day snow model in the light of the ten iterative steps in model development. (2006).Hock, R. Temperature index melt modelling in mountain areas. J. Hydrol. 282, 104–115 (2003).Article 
    ADS 

    Google Scholar 
    McKay, M. D., Beckman, R. J. & Conover, W. J. A comparison of three methods for selecting values of input variables in the analysis of output from a computer code. Technometrics 21, 239–245 (1979).MathSciNet 

    Google Scholar 
    Legates, D. R. & McCabe, G. J. Evaluating the use of goodness-of-fit measures in hydrologic and hydroclimatic model validation. Water Resour. Res. 35, 233–241 (1999).Article 
    ADS 

    Google Scholar 
    Liu, Y., Freer, J., Beven, K. & Matgen, P. Towards a limits of acceptability approach to the calibration of hydrological models: extending observation error. J. Hydrol. 367, 93–103 (2009).Article 
    ADS 

    Google Scholar 
    Hosking, J. R. M. & Wallis, J. R. Regional Frequency analysis: an Approach Based on L-Moments https://doi.org/10.1017/CBO9780511529443 (Cambridge University Press, 1997).Hosking, J. R. M. & L-Moments Analysis and Estimation of distributions using linear combinations of order statistics. J. Royal Stat. Soc. Ser. B (Methodological). 52, 105–124 (1990).Article 
    MathSciNet 

    Google Scholar 
    Jarvis, A., Reuter, H., Nelson, A. & Guevara, E. Hole-filled SRTM for the globe Version 4. (2008).Download referencesAcknowledgementsWe thank the Advanced Research Computing Unit of Durham University for their provision of and support with high-performance computing resources. We also acknowledge two anonymous reviewers whose comments have improved the manuscript.FundingThis research was funded by the Charles Wilson Doctoral Studentship.Author informationAuthors and AffiliationsDepartment of Geography, Durham University; Institute of Hazard, Risk and Resilience, Durham University, Durham, UKIvo Pink, Sim M. Reaney & C. Isabella BovoloDepartment of Geography, Durham University, Durham, UKRichard J. HardyAuthorsIvo PinkView author publicationsSearch author on:PubMed Google ScholarSim M. ReaneyView author publicationsSearch author on:PubMed Google ScholarRichard J. HardyView author publicationsSearch author on:PubMed Google ScholarC. Isabella BovoloView author publicationsSearch author on:PubMed Google ScholarContributionsI.P. performed the simulations and analysed the model results. S.R provided modelling tools. All authors equally participated in generating ideas, designing the research project, and editing various versions of the manuscript.Corresponding authorCorrespondence to
    Ivo Pink.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary InformationBelow is the link to the electronic supplementary material.Supplementary Material 1Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
    Reprints and permissionsAbout this articleCite this articlePink, I., Reaney, S.M., Hardy, R.J. et al. Increased rainfall-runoff drives flood hazard intensification in Central Himalayan river systems.
    Sci Rep 15, 42277 (2025). https://doi.org/10.1038/s41598-025-26815-2Download citationReceived: 21 July 2025Accepted: 30 October 2025Published: 08 December 2025Version of record: 08 December 2025DOI: https://doi.org/10.1038/s41598-025-26815-2Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Global dataset of sand dam features and geographical distribution across drylands

    AbstractSand dams are water infrastructure, built across ephemeral sandy rivers, that increase water supply by creating an artificial sandy aquifer upstream of the dam. Despite their effectiveness and recent traction in the research and development arena, empirical data on their distribution and characteristics are scattered and largely unreported. This gap represents a major barrier for understanding the large-scale potential of such a Nature-based Solution and for planning new installations. This paper presents a global dataset of sand dam locations and dimensions, developed collaboratively by research and development experts. We collected sand dam information from several sources, including local sand dam organizations. The data was reviewed and integrated through visual inspection in Google Earth. Although most georeferenced sand dams are from Eastern and Southern Africa, this dataset is a first global inventory and represents an invitation for others working in sand dams around the world to contribute their data. The dataset supports research on the effectiveness of sand dams and can aid practitioners with science-based criteria for sand dam development.

    Similar content being viewed by others

    The Global Dam Watch database of river barrier and reservoir information for large-scale applications

    Article
    Open access
    08 October 2024

    Assessment of suitable sub-watersheds for sand dams and prioritization of artificial sandy aquifers for sustainable groundwater development

    Article
    Open access
    26 September 2025

    Global Dam Tracker: A database of more than 35,000 dams with location, catchment, and attribute information

    Article
    Open access
    23 February 2023

    Background and SummarySmall water infrastructure, such as ponds and small dams, are fundamental to support agricultural activities, especially in rangelands, where they often represent the only reliable water sourcE1. Sand dams are a particularly promising water infrastructure for drylands, since they provide specific benefits compared to conventional small dams2. In fact, sand dams are concrete walls built across ephemeral streams, which are specifically designed to trap sandy sediments upstream of the dam, thus creating an artificial aquifer able to increase shallow groundwater resources. The main elements of a sand dams are (i) the dam wall, which is anchored to the bedrock (or to a deep compacted clay layer) and raised for at most 2 m above the sand, (ii) spillways to allow runoff to flow downstream together with fines sediments and wings to anchor the structure to lateral rocks (or anchored deeper into the terrain). The water contained within the sand pores is protected from evaporation and, to some extent, from surface contamination, making sand dams a nature-based solution to water scarcity in drylands3. Although the first known sand dams date back to the early 1900s4, their construction is increasing exponentially over the last couple of decades thanks to the effort of several non-governmental organizations (NGOs) working on sustainable rural development.It is helpful to differentiate and contextualize the sand dams within the broader family of in-channel water harvesting structures, which all harness seasonal flows to improve water availability in drylands. Among the most comparable are check dams, subsurface dams, and water spreading weirs, each with different strengths and limitations depending on local hydrology, terrain, resources and intended use.Check dams are widely used for erosion control and some groundwater recharge. They are normally relatively simple and low-cost to construct, and due to their usual small sizes, they have an impact normally if several are constructed in the same area. They effectively reduce runoff speed, promote sediment deposition and landscape stabilization, and enhance some infiltration along the streambed and adjacent banks5,6. Unlike sand dams, they do not typically store significant volumes of accessible water for dry season use, as this is normally not its purpose.Subsurface (or underground) dams, constructed beneath the riverbed, are highly effective at blocking the downstream flow of groundwater and storing water below the surface7. One of their advantages is structural flexibility—they do not require above-ground stability, making them suitable for wider channels and more downstream locations, so they normally have a larger storage capacity than sand dams, which in turn can potentially have more upstream and downstream impacts. Subsurface dams typically require more extensive excavation, and, unlike sand dams, they do not rely on sediment accumulation to form a surface reservoir. As a result, their ability to facilitate lateral connectivity with bank storage depends largely on the geomorphology of the channel and surrounding banks. Accessing water from subsurface dams generally requires deeper wells or more robust pumping systems, making them less accessible and more costly for community use.Last, water spreading weirs are designed to slow floodwaters and spread them laterally across the floodplain, enhancing soil moisture, supporting vegetation regrowth and flood-based agriculture, and recharging shallow aquifers8. Their strength lies in improving agricultural potential and landscape productivity over a broad area. However, they do not store water directly within the channel and are not intended for abstraction.A renewed interest in sand dam structures arose in academia, with several new studies and reviews addressing the many research gaps on sand dams’ hydrological and socio-economic aspects2,9,10,11.However, both research and practice on sand dams are hindered by limited data availability on sand dams’ locations, characteristics, and impacts. The main reason for this data void is because most historical structures were not reported or monitored after they were built12, and recent sand dam development is driven by NGOs, which often do not systematically share their records. While large-scale datasets are often available for traditional dams of varying dimensions, including large and medium dams, both at the national and global scale13,14,15, no datasets have been produced for sand dams, despite representing a pivotal water infrastructure in drylands, where water is most needed. The reasons for this data gap are that i) the majority of sand dams are built by local communities, thus data are not systematically reported, ii) they traditionally have small dimensions, making their identification from satellite images difficult. In fact, most traditional dam datasets are derived from remote sensing of open water bodies, which is impossible for sand dams, since they do not store visible open water except for very short periods (often just a few days per year).Here, we present an extensive and revised dataset with 1006 records of sand dams across the world, the Global Sand Dams Dataset (GSDD). We gathered data from different available sources and developed a network of researchers and practitioners to share, enrich and revise information on dams’ characteristics, such as dams’ crest length, throwback, and stream width. The scope of the dataset is to provide open access data to boost research and improve the understanding of sand dam suitability and impact, which can eventually support the implementation of new sand dam construction programs. The dataset can also serve as a resource for those interested in exploring various factors that influence the success of development projects, including questions from anthropology and other socially engaged research domains. As remote sensing data availability increases and artificial intelligence-supported analysis techniques expand, the dataset also offers an opportunity for researchers across the spectrum of earth, social, and agricultural sciences to conduct novel research into rural food and water security in drylands. Research-quality data from rural areas in the Global South remains rare or at the very least, inaccessible, and this dataset seeks to provide an example of how such data can be stored and shared, in accordance with the Findable, Accessible, Interoperable, and Reusable (FAIR) data principles. The launch of the dataset is also an open call to anyone interested in sharing further data on sand dams, providing a platform for systematically collecting and harmonising sand dams reporting standards to advance research, sharing knowledge and foster implementation and replication/learning from previous experiences.MethodsIn this section, we i) present an overview of the dataset production procedure, ii) describe the data sources, iii) report the data harmonization and enrichment process, and present the GSDD attributes and features.Dataset production procedureThe collection, development, and technical validation process of the GSDD is concisely displayed in Fig. 1 and explained in detail in the sections below.Fig. 1Steps involved in the production of the GSDD.Full size imageThe working group started in May 2021 during an online workshop on sand dam research organised by the Water Harvesting Lab of the University of Florence, Italy (Castelli et al. 2022). The working group later expanded, including Sand Dams Worldwide and Dabane Trust Water Workshop, which are two major NGOs working on the construction of sand dams.A first set of 110 records was collected by merging all data points provided by the working group of international researchers and collaborators based on sand dams they have personally worked on, either during the construction process or during research projects. Since most of those 110 records were shared by the NGOs involved in the funding and construction of sand dams, we opened the collaboration on the dataset to the major NGOs working with sand dams. At this stage we collected the majority of the records in the GSDD, which was later finalized with the integration of records from scientific and grey literature, and open sources online platform. Given the diverse data sources, we designed a multiple step technical validation procedure involving a careful screening and visual inspection of all records and a double check of a random sample of the dataset. These steps are described in detail in the following sections.Data sourcesThe data sources are provided in the attribute “Source” of the dataset. The aim of the GSDD is to make core information on sand dams’ location and characteristics available for researchers and practitioners to eventually improve their spread and effectiveness. Since most sand dams are built by NGOs, the main data sources are the NGOs involved in the funding, planning or construction, mostly in rural communities. The main international NGOs providing data are Sasol Foundation (https://www.sasolfoundation.com/) and Sand Dams Worldwide (https://www.sanddamsworldwide.org.uk/ – previously Africa Sand Dams Foundation), while Dabane Trust (https://dabane.org/) is located in Zimbabwe and No One Out in their local hub in Uganda (https://www.nooneout.org/). These NGOs account for most of the records in the GSDD, which they provided as csv with the available information based on the attribute of the GSDD that we shared with them as a template in advance.The minimum requirement for the records to be included in the dataset is the geographic location coordinate. However, many records collected along the process have more details. For example, the initial set of 110 records shared by the researchers of the GSDD team also have information on construction date, presence, and type of water access (e.g. scoop holes or solar pumps etc.) or purpose of the sand dams (e.g. domestic use or pastoral use etc.). Additionally, the 110 records have additional information on water quality, because of the analysis conducted during a collaboration between the NGO Africa Sand Dam Foundation and researchers from the GSDD team. Although none of the other data collected from other sources have such information on water quality, we kept those attributes as a blueprint for future research.We also included the very little freely available data on sand dams from online open access sources, including two records in the Global Database on Sustainable Land Management (WOCAT16, located in Kenya and three records from the Global Inventory of Managed Aquifer Recharge applications by the International Groundwater Resources Assessment Centre (https://un-igrac.org/our-work/activities/global-inventory-of-managed-aquifer-recharge-schemes/). From these sources we only kept the location and the construction date, before enriching them with the other attributes described in the next sections.From scientific and grey literature there is mention of sand dams in several countries, including Namibia, India, and South Korea4,10, but locations remain mostly unknown. From these sources, we created records of the sand dams with known locations and cited the source in the “source” attribute either as a scientific reference or as a website. Finally, we have checked the Global Dam Watch (GDW) database17 (i.e. the most complete dataset of river barriers and reservoirs) for sand dam structures. Since we found no overlap with our GSDD, we can reasonably deduce that the GDW does not include sand dams records.Dataset harmonizationBecause of the diversity of data sources and reporting standards of the preliminary dataset, a thorough harmonization process was designed and implemented to 1) avoid redundancy in the records, 2) provide a coherent structure, aligning all records and setting common attributes, and 3) check and adjust the reported location (LAT and LON coordinates) of the dams. Since the whole compilation (and review) was based on visual inspection on Google Earth, thus subject to the researchers’ judgement, we organized internal online workshops among the core team of co-authors involved in database compilation and review to ensure consistency across records. During these interactions we collectively defined a set of shared criteria and steps for verifying and enriching records. For example, we agreed on consistent visual inspection procedures using Google Earth to confirm the presence and characteristics of each sand dam. We also established specific guidelines on how to measure dam crest length, throwback distance, and stream width — ensuring that all researchers interpreted and measured these attributes in a uniform way (like throwback measured as straight-line distance upstream from the dam, width measured at regular intervals along the throwback).Specifically, if the sand dam could not be located on Google Earth, the reason was provided as:

    a.

    Not visible from Google Earth – either if the sand dam could not be found or if the quality of the satellite image was not sufficient (e.g. cloud coverage or very low resolution).

    b.

    No image available on Google Earth – the sand dam was built recently, and no satellite image was available.

    c.

    Off-stream location – the coordinates were located off a waterway, and presumed to be inaccurate, since no dams were found in the surrounding area.

    d.

    Broken – the sand dam was reported as broken or the dam is visibly broken from the satellite inspection, either collapsed or partly damaged.

    e.

    Water pond – the coordinate point to a structure which holds water all year round.

    4. If additional sand dams were identified during the process, they were added in the dataset.5. Add comments about the sand dam, if needed.6. Enrich the dataset with additional information, including the following dam and stream characteristics, which are key for estimating the sand and water storage potential:

    a.

    The length of the dam structure, including the dam’s main doby and the visible portion of the wings

    b.

    The throwback, which is the longest straight line from the dam’s body upstream.

    c.

    The average width of the river upstream of the dams along the throwback.

    The dam and stream characteristics were retrieved by visual inspection from Google Earth. Although there are other high resolution satellite products, we use Google Earth, since it is the highest resolution tool among the freely available satellite tools, and widely used in peer-reviewed scientific dataset14,18,19. In fact, it allows for a transparent methodology and easy replicability or check from third parties. Additionally, the “time machine function” on Google Earth was instrumental for estimating the potential construction date of sand dams (i.e. picking the date of the image when the dam is first visible), when this information was missing from the original source.Data RecordsDetails of the attributes included in the GSDD20 are reported in Supplementary Table 1. As the core information of the GSDD are the location and characteristics of the sand dams, the geographic coordinates (i.e. LAT and LON), dam length, throwback and river width are the only complete attributes for all the records. We added a “comment” column in which additional information on data quality or uncertainty in the evaluation of sand dams’ characteristics are reported. For transparency, we also maintained comments related to the review process. The other attributes report information retrieved from the original sources; hence they are available for a subset of records. Values in the “construction date” column mostly consider the original source, except those estimated using Google Earth.The dataset includes 1006 records, spanning across 15 countries and 3 continents (Africa, Asia, and South America), which were built over the last 80 years (Fig. 2). Most of the dams are in Kenya (892), including some of the oldest records from 1952. Although this seems to portray a skewed representation of the global spread of sand dams, it is in line with current scientific literature, which show that most sand dams are actually located in Kenya. For example, Ritchie et al., (2021) states that “the overwhelming majority of sand dams have been built in South-eastern Kenya, perhaps due to unique and favourable physiographic features, and thus the majority of sand dam research has also been conducted in Kenya”. Our dataset widens the range of countries hosting sand dams by including records from literature and research outside Kenya to provide a more complete picture of sand dams global spread and potential. Table 1, in fact, shows that Angola, India, Tanzania and Zimbabwe have over 10 sand dams in their territory.Fig. 2Spatial global distribution of the sand dam records of the GSDD, compared to countries with known presence of sand dams and potential countries suitable for sand dams according to (Yifru et al.10) (a) and the temporal trend of sand dam presence in the period 1952–2023 (b).Full size imageTable 1 Country statistics of total number, mean and standard deviation (sd) values of sand dam characteristics.Full size tableThe median dimensions of the dams in the GSDD are 32 m of dam length, 17 m of stream width (upstream of the dams) and 130 m of throwback, although some sand dams show greater dimensions (Fig. 3). For example, the two sand dams in China are about 150 m in length, built across a stream of about the same width and quite straight, with about 3 km of throwback. The median country statistics are reported in Table 1.Fig. 3Median dams’ characteristics, including throwbacks, stream width, and dam dimensions (a), photo of a sand dam shared by Sand Dams Worldwide and its characteristics (b) and a snapshot with dam’s characteristics from Google Earth (c).Full size imageTechnical ValidationA core team of 10 co-authors had volunteered to perform the technical validation of the dataset. The total records were split equally between the 10 reviewers, who checked all records individually on Google Earth.Although the review process was very detailed and carefully coordinated with the internal workshops, it was still subject to the reviewers’ judgement. To limit uncertainty in records evaluation, a second review round was conducted on a subset of data including i) a set of records judged particularly unclear by the reviewers during the first review round and ii) a randomly selected subset of 200 records (accounting for about 20% of all record), for a total of 263 records. The second review round followed the same procedure as the first review except that it was conducted by different researchers. On average, about 80% of the records were confirmed in the second round, showing a satisfactory homogeneity in the imagery interpretation.Dataset limitationsThis dataset is just a first step towards providing a complete picture of the global distribution of sand dams, although several challenges might hinder a comprehensive assessment. In fact, many sand dam projects, especially those initiated by farmers, local communities, or small NGOs, often lack formal monitoring and evaluation mechanisms, resulting in their exclusion from broader development databases and scientific literature. Many local sand dam projects may not have digital records or may be documented in formats that are not easily accessible or shareable.These local or small-scale initiatives may not attract media or scientific attention, particularly if they are in remote or under-represented areas. Documentation and formal reporting efforts typically focus on larger, high-profile projects funded by international organisations or governments, overshadowing the achievements of smaller, community-led endeavours. Linked to this, efforts to gather data may be concentrated in regions with higher visibility or funding incentives for these topics. Fragmented data systems at local, national, and international levels further contribute to the invisibility of these projects.Additionally, language barriers may impede information sharing about sand dams, with documentation often existing in local languages that do not reach the global community. Moreover, the terminology used to describe sand dams can vary across regions and organisations. Terms like “sand dam”, “subsurface dam”, “sand storage dam”, “ground-water dams”21 or even local names in regional languages may be used interchangeably to refer to similar structures, or these might be even referring to different structures than those inventoried in this work. Depending on the geographical region, there may be localised terms or colloquial expressions used to describe sand dams. An example is the term “barragens”, which in Angolan Portuguese refers to both conventional dams and sand dams interchangeably, while the local term “Chimpaca” is mostly used by Angolan pastoralists to refer to sand dams or cattle water ponds9. The language issue might also fuel misunderstanding between different water infrastructure and their spread. For example, of the many small dams in India documented by Yifru et al.10, the vast majority seem to be check dams or subsurface dams, as mentioned by22, but sometimes included in the geographical domain of sand dams. Moreover, even differences in approaches for construction can lead to labelling the infrastructures as different techniques, making it difficult to standardise data collection and reporting. The researchers used a variety of multilingual keywords and search terms relevant to sand dams in different languages, but the diversity of regional languages and dialects may still result in incomplete data capture.This paper is an initiative to highlight the importance of including local projects that often go unrecognised. The authors aim to motivate others to also incorporate these smaller or community-led interventions into the dataset. Such inclusion and updates are necessary to render the dataset adaptable, transcending its status as a static depiction of sand dam locations.Beyond the challenge of compiling cases, the dataset does not include in-situ measurements to refine the assessment of potential for water harvesting, nor performance metrics to understand if the sand dams are properly functioning (e.g. storing sandy sediments) or fulfilling their intended purposes.Potential Use and future development of the datasetNotwithstanding the mentioned limitations, the GSDD offers a first usable tool to support researchers and practitioners in several endeavours. Researchers can use the GSDD to explore numerous research questions, which are at the forefront of hydrology research in water infrastructures. Some potential macro-research questions include (i) exploring the hydrological, ecological or socio-economic impacts of multiple sand dams at the catchment scale, including effects of upstream-downstream connectivity and water availability, (ii) assessing the effect of sand dams location and development on land use/cover change and social-ecological dynamics of agro-pastoralists, (iii) estimating the potential for water storage increase, or groundwater recharge, and its potential use, (iv) improving or validating large-scale best siting approaches and (v) evaluating socio-economic impacts and potential of sand dams implementations. The mentioned potential research frontiers are in line with recent collaborative scientific work.While the research advances enabled by GSDD can support practitioners in improving sand dam implementation and effectiveness, the dataset can also be used as an operational tool, for (i) exploring new areas for sand dam projects based on geographical similarity and gaps, (ii) assessing the overall impacts of past and current sand dam projects and (iii) estimating dimensional characteristics of construction and its costs.Although the majority of records currently come from Kenya, the dataset is not restricted in scope to this context. In recent years, sand dams have been increasingly promoted and implemented beyond Kenya, driven by international NGOs such as Sand Dams Worldwide and other global actors. Projects have been piloted or expanded in diverse socio-ecological settings, including South Korea23, Zimbabwe24 and Mexico25, among others. Some of these emerging structures are not yet systematically mapped, but their presence illustrates the rapid and ongoing spread of the technology. The very motivation behind the GSDD is therefore to provide a transparent, flexible, and extensible evidence base that not only documents the Kenyan experience but also accommodates and supports the integration of new data from other regions as it becomes available. In this sense, while Kenya dominates the current dataset, the GSDD contributes to the global research and development of sand dams, enabling comparative analysis, supporting scaling out in different contexts, and fostering co-learning for their sustainable implementation.To provide further information in support of the mentioned objectives and beyond, a useful improvement of the database in a future publication could include a variety of hydroclimatic and socio-economic data for each data point, taken from local or global sources. For the hydroclimatic indicators, global data from TerraClimate, for example, can be used to extract variables such as runoff, precipitation minus evapotranspiration or climatic water deficit, helping assess water availability and the potential for water storage across different regions.On the socioeconomic side, integrating globally available data can support characterizing community vulnerability in the environments in which dams are built, and potential impact of them in provision of water security. For example, population density from WorldPop could help estimate the number of actual or potential beneficiaries. The Multidimensional Poverty Index or livelihood information from UNHCR, for instance, could be integrated to provide insight into economic vulnerability. Data on access to improved water sources – for example coming from the WHO/UNICEF – can help contextualize the role of sand dams in improving water access and supporting climate-sensitive livelihoods.Finally, the GSDD has the ambition to provide researchers and practitioners with a platform to improve data collection and monitoring of sand dams by expanding the network of contributors and users of the dataset. The ambition of the GSDD working group is to welcome NGOs and researchers in sharing data and experience on sand dam projects, contributing to improving the dataset’s comprehensiveness and quality with time, while supporting its use in research and implementation projects.

    Data availability

    The GSDD is an open-source platform, which can be expanded by integrating data from other global areas. The dataset is available on Zenodo, and it can be downloaded using the following URL: https://zenodo.org/records/15828863.
    Code availability

    The data of the GSDD were processed and mapped using the software RStudio and QGIS. The R code is publicly available here: https://github.com/piemonteseluigi/Code_GSDD/tree/78147b27f26b08b6854c47e829f244e9abc2bdb0.
    ReferencesPiemontese, L. et al. Over-reliance on water infrastructure can hinder climate resilience in pastoral drylands. Nat. Clim. Change 14, 267–274 (2024).Article 
    ADS 

    Google Scholar 
    Ritchie, H., Eisma, J. A. & Parker, A. Sand Dams as a Potential Solution to Rural Water Security in Drylands: Existing Research and Future Opportunities. Front. Water 3 (2021).Castelli, G. et al. Sand dams for sustainable water management: Challenges and future opportunities. Sci. Total Environ. 156126 https://doi.org/10.1016/j.scitotenv.2022.156126 (2022).Wipplinger, O. Sand storage dams in South West Africa. Civ. Eng. Siviele Ingenieurswese 1974, 135–136 (1974).
    Google Scholar 
    Abbasi, N. A., Xu, X., Lucas-Borja, M. E., Dang, W. & Liu, B. The use of check dams in watershed management projects: Examples from around the world. Sci. Total Environ. 676, 683–691 (2019).Article 
    CAS 
    PubMed 
    ADS 

    Google Scholar 
    Lucas-Borja, M. E., Piton, G., Yu, Y., Castillo, C. & Antonio Zema, D. Check dams worldwide: Objectives, functions, effectiveness and undesired effects. CATENA 204, 105390 (2021).Article 

    Google Scholar 
    Jamali, I. A., Olofsson, B. & Mörtberg, U. Locating suitable sites for the construction of subsurface dams using GIS. Environ. Earth Sci. 70, 2511–2525 (2013).Article 
    ADS 

    Google Scholar 
    Amede, T. et al. Facilitating livelihoods diversification through flood-based land restoration in pastoral systems of Afar, Ethiopia. Renew. Agric. Food Syst. (2020).Piemontese, L., Castelli, G., Limones, N., Grazio, A. & Bresci, E. Large-scale siting of sand dams: A participatory approach and application in Angolan drylands. Land Degrad. Dev. 34, 844–858 (2023).Article 

    Google Scholar 
    Yifru, B. A. et al. Water Storage in Dry Riverbeds of Arid and Semi-Arid Regions: Overview, Challenges, and Prospects of Sand Dam Technology. Sustainability 13, 5905 (2021).Article 
    ADS 

    Google Scholar 
    Ryan, C. & Elsner, P. The potential for sand dams to increase the adaptive capacity of East African drylands to climate change. Reg. Environ. Change 16, 2087–2096 (2016).Article 

    Google Scholar 
    Eisma, J. A. & Merwade, V. A data-driven approach to assessing the impact of water harvesting structures on regional water storage in East Africa. J. Hydroinformatics 23, 352–367 (2021).Article 

    Google Scholar 
    Speckhann, G. A., Kreibich, H. & Merz, B. Inventory of dams in Germany. Earth Syst. Sci. Data 13, 731–740 (2021).Article 
    ADS 

    Google Scholar 
    Mulligan, M., van Soesbergen, A. & Sáenz, L. GOODD, a global dataset of more than 38,000 georeferenced dams. Sci. Data 7, 31 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Paredes-Beltran, B., Sordo-Ward, A. & Garrote, L. Dataset of Georeferenced Dams in South America (DDSA). Earth Syst. Sci. Data 13, 213–229 (2021).Article 
    ADS 

    Google Scholar 
    Liniger, H. et al. Achieving land degradation neutrality: The role of SLM knowledge in evidence-based decision-making. Environ. Sci. Policy 94, 123–134 (2019).Article 

    Google Scholar 
    Lehner, B. et al. The Global Dam Watch database of river barrier and reservoir information for large-scale applications. Sci. Data 11, 1069 (2024).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Shah, J., Hu, J., Edelenbosch, O. Y. & van Vliet, M. T. H. Global dataset combining open-source hydropower plant and reservoir data. Sci. Data 12, 646 (2025).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zhang, A. T. & Gu, V. X. Global Dam Tracker: A database of more than 35,000 dams with location, catchment, and attribute information. Sci. Data 10, 111 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Piemontese, L. et al. Global Sand Dams Dataset. Zenodo https://doi.org/10.5281/zenodo.15828863 (2025).Hanson, G. & Nilsson, Å. Ground-Water Dams for Rural-Water Supplies in Developing Countries. Groundwater 24, 497–506 (1986).Article 

    Google Scholar 
    Parker, A. H. et al. The multiple uses of water derived from managed aquifer recharge systems in Kenya and India. J. Water Sanit. Hyg. Dev. 12, 208–216 (2022).Article 

    Google Scholar 
    Kim, I. H., Chang, S. W., Kim, M.-G. & Chung, I.-M. Effects of a bypass-type multistage sand-filled dam on water supply in a drought-vulnerable area in Mullocheon catchment, South Korea. J. Hydrol. Reg. Stud. 56, 101997 (2024).Article 

    Google Scholar 
    Lazurko, A. et al. Assessing sand dams for contributions to local water security and drought resilience in the semi-arid eastern Shashe catchment, Zimbabwe. Reg. Environ. Change 24, 36 (2024).Article 

    Google Scholar 
    Villarreal-Chávez, G. Impact of the construction of a sand dam on the benthic ecology of an estuary in the mexican north Pacific. Cienc. Mar. 17, 105–119 (1991).Article 

    Google Scholar 
    Download referencesAuthor informationAuthors and AffiliationsDepartment of Agriculture, Food, Environment and Forestry (DAGRI), University of Florence, Florence, ItalyLuigi Piemontese, Lorenzo Villani, Giulio Castelli & Elena BresciDepartment of Water and Climate (HYDR), Vrije Universiteit Brussels, Brussels, BelgiumLorenzo VillaniDepartment of Physical Geography and Regional Geographical Analysis, University of Seville, Seville, SpainNatalia LimonesInstitute for Environmental Studies (IVM), VU University Amsterdam, Deltares Institute, Delft, NetherlandsJeroen. C. J. H. AertsUNESCO Chair in Hydropolitics, University of Geneva, Geneva, SwitzerlandGiulio CastelliInstitute for Environmental Sciences (ISE), University of Geneva, Geneva, SwitzerlandGiulio CastelliDepartment of Civil Engineering, University of Texas at Arlington, Arlington, Texas, USAJessica A. EismaDabane Water Workshops, Bulawayo, ZimbabweBongani MpofuDepartment of Biology, Eastern Mennonite University, Harrisonburg, Virginia, USADoug Graber NeufeldSchool of Water, Energy and Environment, Cranfield University, Cranfield, UKHannah RitchieDepartment of Environmental Science, Auckland University of Technology, Aotearoa, New ZealandCate RyanDepartment of Civil Engineering and Construction Studies, Atlantic Technological University Sligo, Sligo, IrelandRuth QuinnSand Dams Worldwide, London, UKChristine WhinneyAuthorsLuigi PiemonteseView author publicationsSearch author on:PubMed Google ScholarLorenzo VillaniView author publicationsSearch author on:PubMed Google ScholarNatalia LimonesView author publicationsSearch author on:PubMed Google ScholarJeroen. C. J. H. AertsView author publicationsSearch author on:PubMed Google ScholarGiulio CastelliView author publicationsSearch author on:PubMed Google ScholarJessica A. EismaView author publicationsSearch author on:PubMed Google ScholarBongani MpofuView author publicationsSearch author on:PubMed Google ScholarDoug Graber NeufeldView author publicationsSearch author on:PubMed Google ScholarHannah RitchieView author publicationsSearch author on:PubMed Google ScholarCate RyanView author publicationsSearch author on:PubMed Google ScholarRuth QuinnView author publicationsSearch author on:PubMed Google ScholarChristine WhinneyView author publicationsSearch author on:PubMed Google ScholarElena BresciView author publicationsSearch author on:PubMed Google ScholarContributionsAll authors co-designed the study. L.P., L.V., N.L., J.C.J.H.A., G.C., J.E., B.M., D.G.N., H.R., C.R., R.Q. and C.W. collected and reviewed the dataset. L.P. coordinated the dataset production and led the writing of the paper. L.V. and N.L. contributed substantially to the writing, while all authors contributed to reviewing the writing. J.A.E. provided funding for publication. E.B. supervised the work.Corresponding authorCorrespondence to
    Luigi Piemontese.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
    Reprints and permissionsAbout this articleCite this articlePiemontese, L., Villani, L., Limones, N. et al. Global dataset of sand dam features and geographical distribution across drylands.
    Sci Data 12, 1929 (2025). https://doi.org/10.1038/s41597-025-06197-wDownload citationReceived: 28 November 2024Accepted: 22 October 2025Published: 08 December 2025Version of record: 08 December 2025DOI: https://doi.org/10.1038/s41597-025-06197-wShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    An agile benchmarking framework for wastewater resource recovery technologies

    AbstractWater resource recovery facilities (WRRFs) face growing pressures to balance compliance, sustainability, and cost while adapting to evolving treatment needs. To support research, development, and deployment (RD&D) of innovative technological solutions, we developed an open-access benchmarking framework comprised of 18 plant-wide simulation models. Implemented in QSDsan, the framework is validated against GPS-X™ simulations while capturing distinct system behaviors, treatment performance, energy demand, and operational costs across diverse designs. It offers a rigorous and transparent foundation for comparative technology evaluations, guiding RD&D decision-making and advancing sustainable water management.

    Data availability

    The datasets generated and/or analyzed during the current study are available in the EXPOsan repository, https://github.com/QSD-Group/EXPOsan/tree/main/exposan/werf/publication_data.
    Code availability

    The underlying code for this study is available on Github and can be accessed via this link: https://github.com/QSD-Group/EXPOsan/tree/main/exposan/werf.
    ReferencesMcCarty, P. L., Bae, J. & Kim, J. Domestic wastewater treatment as a net energy producer–can this be achieved?. Environ. Sci. Technol. 45, 7100–7106 (2011).
    Google Scholar 
    Guest, J. S. et al. A new planning and design paradigm to achieve sustainable resource recovery from wastewater. Environ. Sci. Technol. 43, 6126–6130 (2009).
    Google Scholar 
    Larsen, T. A., Alder, A. C., Eggen, R. I. L., Maurer, M. & Lienert, J. Source separation: will we see a paradigm shift in wastewater handling?. Environ. Sci. Technol. 43, 6121–6125 (2009).
    Google Scholar 
    Henze, M., Gujer, W., Mino, T. & van Loosedrecht, M. Activated sludge models ASM1, ASM2, ASM2d and ASM3. https://doi.org/10.2166/9781780402369 (IWA Publishing, 2006).Batstone, D. J. et al. The IWA anaerobic digestion model No 1 (ADM1). Water Sci. Technol. 45, 65–73 (2002).
    Google Scholar 
    Alex, J. et al. Benchmark simulation model no. 1 (BSM1). Rep. IWA Taskgroup Benchmarking Control Strateg. WWTPs 1, (2008).Jeppsson, U. et al. Benchmark simulation model no 2: general protocol and exploratory case studies. Water Sci. Technol. 56, 67–78 (2007).
    Google Scholar 
    Gernaey, K. V., Jeppsson, U., Vanrolleghem, P. A. & Copp, J. B. Benchmarking of Control Strategies for Wastewater Treatment Plants. https://doi.org/10.2166/9781780401171 (IWA Publishing,2014).wwtmodels. wwtmodels/Benchmark-Simulation-Models. (2025).Hydromantis. GPS-X Technical Reference – v8.0. https://www.hydromantis.com/help/GPS-X/docs/8.0/Technical/index.html#CHAPTER%2012.Products – EnviroSim. https://envirosim.com/products.Li, Y. et al. QSDsan: an integrated platform for quantitative sustainable design of sanitation and resource recovery systems. Environ. Sci. Water Res. Technol. 8, 2289–2303 (2022).
    Google Scholar 
    Tarallo, S., Shaw, A., Kohl, P. & Eschborn, R. A Guide to Net-Zero Energy Solutions for Water Resource Recovery Facilities. https://iwaponline.com/ebooks/book/293/ (2015).US EPA. Clean Watersheds Needs Survey (CWNS) – 2004 Report and Data (2015).US EPA. Clean Watersheds Needs Survey (CWNS) – 2008 Report and Data (2015).US EPA. Clean Watersheds Needs Survey (CWNS) – 2012 Report and Data (2015).US EPA. Clean Watersheds Needs Survey (CWNS) – 2022 Report and Data (2024).El Abbadi, S. H. et al. Benchmarking greenhouse gas emissions from U.S. wastewater treatment for targeted reduction. Nat. Water (In press) (2025).Q. S. D. Group. werf: Benchmark Models for Water Resource Recovery Facilities (WRRF) in the US. EXPOsan https://github.com/QSD-Group/EXPOsan/tree/werf/exposan/werf (2025).QSD Group. EXPOsan—EXPOsition of sanitation and resource recovery systems.Metcalf & Eddy, Inc. Wastewater Engineering: Treatment and Resource Recovery (McGraw-Hill Higher Education, 2013).US EPA. Analysis of Operations & Maintenance Costs for Municipal Wastewater Treatment Systems. (1978).Otterpohl, R. & Freund, M. Dynamic models for clarifiers of activated sludge plants with dry and wet weather flows. Water Sci. Technol. 26, 1391–1400 (1992).
    Google Scholar 
    Takács, I., Patry, G. G. & Nolasco, D. A dynamic model of the clarification-thickening process. Water Res. 25, 1263–1271 (1991).
    Google Scholar 
    Solon, K. et al. Plant-wide modelling of phosphorus transformations in wastewater treatment systems: Impacts of control and operational strategies. Water Res. 113, 97–110 (2017).
    Google Scholar 
    Henze, M. et al. Activated sludge model No.2d, ASM2. D. Water Sci. Technol. 39, 165–182 (1999).
    Google Scholar 
    Rosen, C., Vrecko, D., Gernaey, K. V., Pons, M. N. & Jeppsson, U. Implementing ADM1 for plant-wide benchmark simulations in Matlab/Simulink. Water Sci. Technol. 54, 11–19 (2006).
    Google Scholar 
    Flores-Alsina, X. et al. Modelling phosphorus (P), sulfur (S) and iron (Fe) interactions for dynamic simulations of anaerobic digestion processes. Water Res. 95, 370–382 (2016).
    Google Scholar 
    Kazadi Mbamba, C., Batstone, D. J., Flores-Alsina, X. & Tait, S. A generalised chemical precipitation modelling approach in wastewater treatment applied to calcite. Water Res. 68, 342–353 (2015).
    Google Scholar 
    Kazadi Mbamba, C., Tait, S., Flores-Alsina, X. & Batstone, D. J. A systematic study of multiple minerals precipitation modelling in wastewater treatment. Water Res 85, 359–370 (2015).
    Google Scholar 
    Kazadi Mbamba, C. et al. Precipitation and dissolution. in Generalised Physicochemical Model (PCM) for Wastewater Processes (IWA Publishing, 2022).Musvoto, E. Integrated chemical–physical processes modelling—I. Development of a kinetic-based model for mixed weak acid/base systems. Water Res. 34, 1857–1867 (2000).
    Google Scholar 
    Musvoto, E. V., Wentzel, M. C. & Ekama, G. A. Integrated chemical–physical processes modelling—II. simulating aeration treatment of anaerobic digester supernatants. Water Res. 34, 1868–1880 (2000).
    Google Scholar 
    Cortes-Peña, Y., Kumar, D., Singh, V. & Guest, J. S. BioSTEAM: a fast and flexible platform for the design, simulation, and techno-economic analysis of biorefineries under uncertainty. ACS Sustain. Chem. Eng. 8, 3302–3310 (2020).
    Google Scholar 
    BioSTEAM Development Group. BioSTEAM: the biorefinery simulation and techno-economic analysis modules. https://github.com/BioSTEAMDevelopmentGroup/biosteam (2023).Seider, W. D. et al. Cost accounting and capital cost estimation. In Product and Process Design Principles 426–485 (Wiley, 2017).Williams, M. J., Bizier, P., Groman, J. & Tippetts, R. Alkaline stabilization. in Design of Water Resource Recovery Facilities (MOP8) (ed. The Water Environment Federation (WEF)) (McGraw-Hill Education, 2018).Beecher, N. et al. National U. S. Biosolids Data – FINAL REPORT: Advancing Understanding of Biosolids Management 30 Years after 40 CFR Part 503 and in the Early Stages of PFAS Concerns. https://www.biosolidsdata.org (2023).Rowles, L. S. et al. Financial viability and environmental sustainability of fecal sludge treatment with pyrolysis omni processors. ACS Environ. Au 2, 455–466 (2022).
    Google Scholar 
    Watabe, S. et al. Advancing the economic and environmental sustainability of the newgenerator nonsewered sanitation system. ACS Environ. Au 3, 209–222 (2023).
    Google Scholar 
    Feng, J., Li, Y., Strathmann, T. J. & Guest, J. S. Characterizing the opportunity space for sustainable hydrothermal valorization of wet organic wastes. Environ. Sci. Technol. 58, 2528–2541 (2024).
    Google Scholar 
    Zhang, X., Arnold, W. A., Wright, N., Novak, P. J. & Guest, J. S. Prioritization of early-stage research and development of a hydrogel-encapsulated anaerobic technology for distributed treatment of high strength organic wastewater. Environ. Sci. Technol. 58, 19651–19665 (2024).
    Google Scholar 
    Download referencesAcknowledgementsWe thank Prof. Peter A. Vanrolleghem (Université Laval) for providing valuable insights and advice on wastewater process modeling and plant-wide simulation platform development. We also thank Jianan Feng (University of Illinois Urbana-Champaign) for sharing compiled data on wastewater treatment process characterization for U.S. facilities. This study was funded by the U.S. Department of Energy Industrial Technologies Office. The views expressed in the article do not necessarily represent the views of DOE or the U.S. Government. The publisher, by accepting the article to publication, acknowledges that U.S. Government retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this work, or allow others to do so, for U.S. Government purposes.Author informationAuthors and AffiliationsThe Grainger College of Engineering, Department of Civil and Environmental Engineering, University of Illinois Urbana-Champaign, Urbana, IL, USAXinyi Zhang, Saumitra Rai, Zixuan Wang & Jeremy S. GuestDepartment of Civil and Environmental Engineering, Rutgers, The State University of New Jersey, Piscataway, NJ, USAYalin LiInstitute for Sustainability, Energy, and Environment, University of Illinois Urbana-Champaign, Urbana, IL, USAJeremy S. GuestAuthorsXinyi ZhangView author publicationsSearch author on:PubMed Google ScholarSaumitra RaiView author publicationsSearch author on:PubMed Google ScholarZixuan WangView author publicationsSearch author on:PubMed Google ScholarYalin LiView author publicationsSearch author on:PubMed Google ScholarJeremy S. GuestView author publicationsSearch author on:PubMed Google ScholarContributionsConceptualization: X.Z., Y.L. and J.S.G.; Funding acquisition: J.S.G.; Methodology: X.Z.; Software: X.Z., Y.L. and S.R.; Validation: Z.W.; Visualization: X.Z. and Y.L.; Manuscript writing: X.Z., S.R. and Z.W. in collaboration with all authors.Corresponding authorsCorrespondence to
    Xinyi Zhang, Yalin Li or Jeremy S. Guest.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary informationZhang_et_al_npg-CW_Supporting_Information_SubmittedZhang_et_al_npg-CW_Supporting_Information_Table_S11.Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleZhang, X., Rai, S., Wang, Z. et al. An agile benchmarking framework for wastewater resource recovery technologies.
    npj Clean Water (2025). https://doi.org/10.1038/s41545-025-00537-4Download citationReceived: 16 October 2025Accepted: 18 November 2025Published: 06 December 2025DOI: https://doi.org/10.1038/s41545-025-00537-4Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Focused groundwater recharge is controlled by landscape and climate

    AbstractFocused groundwater recharge, the concentrated infiltration of water through surface features including streams, depressions, or fractures to the water table, is accepted as the dominant recharge mechanism in arid climates. As climates become increasingly arid, groundwater recharge is expected to shift towards focused mechanisms. Yet the magnitude of focused recharge, its spatial distribution and controls across climate zones remain poorly characterised at the continental scale. Here, we compare historical rainfall tritium with >1700 groundwater tritium measurements to assess the likelihood of focused recharge across the Australian continent, providing important context for water resources management, with global implications. 46% of bores assessed show evidence of focused recharge, suggesting that conventional recharge estimates based on diffuse mechanisms may substantially underestimate total recharge. We show that fractured rock and perennial watercourses are the main landscape features that strongly influence the likelihood of focused recharge. While focused recharge is most common in arid regions, it also occurs in wetter climates where fractured rock enhances subsurface connectivity. As aridity and climate variability intensify, understanding the landscape-climate interactions that enable focused recharge, and how shifts in energy and water availability alter the role of groundwater in the water cycle, will be critical to sustaining groundwater resources.

    Similar content being viewed by others

    Intensifying tropical cyclones in the Arabian Sea replenish depleting aquifers

    Article
    Open access
    08 July 2025

    Groundwater recharge is sensitive to changing long-term aridity

    Article

    12 March 2024

    Decline in Iran’s groundwater recharge

    Article
    Open access
    21 October 2023

    Data availability

    The output data produced in this study are available as supporting information at the following Hydroshare link: http://www.hydroshare.org/resource/a9da2e2a766f403793bca6dc379715af93. Data used to support the findings in this study were obtained from different sources. Groundwater tritium data sources are listed in Table S1. Rainfall tritium data was provided by ANSTO52. The surface geology of Australia shapefile can be accessed at: https://ecat.ga.gov.au/geonetwork/srv/api/records/c8856c41-0d5b-2b1d-e044-00144fdd4fa685. The hydrogeology map of Australia shapefile can be accessed at: https://ecat.ga.gov.au/geonetwork/srv/api/records/2da7c234-63e9-10b2-e053-12a3070a174b86. The national surface hydrology lines dataset can be accessed at: https://ecat.ga.gov.au/geonetwork/srv/eng/catalog.search#/metadata/8313087. The gridded predicted rainfall tritium map can be accessed at: https://isotopehydrologynetwork.iaea.org/57. The NGIS bore logs from the Australian Groundwater Explorer provided by the Bureau of Meteorology are available at: http://www.bom.gov.au/water/groundwater/explorer/83. Some data presented in this paper has been visualised using scientific colour maps created by Crameri94.
    Code availability

    The Python script used for data analysis is available at the following Hydroshare link:
    http://www.hydroshare.org/resource/a9da2e2a766f403793bca6dc379715af93.
    ReferencesGreen, T. R. et al. Beneath the surface of global change: Impacts of climate change on groundwater. J. Hydrol. 405, 532–560 (2011).
    Google Scholar 
    Wada, Y., Van Beek, L. P. H. & Bierkens, M. F. P. Modelling global water stress of the recent past: on the relative importance of trends in water demand and climate variability. Hydrol. Earth Syst. Sci. 15, 3785–3808 (2011).
    Google Scholar 
    Morin-Crini, N. et al. Worldwide cases of water pollution by emerging contaminants: a review. Environ. Chem. Lett. 20, 2311–2338 (2022).
    Google Scholar 
    Schwarzenbach, R. P., Egli, T., Hofstetter, T. B., Von Gunten, U. & Wehrli, B. Global Water Pollution and Human Health. Annu. Rev. Environ. Resour. 35, 109–136 (2010).
    Google Scholar 
    Gleeson, T., Wada, Y., Bierkens, M. F. P. & van Beek, L. P. H. Water balance of global aquifers revealed by groundwater footprint. Nature 488, 197–200 (2012).
    Google Scholar 
    Wada, Y. et al. Global depletion of groundwater resources. Geophys. Res. Lett. 37, n/a-n/a (2010).
    Google Scholar 
    Werner, A. D. et al. An Initial Inventory and Indexation of Groundwater Mega-Depletion Cases. Water Resour. Manag. 27, 507–533 (2013).
    Google Scholar 
    Feng, S. & Fu, Q. Expansion of global drylands under a warming climate. Atmos. Chem. Phys. 13, 10081–10094 (2013).
    Google Scholar 
    Huang, J., Yu, H., Guan, X., Wang, G. & Guo, R. Accelerated dryland expansion under climate change. Nat. Clim. Change 6, 166–171 (2016).
    Google Scholar 
    Koutroulis, A. G. Dryland changes under different levels of global warming. Sci. Total Environ. 655, 482–511 (2019).
    Google Scholar 
    Loaiciga, H. A. & Doh, R. Groundwater for People and the Environment: A Globally Threatened Resource. Groundwater 62, 332–340 (2024).
    Google Scholar 
    Acworth, R. I., Rau, G. C., Cuthbert, M. O., Leggett, K. & Andersen, M. S. Runoff and focused groundwater-recharge response to flooding rains in the arid zone of Australia. Hydrogeol. J. 29, 737–764 (2021).
    Google Scholar 
    de Vries, J. J. & Simmers, I. Groundwater recharge: an overview of processes and challenges. Hydrogeol. J. 10, 5–17 (2002).
    Google Scholar 
    Lerner, D. N., Issar, A. S. & Simmers, I. Groundwater recharge: A Guide to Understanding and Estimating Natural Recharge, International Contributions to Hydrogeology. International Association of Hydrogeologists, Goring, UK 8, (1990).Lee, S., Irvine, D. J., Duvert, C., Rau, G. C. & Cartwright, I. A high-resolution map of diffuse groundwater recharge rates for Australia. Hydrol. Earth Syst. Sci. 28, 1771–1790 (2024).
    Google Scholar 
    Scanlon, B. R. et al. Global synthesis of groundwater recharge in semiarid and arid regions. Hydrol. Process. 20, 3335–3370 (2006).
    Google Scholar 
    Scanlon, B. R., Langford, R. P. & Goldsmith, R. S. Relationship between geomorphic settings and unsaturated flow in an arid setting. Water Resour. Res. 35, 983–999 (1999).
    Google Scholar 
    Healy, R. W. Estimating Groundwater Recharge. (Cambridge University Press, United Kingdom, 2010).Meixner, T. et al. Implications of projected climate change for groundwater recharge in the western United States. J. Hydrol. 534, 124–138 (2016).
    Google Scholar 
    Hartmann, A., Goldscheider, N., Wagener, T., Lange, J. & Weiler, M. Karst water resources in a changing world: Review of hydrological modeling approaches. Rev. Geophysics 52, 218–242 (2014).
    Google Scholar 
    Armengol, S. et al. Isogeochemical characterization of mountain system recharge processes in the Sierra Nevada, California. Water Resour. Res. 60, e2023WR035719 (2024).
    Google Scholar 
    Gee, G. W. & Hillel, D. Groundwater recharge in arid regions: Review and critique of estimation methods. Hydrological Process. 2, 255–266 (1988).
    Google Scholar 
    Allison, G. B., Gee, G. W. & Tyler, S. W. Vadose-Zone Techniques for Estimating Groundwater Recharge in Arid and Semiarid Regions. Soil Sci. Soc. Am. J. 58, 6–14 (1994).
    Google Scholar 
    Li, M., Xie, Y., Dong, Y., Wang, L. & Zhang, Z. Review: Recent progress on groundwater recharge research in arid and semiarid areas of China. Hydrogeol. J. 32, 9–30 (2024).
    Google Scholar 
    Owuor, S. O. et al. Groundwater recharge rates and surface runoff response to land use and land cover changes in semi-arid environments. Ecol. Process 5, 16 (2016).
    Google Scholar 
    Xu, Y. & Beekman, H. E. Review: Groundwater recharge estimation in arid and semi-arid southern Africa. Hydrogeol. J. 27, 929–943 (2019).
    Google Scholar 
    Berghuijs, W. R., Luijendijk, E., Moeck, C., van der Velde, Y. & Allen, S. T. Global recharge data set indicates strengthened groundwater connection to surface fluxes. Geophys. Res. Lett. 49, e2022GL099010 (2022).West, C. et al. Ground truthing global-scale model estimates of groundwater recharge across Africa. Sci. Total Environ. 858, 159765 (2023).
    Google Scholar 
    Reinecke, R. et al. Uncertainty in model estimates of global groundwater depth. Environ. Res. Lett. 19, 114066 (2024).
    Google Scholar 
    Cuthbert, M. O. et al. Observed controls on resilience of groundwater to climate variability in sub-Saharan Africa. Nature 572, 230–234 (2019).
    Google Scholar 
    Baker, A. et al. An underground drip water monitoring network to characterize rainfall recharge of groundwater at different geologies, environments, and climates across Australia. Geosci. Instrum. Method. Data Syst. 13, 117–129 (2024).
    Google Scholar 
    Barua, S., Cartwright, I., Dresel, P. E. & Daly, E. Using multiple methods to investigate the effects of land-use changes on groundwater recharge in a semi-arid area. Hydrol. Earth Syst. Sci. 25, 89–104 (2021).
    Google Scholar 
    Boas, T. & Mallants, D. Episodic extreme rainfall events drive groundwater recharge in arid zone environments of central Australia. J. Hydrol.: Regional Stud. 40, 101005 (2022).
    Google Scholar 
    Cartwright, I., Werner, A. & Woods, J. Using geochemistry to discern the patterns and timescales of groundwater recharge and mixing on floodplains in semi-arid regions. J. Hydrol. 570, 612–622 (2019).
    Google Scholar 
    Crosbie, R. S., Holland, K. L. & McVicar, T. R. Regional-scale partitioning of transmission losses and groundwater recharge using satellite estimates of actual evapotranspiration in an arid environment. Ecohydrology 16, e2490 (2023).
    Google Scholar 
    Hocking, M. & Kelly, B. F. J. Groundwater recharge and time lag measurement through Vertosols using impulse response functions. J. Hydrol. 535, 22–35 (2016).
    Google Scholar 
    Priestley, S. C. et al. A 35 ka record of groundwater recharge in south-west Australia using stable water isotopes. Sci. Total Environ. 717, 135105 (2020).
    Google Scholar 
    Skrzypek, G., Siller, A., McCallum, J. L. & Dogramaci, S. Groundwater recharge through internally drained basins in a semiarid climate, Western Australia. J. Hydrol.: Regional Stud. 47, 101388 (2023).
    Google Scholar 
    Wood, C., Cook, P. G., Harrington, G. A. & Knapton, A. Constraining spatial variability in recharge and discharge in an arid environment through modeling carbon-14 with improved boundary conditions. Water Resour. Res. 53, 142–157 (2017).
    Google Scholar 
    Xie, Y., Crosbie, R., Yang, J., Wu, J. & Wang, W. Usefulness of Soil Moisture and Actual Evapotranspiration Data for Constraining Potential Groundwater Recharge in Semiarid Regions. Water Resour. Res. 54, 4929–4945 (2018).
    Google Scholar 
    Baker, A. et al. Rainfall recharge thresholds in a subtropical climate determined using a regional cave drip water monitoring network. J. Hydrol. 587, 125001 (2020).
    Google Scholar 
    Chmielarski, M. et al. Identifying the influence of episodic events on groundwater recharge in semi-arid environments using environmental tracers. J. Hydrol. 633, 130848 (2024).
    Google Scholar 
    Cuthbert, M. O. et al. Understanding and quantifying focused, indirect groundwater recharge from ephemeral streams using water table fluctuations. Water Resour. Res. 52, 827–840 (2016).
    Google Scholar 
    Robson, T. C. & Webb, J. A. The use of environmental tracers to determine focused recharge from a saline disposal basin and irrigation channels in a semiarid environment in Southeastern Australia. J. Hydrol. 538, 326–338 (2016).
    Google Scholar 
    Allison, G. & Hughes, M. The use of environmental chloride and tritium to estimate total recharge to an unconfined aquifer. Soil Res. 16, 181 (1978).
    Google Scholar 
    Allison, G. B. & Hughes, M. W. Comparison of recharge to groundwater under pasture and forest using environmental tritium. J. Hydrol. 17, 81–95 (1972).
    Google Scholar 
    Cartwright, I. & Morgenstern, U. Constraining groundwater recharge and the rate of geochemical processes using tritium and major ion geochemistry: Ovens catchment, southeast Australia. J. Hydrol. 475, 137–149 (2012).
    Google Scholar 
    Cook, P. G. & Robinson, N. I. Estimating groundwater recharge in fractured rock from environmental 3H and 36Cl, Clare Valley, South Australia. Water Resour. Res. 38, 111–1113 (2002).
    Google Scholar 
    Cartwright, I., Cendón, D., Currell, M. & Meredith, K. A review of radioactive isotopes and other residence time tracers in understanding groundwater recharge: Possibilities, challenges, and limitations. J. Hydrol. 555, 797–811 (2017).
    Google Scholar 
    Li, Z., Chen, X., Liu, W. & Si, B. Determination of groundwater recharge mechanism in the deep loessial unsaturated zone by environmental tracers. Sci. Total Environ. 586, 827–835 (2017).
    Google Scholar 
    Telloli, C. et al. Characterization of groundwater recharge through tritium measurements. Adv. Geosci. 57, 21–36 (2022).
    Google Scholar 
    Tadros, C. V., Hughes, C. E., Crawford, J., Hollins, S. E. & Chisari, R. Tritium in Australian precipitation: A 50 year record. J. Hydrol. 513, 262–273 (2014).
    Google Scholar 
    Jasechko, S. et al. Global aquifers dominated by fossil groundwaters but wells vulnerable to modern contamination. Nat. Geosci. 10, 425–429 (2017).
    Google Scholar 
    Cabrera, A., Cendón, D. I., Aparicio, V. & Currell, M. J. Intensive agriculture, a pesticide pathway to >100 m deep groundwater below dryland agriculture, Cordoba Pampas, Argentina. J. Hydrol. 643, 131989 (2024).
    Google Scholar 
    Taufiq, A. et al. Impact of excessive groundwater pumping on rejuvenation processes in the Bandung basin (Indonesia) as determined by hydrogeochemistry and modeling. Hydrogeol. J. 26, 1263–1279 (2018).
    Google Scholar 
    Thaw, M., GebreEgziabher, M., Villafañe-Pagán, J. Y. & Jasechko, S. Modern groundwater reaches deeper depths in heavily pumped aquifer systems. Nat. Commun. 13, 5263 (2022).Terzer-Wassmuth, S., Araguás-Araguás, L. J., Copia, L. & Wassenaar, L. I. High spatial resolution prediction of tritium (3H) in contemporary global precipitation. Sci. Rep. 12, 10271 (2022).
    Google Scholar 
    Pinti, D. L. et al. Origin of volatile species and aqueous fluids in the Los Humeros Geothermal Field, Mexico. Chem. Geol. 584, 120539 (2021).
    Google Scholar 
    Poulsen, D. L., Cook, P. G., Simmons, C. T., Solomon, D. K. & Dogramaci, S. Depth-Resolved Groundwater Chemistry by Longitudinal Sampling of Ambient and Pumped Flows Within Long-Screened and Open Borehole Wells. Water Resour. Res. 55, 9417–9435 (2019).
    Google Scholar 
    Reinecke, R. et al. Uncertainty of simulated groundwater recharge at different global warming levels: a global-scale multi-model ensemble study. Hydrol. Earth Syst. Sci. 25, 787–810 (2021).
    Google Scholar 
    Döll, P. & Fiedler, K. Global-scale modeling of groundwater recharge. Hydrol. Earth Syst. Sc. 12, 863–885 (2008).
    Google Scholar 
    Graaf, de et al. A high-resolution global-scale groundwater model. Hydrol. Earth Syst. Sc. 19, 823–837 (2015).
    Google Scholar 
    Müller Schmied, H. et al. The global water resources and use model WaterGAP v2. 2d: Model description and evaluation. Geosci. Model Dev. 14, 1037–1079 (2021).
    Google Scholar 
    Wan, W., Döll, P. & Müller Schmied, H. Global-Scale Groundwater Recharge Modeling Is Improved by Tuning Against Ground-Based Estimates for Karst and Non-Karst Areas. Water Resour. Res. 60, e2023WR036182 (2024).
    Google Scholar 
    Cendón, D. I. et al. Freshwater recharge into a shallow saline groundwater system, Cooper Creek floodplain, Queensland, Australia. J. Hydrol. 392, 150–163 (2010).
    Google Scholar 
    Jasechko, S. et al. The pronounced seasonality of global groundwater recharge. Water Resour. Res. 50, 8845–8867 (2014).
    Google Scholar 
    Guinea, A., Hollins, S., Meredith, K., Hankin, S. & Cendón, D. I. Characterization of the subsurface architecture and identification of potential groundwater paths in a clay-rich floodplain using multi-electrode resistivity imaging. Hydrological Sci. J. 63, 909–925 (2018).
    Google Scholar 
    Lapworth, D. J. et al. Deep urban groundwater vulnerability in India revealed through the use of emerging organic contaminants and residence time tracers. Environ. Pollut. 240, 938–949 (2018).
    Google Scholar 
    Cartwright, I., Weaver, T., Cendón, D. & Swane, I. Environmental isotopes as indicators of inter-aquifer mixing, Wimmera region, Murray Basin, Southeast Australia. Chem. Geol. 277, 214–226 (2010).
    Google Scholar 
    Meredith, K. T. et al. Evaporation and concentration gradients created by episodic river recharge in a semi-arid zone aquifer: Insights from Cl−, δ18O, δ2H, and 3H. J. Hydrol. 529, 1070–1078 (2015).
    Google Scholar 
    Jasechko, S. et al. Terrestrial water fluxes dominated by transpiration. Nature 496, 347–350 (2013).
    Google Scholar 
    Gleeson, T., Novakowski, K. & Kurt Kyser, T. Extremely rapid and localized recharge to a fractured rock aquifer. J. Hydrol. 376, 496–509 (2009).
    Google Scholar 
    Lall, U., Josset, L. & Russo, T. A Snapshot of the World’s Groundwater Challenges. Annu. Rev. Environ. Resour. 45, 171–194 (2020).
    Google Scholar 
    Gorelick, S. M. & Zheng, C. Global change and the groundwater management challenge. Water Resour. Res. 51, 3031–3051 (2015).
    Google Scholar 
    Bierkens, M. F. P. & Wada, Y. Non-renewable groundwater use and groundwater depletion: a review. Environ. Res. Lett. 14, 063002 (2019).
    Google Scholar 
    Gleeson, T., Cuthbert, M., Ferguson, G. & Perrone, D. Global Groundwater Sustainability, Resources, and Systems in the Anthropocene. Annu. Rev. Earth Planet. Sci. 48, 431–463 (2020).
    Google Scholar 
    Wada, Y. & Heinrich, L. Assessment of transboundary aquifers of the world—vulnerability arising from human water use. Environ. Res. Lett. 8, 024003 (2013).
    Google Scholar 
    Clark, I. D. & Fritz, P. Environmental Isotopes in Hydrogeology. (CRC Press, 1997).Duvert, C., Stewart, M. K., Cendón, D. I. & Raiber, M. Time series of tritium, stable isotopes and chloride reveal short-term variations in groundwater contribution to a stream. Hydrol. Earth Syst. Sci. 20, 257–277 (2016).
    Google Scholar 
    Morgenstern, U. & Daughney, C. J. Groundwater age for identification of baseline groundwater quality and impacts of land-use intensification – The National Groundwater Monitoring Programme of New Zealand. J. Hydrol. 456–457, 79–93 (2012).
    Google Scholar 
    Stewart, M. K., Morgenstern, U. & McDonnell, J. J. Truncation of stream residence time: how the use of stable isotopes has skewed our concept of streamwater age and origin. Hydrological Process. 24, 1646–1659 (2010).
    Google Scholar 
    Geoscience Australia. Grid Coordinate Application Batch Processing. Geodetic Calculators: Grid Coordinate Application Batch Processing https://geodesyapps.ga.gov.au/grid-coordinate-batch-processing (2021).Bureau of Meteorology. Australian Groundwater Explorer. http://www.bom.gov.au/water/groundwater/explorer/map.shtml (2025).Taylor, C. B. & Fletcher, E. H. Assessment of Data for Tritium in Southern Hemisphere Precipitation, 1960-1983. 20 (1988).Raymond, Liu, Gallagher, Zhang, & Highet. Surface Geology of Australia 1:1 million scale dataset 2012 edition. https://doi.org/10.26186/74619 (2012).BrodieR., Kilgour, B., Jacobson, G. & Lau, J. Hydrogeology Map of Australia (G.Jacobson and JE.Lau Hydrogeology Map). (2019).Crossman, S. & Li, O. Surface Hydrology Lines (National). (2015).Bureau of Meteorology. Australian Water Outlook. https://awo.bom.gov.au/ (2022).Bureau of Meteorology. Decadal and multi-decadal rainfall averages maps. http://www.bom.gov.au/climate/maps/averages/decadal-rainfall/ (2023).United Nations Environment Programme. World Atlas of Desertification: Second Edition. (1997).Cartwright, I. & Morgenstern, U. Transit times from rainfall to baseflow in headwater catchments estimated using tritium: The Ovens River, Australia. Hydrol. Earth Syst. Sci. 19, 3771–3785 (2015).
    Google Scholar 
    Duvert, C. et al. Seasonal shift from biogenic to geogenic fluvial carbon caused by changing water sources in the wet-dry tropics. JGR Biogeosci. 125, e2019JG005384 (2020).
    Google Scholar 
    Lee, S. Focused groundwater recharge driven by landscape controls and climate. HydroShare, http://www.hydroshare.org/resource/a9da2e2a766f403793bca6dc379715af (2025).Crameri, F. Scientific colour maps. https://doi.org/10.5281/zenodo.1243862 (2018).Download referencesAcknowledgementsWe would like to acknowledge the Traditional Owners of the lands related to any aspects of this study. We would like to acknowledge the institutions and individuals that collected the data used in this study originally. We are grateful to the three anonymous peer reviewers for their detailed reviews and valuable suggestions. Stephen Lee was supported by a Research Training Programme scholarship (doi.org/10.82133/C42F-K220) through Charles Darwin University and by the Australian Geoscience Council Lee Parkin Australian Geoscience Information Association Grant.Author informationAuthors and AffiliationsFaculty of Science and Technology and Research Institute for the Environment and Livelihoods, Charles Darwin University, Brinkin, NT, AustraliaStephen Lee, Dylan J. Irvine & Clément DuvertNational Centre for Groundwater Research and Training, Bedford Park, SA, AustraliaStephen Lee, Dylan J. Irvine, Gabriel C. Rau, Matthew Currell & Clément DuvertSchool of Environmental and Life Sciences, The University of Newcastle, Callaghan, NSW, AustraliaGabriel C. RauCentre for Integrated Resilience: Coasts, Water and Climate, The University of Newcastle, Callaghan, NSW, AustraliaGabriel C. RauSchool of Engineering and Built Environment, and Australian Rivers Institute, Griffith University, Nathan, QLD, AustraliaMatthew CurrellAustralian Nuclear Science and Technology Organisation (ANSTO), Lucas Heights, NSW, AustraliaCarol V. TadrosSchool of Biological, Earth and Environmental Sciences, UNSW Sydney, Sydney, NSW, AustraliaCarol V. TadrosCollege of Science and Engineering, James Cook University, Cairns, QLD, AustraliaClément DuvertAuthorsStephen LeeView author publicationsSearch author on:PubMed Google ScholarDylan J. IrvineView author publicationsSearch author on:PubMed Google ScholarGabriel C. RauView author publicationsSearch author on:PubMed Google ScholarMatthew CurrellView author publicationsSearch author on:PubMed Google ScholarCarol V. TadrosView author publicationsSearch author on:PubMed Google ScholarClément DuvertView author publicationsSearch author on:PubMed Google ScholarContributionsS.L. contributed to research design, collated data, conducted all data analyses and wrote the original manuscript. D.I. conceptualised the idea for the study, providing primary supervision and supported the writing and review of the manuscript. G.R. supported the writing and review of the manuscript. M.C. supported the writing and review of the manuscript. C.T. supported the writing and review of the manuscript. C.D. conceptualised the idea for the study, providing secondary supervision, and supported the writing and review of the manuscript.Corresponding authorCorrespondence to
    Stephen Lee.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Communications Earth & Environment thanks Nadim K. Copty and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editor: Nicola Colombo. A peer review file is available.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary informationTransparent Peer Review fileSupplementary information of Focused groundwater recharge is controlled by landscape and climateRights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleLee, S., Irvine, D.J., Rau, G.C. et al. Focused groundwater recharge is controlled by landscape and climate.
    Commun Earth Environ (2025). https://doi.org/10.1038/s43247-025-03063-wDownload citationReceived: 06 August 2025Accepted: 25 November 2025Published: 06 December 2025DOI: https://doi.org/10.1038/s43247-025-03063-wShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Enhancing seed, land and water management in humanitarian interventions for urban areas under siege

    In areas under siege, the growing of fruits and vegetables and the keeping of livestock have always provided a lifeline for desperate urban populations. Lessons from siege warfare in modern times should be applied to the development of innovative humanitarian interventions aimed at facilitating urban agriculture and food security programmes during future sieges.

    Access through your institution

    Buy or subscribe

    This is a preview of subscription content, access via your institution

    Access options

    Access through your institution

    Subscribe to this journal

    Receive 12 digital issues and online access to articles

    $119.00 per year
    only $9.92 per issue

    Learn more

    Buy this articlePurchase on SpringerLinkInstant access to the full article PDF.USD 39.95Prices may be subject to local taxes which are calculated during checkout

    Additional access options:

    Log in

    Learn about institutional subscriptions

    Read our FAQs

    Contact customer support

    Fig. 1: Urban timeline of sieges that have affected urban food systems.

    ReferencesParkin, S. The Forbidden Garden of Leningrad: A True Story of Science and Sacrifice in a City Under Siege (Sceptre, 2024).Abadžić, A. Sarajevo: The Longest Siege 39–40 (Modul Memorije, 2022).Download referencesAuthor informationAuthors and AffiliationsIndependent consultant and researcher, Runnymede, Surrey, UKAndrew Adam-BradfordAuthorsAndrew Adam-BradfordView author publicationsSearch author on:PubMed Google ScholarCorresponding authorCorrespondence to
    Andrew Adam-Bradford.Ethics declarations

    Competing interests
    The author declares no competing interests.

    Rights and permissionsReprints and permissionsAbout this articleCite this articleAdam-Bradford, A. Enhancing seed, land and water management in humanitarian interventions for urban areas under siege.
    Nat Water (2025). https://doi.org/10.1038/s44221-025-00556-yDownload citationPublished: 05 December 2025Version of record: 05 December 2025DOI: https://doi.org/10.1038/s44221-025-00556-yShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    High-frequency monitoring reveals a CO2 source-sink shift in a subtropical eutrophic urban lake

    AbstractEutrophic urban lakes with CO2-supersaturation represent potential carbon (C) sources; however, the drivers behind the reported C-source–sink shift remain poorly understood. This study provides a systematic assessment of daytime/seasonal pCO2 and fCO2 dynamics in a subtropical moderately eutrophic urban lake (Bailuwan, China), based on over a year of high-frequency monitoring, aiming to clarify the mechanisms regulating CO2 exchange at the water–air interface in such ecosystems. Our work revealed consistent daytime declines in pCO2 (and fCO2) on 12 sampling days, though morning–afternoon differences were not significant (n = 24). Novel episodic undersaturation events were newly observed in October 2020 and March 2021, contrasting with the prevailing supersaturation. Annual mean values (n = 48) reached 1789 µatm (pCO2) and 130 mmol m−2 h−1 (fCO2). Critically, we identified a pronounced semi-annual divergence: pCO2 from January to June significantly exceeded values from July to November. Both periods maintained a net source status (> 420 µatm), lacking the typical spring-sink/summer-source transition reported in previous studies. Key regulators, such as pH, chlorophyll a, and dissolved oxygen, influence C-sink-source dynamics, with eutrophication further modulating these shifts. These original findings highlight the need for targeted strategies to reduce pollutants and enhance carbon sequestration in urban lakes.

    Similar content being viewed by others

    Seasonal and diurnal changes of pCO2 in the lower Brahmaputra River, Bangladesh

    Article
    Open access
    21 November 2024

    Emerging climate impact on carbon sinks in a consolidated carbon budget

    Article

    12 November 2025

    Freeze-thaw seasonal variations and environmental controls of CO2 and CH4 diffusive emissions from reservoirs in the upper Yellow River

    Article
    Open access
    03 October 2025

    IntroductionFreshwater ecosystems, encompassing lakes, rivers, and wetlands, are pivotal in the biogeochemical dynamics of greenhouse gases (GHGs), with impacts extending from local to planetary scales1,2,3,4,5,6. In particular, inland lakes, despite their limited spatial coverage (~ 3.7% of non-glacial terrestrial surfaces)7, are key players in carbon cycling through sequestration, transport, and transformation mechanisms8,9,10. Empirical studies have consistently demonstrated that approximately 90% of global inland lakes exhibit carbon dioxide (CO2) supersaturation relative to atmospheric equilibrium11,12. Global estimates suggest that CO2 emissions from lakes range from 0.07 to 0.15 Pg C year-1, with upper estimates reaching as high as 0.57 Pg C year-113,14,15. With a spatial extent representing 6.2% of global lake coverage (25°N–54°N), lakes in China release roughly 15.98 Tg C year− 1 in the form of atmospheric CO2 emissions16,17,18. Unexpectedly, shallow lakes with extensive surface areas but depths of less than 3 m have been identified as critical hot-spots for CO2 emissions owing to their high biogeochemical activity and efficient gas exchange19,20. The consistent pattern of CO2 supersaturation in these aquatic environments, with CO2 levels exceeding the atmospheric equilibrium range of 380−420 µatm, provides clear evidence of their role as net CO2 sources to the atmosphere21,22,23,24.From a global spatiotemporal perspective, the CO2 emitted into the atmosphere is rebalanced through biogeochemical mechanisms such as photosynthetic assimilation, sediment burial, and oceanic uptake, collectively contributing to the dynamic equilibrium of the global carbon cycle25,26. From 2007 onward, the oceans have sequestered about 56% of human-induced carbon emissions, resulting in detectable ocean acidification (pH), with the residual 44% accumulating in the atmosphere3,5,27,28. Growing empirical evidence documents the conversion of particular lake ecosystems from net carbon sources to net sinks, with some currently exhibiting transitional carbon dynamics29,30,31. Systematic monitoring data presented by Xiao et al.32 showed a significant downward trend in CO2 outgassing rates from Chinese lakes between the 1980s and 2000s, pointing to their gradual transformation from C-releasing to C-sequestering ecosystems. This transformation may be attributed to synergistic interactions among biogeochemical processes (e.g., enhanced C-sequestration and organic matter burial), shifting environmental conditions (e.g., nutrient loading and hydrological regimes), and targeted anthropogenic interventions (e.g., ecological restoration and eutrophication control)25,33,34,35,36. Consequently, systematic documentation of C-source-sink dynamics and their regulatory controls in lacustrine ecosystems is imperative to refine predictive models of C-exchange and quantify their contributions to regional C-budgets under evolving climatic and anthropogenic pressures.Watershed urbanization generates synergistic perturbations to aquatic ecosystem processes, wherein combined effects of anthropogenic nutrient loading and riparian habitat fragmentation lead to fundamental alterations in carbon transformation pathways and greenhouse gas exchange in urban water systems37,38. Urban lakes, as highly sensitive freshwater ecosystems, are particularly susceptible to algal blooms, resulting in eutrophication35,39,40,41. In this study, an urban lake is defined as a lentic water body situated entirely within a metropolitan area, whose hydrological processes, water quality, and ecological functions are primarily governed by anthropogenic activities15,41,56,89. Key characteristics include: (i) heavily modified hydrology through water level control and engineered shorelines; (ii) significant nutrient inputs from urban runoff and wastewater; (iii) altered ecological communities due to habitat modification and recreational use; and (iv) serving dual roles in both receiving urban discharges and providing ecosystem services such as flood mitigation and recreation25,26,42. Comparative studies reveal dramatic differences in carbon emissions between two Indian lakes: the hypereutrophic Belandur Lake exhibits exceptionally high CO2 efflux rates (5711 ± 844 Tg C year− 1), while Jakkur Lake, currently under ecological rehabilitation, demonstrates substantially lower emissions (24 ± 10 Tg C year− 1)42. While empirical evidence suggests that eutrophication could exert a modest positive influence on CO2 sequestration in non-urban lakes29,36, its role and mechanistic pathways in urban lakes remain inadequately elucidated. Consequently, to advance our comprehension of the underlying mechanisms regulating CO2 exchange in urban lakes, particularly shallow lakes, it is imperative to conduct more extensive field measurements and systematic analyses.Inland water CO2 fluxes (fCO2) are predominantly controlled by the interplay between aqueous CO2 partial pressure (pCO2) and the rate of gas transfer (kCO2) across the water-air boundary in freshwater ecosystems17,43,44. The pCO2 parameter, recognized as a pivotal determinant in deciphering carbon cycle variability across urban water bodies at multiple spatial scales11,45, demonstrates multifactorial regulation through: (i) environmental drivers (e.g., solar irradiance)4,9,10, (ii) biogeochemical processes (particularly aquatic metabolism involving photosynthesis-respiration/P-R coupling)46, (iii) hydrological dynamics (including thermal stratification and mixing regimes)47,48,49, and (iv) allochthonous carbon inputs from catchment areas50,51,52,53. Correspondingly, kCO2 variability exhibits primary dependence on wind shear stress and thermal conditions54. Distinct from the relative homogeneity of atmospheric pCO2, lacustrine pCO2 manifests pronounced spatiotemporal heterogeneity across daytime, monthly, and seasonal scales, with system-specific characteristics strongly influenced by morphometric parameters and trophic status4,5,6,55,56. Empirical evidence from Lake Ulansuhai, a shallow urban waterbody, reveals eutrophication-induced functional shifts from CO2 source to sink conditions57. Moreover, rapid urban expansion around lacustrine environments introduces substantial complexity to carbon cycling processes, with notable impacts on CO2 emission patterns5,58,59. Despite these insights, critical knowledge gaps persist regarding high temporal resolution characterization of pCO2 dynamics in urban shallow lakes, necessitating systematic investigations into urbanization-CO2 emission synergies through integrated observational and modeling approaches.Building upon the aforementioned backgrounds, this work aims to address a central scientific question: how and to what extent do eutrophic urban lakes modulate the dynamics of CO2 exchange across the water–air interface under high-frequency monitoring? To resolve the gap, we implemented a comprehensive study assessing daytime CO2 patterns in a shallow, subtropical urban lake with elevated nutrient levels in southwest China. This work pursues three principal aims: (i) performing monthly diurnal monitoring of pCO2/fCO2 and associated physicochemical variables between October 2020 and November 2021; (ii) identifying key hydrological and environmental controls on aquatic carbon cycling; and (iii) evaluating the premise that subtropical eutrophic lakes demonstrate alternating C-source-sink behavior under high-frequency investigation. The findings of this work are expected to provide novel insights into the mechanisms governing CO2 exchange in lacustrine systems and to refine the quantification of CO2 emissions from urban lakes, thereby reducing uncertainties in regional/global C-budget assessments.ResultsVariations in pCO2
    Throughout our investigation period (07:00–18:00 CST), sustained/significant (p > .05) daytime pCO2 decreases were observed during nine sampling days, with the exception of three dates (February 28, March 30, and November 22, 2021) that exhibited gradual increasing trends (n = 1; Fig. 1A–L). Considering individual field investigation days, the declining pattern remained predominant across most sampling occasions, except those illustrated in Fig. 1D, E, I, J, L. The composite analysis of four temporal sampling points revealed a statistically remarkedly (p < 0.05) reduction in mean pCO2 levels (n = 12) between 11:00 and 18:00 CST (Fig. 2M; Table 1). Relative to the 07:00 CST reference (2043.91 ± 2033.26 µatm), pCO2 exhibited pronounced (p < 0.05) daytime fluctuations as 42.63% at 11:00 CST, followed by successive reductions of 49.18% (14:00 CST) and 43.42% (18:00 CST). Interestingly, non-significant (p > 0.05) diurnal variation was detected between morning (07:00–11:00 CST; 2479.56 ± 3971.80 µatm) and afternoon periods (14:00–18:00 CST; 1097.66 ± 803.29 µatm) when analyzing the combined dataset (n = 24; Fig. 2A).Fig. 1Hourly variations of pCO2 (solid line) and fCO2 (dashed line) in the studied lake. Hollow circles denote individual measurements (n = 1; Fig. 2A–L) and mean values (n = 12; Fig. 2M–N). Field measurements originally scheduled for July 2021 were conducted on August 2 due to meteorological constraints.Full size imageFig. 2Diurnal comparison of mean pCO2 (A), fCO2 (B), Twater (C), pH (D), Chla (E) and DO (F) between morning (07:00–11:00 CST) and afternoon (14:00–18:00 CST) periods (n = 24). Distinct lowercase letters denote statistically significant differences (p < 0.05) between temporal periods. The vertical line on the bars represents the mean ± standard deviations (SDs).Full size imageTable 1 Temporal correlation analysis of daily pCO2 and fCO2 in Bailuwan lake (n = 4). x, time; y1, pCO2; y2, fCO2; r2, regression coefficients.Full size tableMonthly analysis revealed unexpected temporal patterns, with peak pCO2 concentrations (n = 4; 8148.02 ± 1882.12 µatm) occurring in October 2021, contrasting sharply with the minimum values recorded in October 2020 (198.14 ± 65.20 µatm). Throughout the annual cycle (January–November 2021), mean pCO2 levels during the first half of the year (January-June; n = 28; 1239.34 ± 781.00 µatm) generally lowered (p > 0.05) those of the latter half (July-November; n = 16; 3143.15 ± 4673.69 µatm). However, January and March exhibited the lower mean concentrations (510.20 ± 101.94 µatm and 314.92 ± 145.48 µatm, respectively; Fig. 3A). The overall daytime mean pCO2 across all measurements was 1788.61 ± 2919.44 µatm (n = 48).Fig. 3Monthly variations of pCO2 (A) and fCO2 (B) at the water–air interface in Bailuwan Lake. Upper panel displays median (black line) and mean (red line, n = 4), with whiskers spanning the range between [Q1 − 1.5×IQR] (lower bound) and [Q3 + 1.5×IQR] (upper bound). Q1: first quartile; Q3: third quartile; IQR: interquartile range (Q3 − Q1). Oct.0 and Dec.0 denote October and December 2020 sampling campaigns, respectively; remaining data were obtained in 2021. The vertical line on the bars represents the mean ± standard deviations (SDs).Full size imageAlterations in fCO2
    Similar to the pCO2 patterns, the daytime mean fCO2 concentrations (n = 12) exhibited a characteristic pattern of initial increase significantly (p < 0.05) followed by subsequent decrease. Specifically, compared to the baseline measurement at 07:00 CST (139.16 ± 175.86 mmol m− 2 h− 1), fCO2 concentrations showed a 76.13% increase by 11:00 CST, followed by significant reductions of 55.93% and 45.78% at 14:00 and 18:00 CST, respectively. Analysis of individual sampling day (n = 1) revealed consistent and statistically significant (p < 0.05) daytime fCO2 decreases across nine sampling dates, with the exception of February 28, March 30, and November 22, 2020 (Fig. 1A–L). Further, comparative analysis between morning (07:00–11:00 CST; 192.14 ± 418.09 mmol m− 2 h− 1) and afternoon periods (14:00–18:00 CST; 68.39 ± 76.65 mmol m− 2 h− 1) demonstrated non-significant (p > 0.05) diurnal variation in mean fCO2 concentrations (n = 24; Fig. 2B).Monthly analysis revealed distinct temporal patterns, with negative daytime mean fCO2 values (n = 4) recorded in October 2020 and March 2021, contrasting sharply with the peak concentration observed in October 2021 (n = 4; 775.16 ± 873.01 mmol m− 2 h− 1; Fig. 3B). Remarkably, the calculated mean fCO2 concentration was 130.26 ± 303.85 mmol m− 2 h− 1 (n = 48) across the entire study period.Classification of trophic stateThe Carlson’s trophic state index (TSI) was quantitatively assessed through the established methodology in Method S1, incorporating key limnological parameters including total phosphorus (TP), total nitrogen (TN), water transparency (TPC), chlorophyll a (Chla), and chemical oxygen demand (CODMn) as detailed in Table S1 and Fig. 4. The comprehensive TSI(∑) analysis yielded a mean value of 63.04 (Table 2), categorizing the lake within the moderately-eutrophic classification according to the standard limnological criteria. Temporal analysis of TSI(∑) dynamics revealed sustained values exceeding the eutrophication threshold of 60 across the majority of sampling intervals. However, notable deviations (p < 0.05) were observed during the spring sampling campaigns of March and April 2021, during which TSI(∑) values fell below this critical benchmark, as documented in Table S2.Fig. 4Temporal variations in water quality parameters of Bailuwan Lake: TPC (A,B), Chla (C,D), TN (E,F), TP (G,H), and CODMn (I,J). Additional details are provided in Fig. 3.Full size imageTable 2 Comprehensive TSI of Bailuwan lake (n = 48). The table presents NT, TP, CODMn, total dissolved nitrogen (TDN) and eutrophication evaluation criteria, as detailed in methods S2–S7.Full size tableEnvironmental parametersRegarding diurnal variations, water quality parameters including aqueous temperature (Twater), TPC, dissolved oxygen (DO), chloride ion (Cl⁻), nitrate ion (NO3⁻), and sulfate ion (SO42⁻) exhibited a progressive increase from 07:00 to 18:00 CST (Figs. 4, 5 and 6). Surprisingly, DO concentrations demonstrated a pronounced surge of 57.33% over this period, escalating from baseline levels at 07:00 CST to maximum values observed at 18:00 CST. This trend was corroborated by comparative analyses, which revealed significantly (p < 0.05) elevated afternoon DO levels relative to morning measurements (Fig. 2F). Conversely, non-statistically significant (p > 0.05) diurnal variations were detected in Twater, pH, or Chla concentrations between morning and afternoon sampling intervals (Figs. 2C–E; p > 0.05).Fig. 5Temporal variations in physicochemical parameters of Bailuwan Lake: Twater (A,B), pH (C,D), FNU (E,F), EC (G,H), DO (I,J), and TOC (K,L). Left panels (A,C,E,G,I,K): red lines indicate mean values per sampling time (n = 12); and right panels (B,D,F,H,J,L): black dots denote monthly means (n = 4) with whiskers representing mean ± SDs. Additional details are provided in Fig. 3.Full size imageFig. 6Temporal variations in anionic species of Bailuwan Lake: F− (A,B), Cl− (C,D), NO3− (E,F) and SO42− (G,H). Additional details are provided in Fig. 3.Full size imageMonthly monitoring revealed a shared temporal trajectory among water quality indicators including TPC, turbidity (FNU), electrical conductivity (EC), anion concentrations (F−, Cl−, and SO42−), and metal levels (potassium/K, sodium/Na, chromium/Cr) all followed an initial ascending phase followed by measurable decreases. Conversely, Chla and TN demonstrated an initial decrease followed by an increase. The parameters pH and NO3− showed a gradual decline, while magnesium/Mg displayed a consistent upward trend (Figs. 4, 5, 6 and 7). Additionally, our analysis of monthly variations in CO32− and HCO3− levels revealed distinct patterns. The CO32− generally exhibited an initial decrease followed by an increase, with a notable surge observed in March 2021. In contrast, the HCO3− reached the lowest value in July 2021 (Fig. S1).Fig. 7Monthly changes in aquatic metals of Bailuwan Lake: K (A), Na (B), Mg (C), Cu (D), Zn (E), Fe (F), Mn (G), and Cr (H). Additional details are provided in Fig. 3.Full size imageStatistical analysis demonstrated strong inverse relationships of pCO2 with pH and DO (p < 0.01, Table 3), contrasted by a direct positive association with Chla (p < 0.05, Table 4). These correlations facilitated the development of predictive linear models linking pCO2 to the key parameters (pH, DO, Chla, solar radiation/SR, Twater, and CODMn), as visualized in Figure S2.Table 3 Correlation analysis of pCO2/fCO2 with water quality parameters in the studied lakes. * and ** denote significant correlations at the 0.05 and 0.01 levels (two–tailed test), respectively. Twater, water temperature; EC, electrical conductivity; TPC, transparency; FNU, turbidity; DO, dissolved oxygen; SR, solar radiation.Full size tableTable 4 Correlation analysis of pCO2/fCO2 with nutrient status indices in the studied lake. Chla, chlorophyll a; TOC, total organic carbon; F−, fluoride; Cl−, chloride; NO3−, nitrate; SO42−, sulfate; TN, total nitrogen; TP, total phosphorus; CODMn, permanganate index. Additional details are provided in Table 3.Full size tableDiscussionNotable shiftting patterns of C-sink-source in the investigated lakeAccording to recent data released by the United Nations, the global urbanization rate has increased from 30% in 1950 to 56% in 2020, and is projected to reach 68% by 205060,61. In China, more than 60% of the permanent population had achieved urbanization by the end of 201922,38. Previous studies have demonstrated that urban lacustrine systems receive substantial inputs of exogenous labile organic carbon derived from diverse anthropogenic activities, which significantly enhances heterotrophic metabolism in these urbanized aquatic ecosystems, consequently resulting in elevated CO2 emissions. These mechanistic insights explain why metropolitan water bodies have been consistently documented42,62,63,64 as focal points for carbon release within anthropogenic landscapes. Consequently, investigating the mechanisms by which urban lakes respond to urbanization is of significant importance for predicting urban greenhouse gas emissions, particularly CO2.Across all multi-year sampling days, the pCO2 in the morning hours (07:00–11:00 CST; 2480 ± 3972 µatm; n = 22) was numerically higher than that in the afternoon periods (14:00–18:00 CST; 1098 ± 803 µatm; n = 22; Figs. 1 and 2), but the difference was not statistically significant (p > 0.05). Conversely, afternoon DO concentrations demonstrated a significant elevation compared to morning values (p < 0.05; Fig. 3F), whereas Chla levels remained stable without observable daytime fluctuations (p > 0.05; Figs. 3 and 4). These patterns align with prior work in local eutrophic lakes4,5,6. Increased solar radiation after 07:00 CST spurs photosynthesis (P); later, declining light toward 18:00 CST shifts the balance toward respiration (R). The resultant drop in dissolved oxygen is documented in Table 1 and Fig. S3. Noteworthy, temperature-driven pCO2 fluctuations demonstrated no direct significance65 (Fig. 2C). Parallel observations by Potter and Xu in a subtropical North American lake revealed pronounced diurnal pCO2 dynamics, with predawn peaks and evening troughs58. Intriguingly, nocturnal CO2 efflux rates nearly tripled daytime values, underscoring distinct day-night emission patterns. In our dataset, anomalous pCO2 behavior was observed on February 28, March 30, and November 22, 2021, where morning pCO2 (07:00–11:00 CST) slightly decreased compared to afternoon levels (Fig. 1). Measurements from February 28 and November 21 consistently demonstrated pCO2 supersaturation, with all recorded values surpassing the characteristic atmospheric CO2 range (380–420 µatm)66,67. In contrast, significantly lower pCO2 levels were recorded on March 30, 2021 (198 ± 65 µatm) and October 30, 2020 (315 ± 145 µatm; n = 4; Fig. 1), reflecting dynamic C-sink-source transitions. The underlying mechanism may be attributed to extreme precipitation events acting as a trigger. On one hand, rainfall directly reduced pCO2 through the dilution effect. On the other hand, it introduced allochthonous nutrients and enhanced water column mixing, thereby stimulating intense algal blooms. The resulting high level of photosynthetic activity served as the key biological driver leading to significantly decreased pCO2. Furthermore, as an urban wetland, its hydrology is likely influenced by anthropogenic regulation. The supplemental inflow of low-CO2 external water (e.g., reclaimed water) during this period may have further reinforced and amplified the decline in pCO26,15,68.In our study, the overall diurnal mean pCO2 level in the investigated lake (n = 48; 1789 ± 2919 µatm) was significantly higher than typical equilibrium CO2 thresholds (p < 0.05) but markedly lower than global lacustrine average pCO2 of 3230 µatm69. This finding indicates net CO2 supersaturation within the lake during our study, where in-lake CO2 production exceeded consumption, driving a net efflux of CO2 to the atmosphere and thus classifying the lake as a C-source31,44,70,71. These results align with prior investigations of eutrophic urban lakes, including Beihu Lake (~ 960 µatm)6 in the same metropolitan region, as well as Capitol Lake (~ 736 µatm)9 and University Lake (~ 630 µatm)48 in Louisiana, USA. However, the observed pCO2 levels were significantly elevated compared to our earlier findings in the same lake from January to September 2020 (~ 707 µatm)4. Importantly, pCO2 in October 2021 reached an exceptionally high value of 8148 ± 1882 µatm (n = 4; Fig. 3A), a period not captured in our prior study in 20204. However, a prior two-year comparative study of CO2 fluxes across different habitats in Lake Võrtsjärv also revealed pronounced spatial, seasonal, and interannual variability72. This anomaly underscores the potential for pronounced seasonal variability in water-air interface pCO2 dynamics, likely influenced by temporal environmental drivers48,55,59.Previous work by Wang et al., investigating 43 eutrophic lakes across China’s climatic zones56, revealed pronounced seasonal variability in pCO2 across all studied systems, with lower mean values in summer and autumn, a pattern consistent with most lacustrine studies11,12. This seasonality likely stems from synergistic effects of increased phytoplankton and submerged macrophyte biomass, coupled with thermal stratification, pH dynamics, solar radiation, and anthropogenic activities4,9,73,74. Specifically, pCO2 fluctuations in aquatic systems are governed by the balance between biological production (P; positive correlation with Chla in Table 4) and respiration (R; negative correlation with DO in Table 3)21,75. Thus, we reason that increased primary productivity during warm seasons could facilitate the transition of eutrophic lakes from net CO2 sources to sinks, whereas microbial and/or photochemical mineralization of organic carbon (e.g., TOC in Fig. 5) during cooler seasons may surpass photosynthetic CO2 uptake (also known as the biological C-pump-effect)76, thereby driving seasonal source-sink shifts6,77. Specifically, intense photosynthesis during warm seasons significantly consumes dissolved CO2, lowering pCO2 below atmospheric equilibrium and leading to CO2 influx; the lake functions as net autotrophic (i.e., an atmospheric C-sink) when gross primary production (GPP) exceeds the carbon released through ecosystem respiration. Moreovoer, our year-round monitoring (January–November 2021; Fig. 3A) showed lower mean pCO2 in the first half-year (1239 ± 781 µatm, January–June) compared to the latter period (3143 ± 4674 µatm, July–November), though both phases exceeded atmospheric equilibrium, confirming persistent CO2 supersaturation. Intriguingly, episodic CO2 undersaturation occurred in October 2020 (winter for 198 ± 65 µatm) and March 2021 (spring for 315 ± 145 µatm), temporarily converting the system to a net sink (Fig. 3). Collectively, these findings highlight diurnal/seasonal source-sink transitions in subtropical urban eutrophic lakes, modulated by dynamic biogeochemical drivers.Drivers of CO2 uptake and release in response to environmental conditionsFurther correlation analyses revealed statistically significant negative relationships (p < 0.05) between pCO2/fCO2 and both pH (− 0.727**/−0.681**) and DO (− 0.311*), alongside significant positive correlations with Chla (0.295* and 0.287*, respectively; Tables 3 and 4, and Figs. 5, S2). These findings underscore that CO2 uptake/emission dynamics in urban lakes are governed by multifaceted controls from aquatic environmental factors. In this moderately eutrophic autotrophic lake system, we posit that biological drivers, particularly Chla (as P) and DO (as R), play pivotal roles in modulating CO2 fluxes78. The enrichment of nutrients intensifies the coupling between CO2 fluxes and biogeochemical cycling, as demonstrated by strong correlations with both biological indicators (e.g., Chla) and chemical factors (e.g., TP, TN and pH; Table 2, and Fig. 4), rendering the carbon dynamics more responsive to environmental changes15,18. Specifically, enhanced organic matter mineralization increases the bioavailability of TP, thereby stimulating CO2 emissions14,49,79. The mechanistic linkage aligns with the observed inverse correlation between DO and pCO2 in our study.Prior studies have demonstrated that pH regulates the physicochemical environment of lakes by mediating the dynamic equilibrium and spatial distribution of carbonate species (CO2, CO32−, and HCO3−), thereby influencing CO2 fluxes (quantified as pCO2 and fCO2) at the water-air interface4,80,81. This chemical control is particularly pronounced in lakes with elevated pH (> 8), where aqueous CO2 concentrations exhibit marked sensitivity to alkaline conditions12. Mechanistically, higher pH promotes the conversion of free aqueous CO2 into carbonate ions, reducing pCO2 and creating undersaturation that enhances atmospheric CO2 absorption. Conversely, lower pH destabilizes dissolved inorganic carbon species, driving CO2 efflux to the atmosphere82. In our study, pH values ranged from 6.5 to 9.0, displaying distinct seasonal variability: maximal fluctuations occurred in autumn-winter (August–March), while minimal variability was observed in spring-summer (April–July; Fig. 5D). This pattern may reflect temperature-mediated modulation of hydrogen ion activity83, which is related to pH, even in the absence of statistical significance (R = − 0.095 for Twater/pH in Table 3), a finding consistent with our earlier observations4. Notably, CO2 flux exhibits strong pH dependence across diverse lacustrine systems. For instance, global analyses of 196 saline lakes revealed that lakes with pH ≥ 9 typically function as weak CO2 sinks84, while a 14-year study of six hardwater lakes in Canada’s Northern Great Plains identified a critical pH threshold of 8.6 for source-to-sink transitions85. In our dataset, seasonal shifts between C-source-sink suggest the existence of a comparable pH threshold governing flux reversals, though its precise value requires further investigation.Chla, a principal photosynthetic pigment in algae and phytoplankton, serves as a critical proxy for freshwater lake productivity79. Empirical studies have demonstrated that correlations between Chla concentrations and greenhouse gas emissions reflect the metabolic equilibrium of lacustrine ecosystems15,86,87, particularly in urban lakes influenced by industrial and domestic wastewater discharges88. In our investigation of a moderately eutrophic lake during non-bloom conditions, significant positive correlations were observed between Chla and both pCO2 (R = 0.295*) and fCO2 (R = 0.287*; Table 4), aligning with our previous findings from Beihu Lake in the same urban system6. However, earlier studies on the lake failed to detect statistically significant associations (pCO2/Chla with R = 0.202; fCO2/Chla with R = 0.213)6. Such inconsistent patterns have been documented in other lacustrine systems. Interestingly, Xu and Xu reported substantial spatiotemporal variability in Chla concentrations in University Lake89, a small urban waterbody in the southern United States, yet subsequent CO2 investigations at the same site revealed no significant Chla-CO2 relationships48. Recent analyses of 33 small artificial lakes by Wang et al.64 further demonstrate this complexity: spring algal blooms in over half of these systems resulted in annual mean Chla concentrations exceeding 10 µg L−1, coinciding with enhanced CO2 sequestration. These paradoxical findings likely stem from multiple interacting mechanisms, including vertical and lateral chemical transport dynamics in open aquatic systems, phytoplankton community composition and density variations, and methodological challenges in modeling Chla-CO2 interactions6,86. Among those, solar radiation exerts a critical regulatory influence on Chla concentrations in urban lacustrine systems through its modulation of phytoplankton photosynthetic activity89,90. Diurnally, photosynthetic efficiency follows a parabolic trajectory: morning radiation intensification stimulates photosynthetic activation, reaching maximum capacity at solar noon before subsequent photoinhibition reduces both photosynthetic performance and Chla levels through late afternoon (ref., Fig. 2E)91,92. Seasonally, spring radiation intensification creates optimal photothermal conditions for phytoplankton proliferation, enhancing lacustrine carbon sequestration potential. Conversely, summer hyper-radiation events coupled with elevated water temperatures may induce thermal stress responses, potentially triggering CO2 re-release mechanisms that could transition these aquatic systems from C-sink to -source58,93.DO, serving as a critical indicator of aquatic metabolic activity, is regulated by multiple environmental drivers including water temperature, organic matter, biological photosynthesis and water dynamics. Its dynamic equilibrium reflects compensatory atmospheric exchange and respiratory consumption processes91,94,95. In moderately eutrophic lakes (Chla < 30 µg L−1), algal growth exhibits relative moderation compared to hyper-eutrophic systems, where biogeochemical processes (particularly N–P interactions involving NO3− and NH4⁺) dominate over Chla in regulating fCO2 dynamics15,96. This regulatory dominance is particularly pronounced in urban lacustrine systems along anthropogenic disturbance gradients, where significant linear correlations between nutrient concentrations and fCO2 variations have been documented12,97. Empirical evidence from Taihu Lake demonstrates that anthropogenic nutrient loading elevates CO2 emissions through enhanced aquatic respiration40. While theoretical models suggest nutrient enrichment could reduce CO2 emissions via boosted primary production, most field observations indicate that N and P inputs predominantly amplify CO2 release through stimulation of heterotrophic respiration in both water column and sediments17,35,42,98. Our findings reveal a significant negative correlation between DO and pCO2 (p < 0.05; Table 3), indicating that, in addition to biological factors (where elevated productivity reduces CO2 efflux, leading to higher pH, increased oxygen, and lower pCO2), anthropogenically mediated heterotrophic respiration may also serve as a driver of dissolved CO2 supersaturation68,99. Further, Zhang et al. propose that DO (or apparent oxygen utilization, AOU) may serve as a more direct predictor of CO2 variability than Chla in moderately eutrophic lakes during non-bloom periods, demonstrating greater independent explanatory power for fCO2 fluctuations at the water-air interface15.Moreover, in shallow lakes where hydrodynamics are the dominant force, the movement and mixing of water constitute the core physical mechanism regulating the oxygen (O2) budget6. The continuous inflow of riverine water imparts kinetic energy, and the resulting turbulence significantly enhances gas exchange efficiency at the water-air interface, thereby promoting O2 input9. Concurrently, such hydraulic disturbance effectively disrupts thermal stratification, leading the lake toward a fully mixed state50. This process transports O2-rich surface waters to the lake bottom, preventing the formation of hypoxic conditions in the benthic zone. Further, hydrodynamics govern a relatively short hydraulic retention time, which not only exports partially decomposed organic matter18,27, reducing the internal O2 demand, but also continually replenishes nutrients to support moderate levels of photosynthetic O2 production. Therefore, through these three synergistic mechanisms, enhancing reaeration, optimizing vertical distribution, and reducing net consumption, intense hydrodynamic processes positively maintain the high-O2 equilibrium in shallow lakes.Uncertainties in CO2 evasion from subtropical lake systems and future workOur synthesis demonstrates that CO2 exchange rates in anthropogenically-impacted subtropical lakes vary considerably (− 15 to 130 mmol m−2 h−1; Table 5), reflecting strong geographic and temporal dependencies in emission patterns. The mean fCO2 in our study markedly exceeded values reported for urban-dominated lakes in other regions, even surpassing previous findings by Yang et al.4 for the same lake system (29 ± 67 mmol m−2 h−1). However, our results align with the previous observations in Qinglonghu Lake (a moderately eutrophic urban lake in the same region; 108 ± 101 mmol m⁻2 h−1)10, likely attributable to intensified anthropogenic pollution inputs and aggravated eutrophication in our study lakes during the monitoring period. Notably, fCO2 values from the highly urbanized tropical Rio Grande Reservoir in South America95, i.e., 5.14 and 3.18 mmol m−2 h−1 for hypereutrophic and moderately eutrophic zones, respectively, were substantially lower than our measurements for the moderately eutrophic in this lake. These discrepancies may reflect not only differences in eutrophication status and nutrient levels but also hydrological seasonality, as exemplified by the previous work on subtropical University Lake in Louisiana, USA48. While local anthropogenic forces are often the dominant driver of CO2 dynamics in human-modified lakes, we hypothesize that a portion of the residual uncertainties in regional-scale CO2 emission estimates for subtropical lakes is associated with patterns of regional climate change15,31. Specifically, monsoon-driven climatic patterns characterized by high temperatures and/or heavy rainfall amplify interannual fluctuations in hydrological conditions (e.g., rainfall-evaporation balance), thereby modulating the dissolution-release equilibrium of CO2.Previously, Zhang et al.15 demonstrated that most lakes across diverse geographical regions including Taihu Lake, Lake Guadalcacín, Lake Bornos, Lake Alexandrina, and urban artificial ponds function as atmospheric CO2 sources, with annual mean fCO2 increasing alongside Chla concentrations (Table 5). These comparative findings further emphasize the variability of CO2 dynamics across interannual, seasonal, and hour scales6,58,100. For instance, seasonal analyses revealed a C-source-sink alternation pattern in subtropical urban lakes, where winter and spring (periods of low algal biomass) exhibit CO2 production and release, while elevated summer primary productivity facilitates a transition to CO2 sinks18. Contrary to this typical seasonal pattern, our study observed no strict as the summer-sink and winter/spring-source dynamic. Despite two months of pCO2 levels significantly below atmospheric equilibrium (Fig. 3), the annual mean pCO2 (1789 µatm) remained markedly supersaturated, indicating persistent CO2 emissions. Further, discrepancies emerged when comparing our with the previous findings of Yang et al.4 for the same lake system (707 µatm), suggesting that such divergence in seasonal patterns contributes to uncertainties in urban lake CO2 emission assessments. Similarly, Potter and Xu58 highlighted that while summer predominantly manifests as a C-source and winter as a sink, spring sampling may prove valuable for assessing CO2 evasion dynamics in shallow trophic lakes. As a hypothesis, early-spring likely represents a critical transitional window, our future studies accordingly will prioritize through systematic monitoring of this period.Prior studies has established that while long-term pCO2 variations (days to months) significantly influence evasion rate estimates, gas transfer velocity (k600) emerges as a critical regulator of diurnal CO2 evasion at shorter timescales (minutes to hours)100,101. Building on this understanding, both previous studies and our work observed distinct diurnal water-air CO2 gradients from dawn to dusk, yet collectively underestimated pCO2 owing to insufficient consideration of nocturnal dynamics. Through 24-h monitoring of subtropical urban lakes, previous study revealed pronounced diurnal fluctuations in pCO2 and CO2 degassing, particularly marked by nocturnal pCO2 surges58. Their findings identified 10:00 and 22:00 CST as periods of minimal deviation from daily mean pCO2, recommending optimized sampling between 09:00–11:00 CST to balance accuracy and operational feasibility. Our study validated the efficacy of sensor-based continuous monitoring in lacustrine systems. Similarly, Wang et al.55 demonstrated that daytime CO2 fluxes in Tangxun Lake (ca. 8 mmol m−2 day−1) were remarkedly lower than nocturnal fluxes (ca. 10 mmol m−2 day−1), with 11:00–12:00 measurements best approximating daily means. It could therefore be inferred that during nighttime hours, when photosynthetic activity is minimal, enhanced CO2 evasion is likely to occur. Nevertheless, only a limited number of studies have focused on nocturnal CO₂ release dynamics. For instance, analysis of long-term limnological data (1987–2006) from Lake Apopka, Florida, by Gu et al.102 revealed consistently higher average partial pressure of CO2 (224 µatm) at nighttime compared to daytime levels. Similarly, Reis and Barbosa103 reported significantly elevated mean nocturnal pCO2 (565 µatm) relative to daytime values (436 µatm) in a tropical productive lake in southeastern Brazil. More recently, Reiman and Xu104 further corroborated a consistent daytime pattern of pCO2 in the Lower Mississippi River, characterized by a peak prior to sunset and a minimum during periods of maximum solar irradiance. These collective findings underscore the critical role of temporal resolution in evasion rate quantification.Methodologically constrained by funding limitations, manpower shortages, and site accessibility challenges, our study lacked nocturnal monitoring and employed a limited sample size. While these findings provide preliminary insights into CO2 dynamics at water-air interfaces, representing an initial exploratory step. For example, there are still the following urgent challenges that need to be addressed in our current work:

    (i)

    This study conducted high-frequency monitoring at a representative site located 2 m from the lake’s shoreline. This location was selected for its capacity to reflect the dominant air-water exchange processes in the lake’s open waters while avoiding known, strong point-source disturbances. Although spatial variability of CO2 concentrations in the well-mixed central basin of this moderately sized urban lake is likely limited in the absence of major perturbations, we acknowledge that a single sampling point cannot fully capture the potential spatial heterogeneity of the entire lake. Further, its proximity to the shore may not adequately represent biogeochemical processes in the pelagic zone. For instance, nearshore areas susceptible to groundwater inflow or sediment resuspension may develop localized pCO2 hotspots, which were not directly monitored in this study design. Additionally, the presence of submerged aquatic vegetation in the sediments at the sampling site can influence dissolved CO2 concentrations and their diel fluctuations. The spatial heterogeneity of aquatic vegetation introduces uncertainty when extrapolating discrete point measurements to whole-ecosystem fCO2. While the current sampling strategy adheres to standard protocols, it may not fully represent the metabolic diversity across different habitats. In essence, the net CO2 flux in vegetated areas represents a dynamic balance between photosynthetic uptake and respiration, which varies nonlinearly with environmental conditions. Moreover, the inhibitory effect of vegetation canopies on the gas transfer velocity (k) is often overlooked in flux calculations, potentially leading to systematic overestimation in these areas. Consequently, the CO2 values reported herein should be interpreted as the best available estimate of the dominant fluxes in the lake’s nearshore zone, rather than an absolute and precise whole-lake average. Future investigation should aim to better quantify this spatial variability and its impact on the integrated lake carbon flux budget by deploying more extensive sensor networks and integrating high-resolution pCO2 mapping with habitat-specific parameterization of k.

    (ii)

    As noted previously, safety and technical constraints prevented our high-frequency monitoring from covering the nocturnal period. This may introduce uncertainty in our estimates of diel CO2 fluxes, particularly in quantifying nighttime CO2 emissions dominated by ecosystem respiration, potentially leading to an underestimation of total daily CO2 emissions. To assess this uncertainty and provide reasonable flux estimates to the extent possible, we propose that the daily mean flux can be extrapolated based on a conservative estimate of nighttime flux, for instance, by assuming that nighttime fluxes are similar to the lower flux levels observed around sunset. An analysis of the potential systematic bias introduced by the absence of nighttime data indicates that, even in seasons when the lake consistently acts as a CO2 source, and under a worst-case scenario where nighttime fluxes reach daytime peak levels, the core conclusion, that the lake may shift from a CO2 source to a sink in certain seasons (e.g., during summer algal blooms), remains valid. This is because the strong daytime photosynthetic uptake is sufficient to offset the estimated upper bound of respiratory emissions at night. Therefore, we emphasize that the key finding of this study, the observed source-sink transition dynamics of the lake, is primarily driven by strong daytime biogeochemical processes (i.e., photosynthesis vs. respiration/chemical equilibria). Although the lack of direct nighttime observations introduces a degree of uncertainty, it does not undermine our understanding of the principal mechanisms driving these dynamics. Future investigations should place greater emphasis on investigating CO2 dynamics during the nighttime.

    (iii)

    The pCO2 in this study was calculated from pH, temperature, and alkalinity using thermodynamic equilibrium equations. This approach may introduce significant deviations under conditions of dissolved organic carbon (DOC) or extreme pH, potentially leading to overestimation of pCO2. The combination of in-situ direct measurements and multi-method comparisons is necessary, mainly to reduce uncertainties and enhance data reliability. Moreover, the constraining the precise contribution of these allochthonous inputs to the lake’s CO2 emissions entails considerable uncertainties. The pulsed nature of carbon delivery during rainfall events is poorly captured by our monthly sampling, likely leading to an underestimation of episodic inputs. Further, the heterogeneous composition and bioavailability of imported carbon (e.g., labile DOC from sewage vs. refractory DOC from soil erosion) make it difficult to predict its mineralization efficiency and thus its ultimate contribution to CO2 evasion. Disentangling the effects of external carbon from in-lake processes remains a major challenge, as these drivers are often coupled (e.g., nutrient inputs stimulating productivity that consumes CO2).

    Overall, future work require standardized protocols incorporating temporal (diurnal/seasonal) and spatial (cross-regional urban lake selections) dimensions to reduce uncertainties in CO2 flux estimates for urban lacustrine systems. Given the critical role of eutrophication in modulating C-source-sink transitions in urbanized lakes, we propose implementing dual-objective environmental management strategies, such as pollution mitigation (intercepting pollutant inputs through watershed management), and C-conscious restoration including optimizing hydrophyte communities through, C-sequestration species selection, biodiversity enhancement, and spatial configuration optimization. These measures should be prioritized in subtropical regions prone to algal blooms and climatic warming, aiming to simultaneously improve water quality and align lacustrine carbon budgets with global carbon neutrality targets.Table 5 The global comparison of CO2 fluxes heterogeneity in subtropical lakes.Full size tableConclusionsThis work employed a high-frequency observational program to examine temporal and spatial patterns of pCO2 and fCO2 in relation to key environmental variables within a subtropical urban lake with moderate eutrophic status. Temporal analysis identified consistent afternoon reductions in aquatic CO2 parameters during daytime (07:00–18:00 CST), contrasting with anomalous measurements recorded on three specific dates. However, no statistically significant differences in pCO2/fCO2 levels were detected between morning and afternoon (p > 0.05, n = 24). During the investigation, the average pCO2 and fCO2 levels were measured at 1788.61 µatm and 130.26 mmol m⁻2 h⁻1 (n = 48), respectively. Monthly analyses revealed substantial variability in both parameters, with pCO2 ranging from 198.14 to 8148.02 µatm, and fCO2 from − 16.14 to 775.16 mmol m−2 h−1 (n = 4, monthly). The annual cycle (January–November 2021) showed significantly lower mean pCO2 during the first half-year (1239.34 µatm, January–June) compared to the latter period (3143.15 µatm, July–November), both exceeding atmospheric equilibrium to function as net CO2 source. Crucially, episodic CO2 undersaturation occurred in October 2020 (198.14 µatm) and March 2021 (314.92 µatm), temporarily converting the system to a carbon sink. Statistical analyses identified pH, Chla (P), and DO (R) as key environmental drivers of pCO2 and CO2 flux variability, while eutrophication status and anthropogenic disturbances critically modulated source-sink transitions. These findings highlight the urgent need for improved management strategies in urban lake systems, such as reducing pollutants and mitigating carbon emissions, supported by standardized protocols that account for temporal (especially nocturnal), seasonal, and regional variations. Such integrated approaches will enhance the accuracy of CO2 flux estimates and contribute to global carbon neutrality goals.MethodsSite descriptionThe study was conducted at Bailuwan Lake (104° 7′ 40.5″ E, 30° 34′ 56.01″ N), an urban water body situated in the peri‑urban transition zone of Chengdu, Sichuan, China (Fig. S4). This artificially established aquatic system functions as an integrated urban eco‑wetland complex under the management of the Jinjiang District Government, combining tourism with ecological conservation. Designated in 2017 as Chengdu’s first National Urban Wetland Park, the lake represents a typical urban water body in a western Chinese megacity and offers a valuable case study for examining common features and challenges, such as ecological functions, environmental pressures, and management practices, of urban lakes globally.The lake covers a total surface area of 200 ha, with open water bodies accounting for approximately 33.5% of the area and exhibiting depth gradients ranging from 0.5 to 6.5 m. According to our previous study based on 2021 data, vegetation represented the largest land cover type (56.2%) in the study area, followed by bare land, lake surface, and roads (6.8%)105. Hydrologically, the main inflow is an engineered tributary of the Dongfeng Canal system, while water export occurs mainly through evaporation and controlled discharge via constructed drainage infrastructure.The study area is situated within a humid subtropical monsoon climate zone, characterized by pronounced seasonal thermal variability. Meteorological records indicate a mean annual temperature of 16.5 °C, with a distinct seasonal pattern featuring the lowest monthly temperatures in January and peak values during July–August, aligning with the broader regional climate regime. Throughout the monitoring campaign, diurnal air temperature fluctuations in the lake vicinity were substantial, varying from 0 °C to 35 °C (Fig. S3A), reflecting strong day-night thermal dynamics. In addition, the region experiences considerable solar exposure, with an annual cumulative solar radiation measuring approximately 161 kJ cm− 2 (Fig. S3B). This high level of irradiance plays a critical role in driving both hydrological and ecological processes, underscoring the distinctive energy budget setting of this subtropical lacustrine environment.Field measurementsThis investigation employed a monthly field monitoring protocol to collect essential hydrological parameters from October 2020 through November 2021, with the exception of November 2020 and September 2021 due to logistical constraints. Additionally, the scheduled July 2021 field campaign was administratively rescheduled to August 2, 2021. In other words, a total of 12 in-situ measurements were collected over 13 months of investigation. To ensure methodological rigor and data reliability, a triplicate sampling approach (n = 3) was systematically implemented for each field collection event, with samples subsequently subjected to both in-situ measurements and comprehensive laboratory analyses.To maintain rigorous data quality standards and ensure cross-comparability, standardized sampling was performed at four fixed time points daily (07:00, 11:00, 14:00, and 18:00 CST) using a plastic grab-sampler at a depth of 30–50 cm below the water surface. All trips were made on sunny days to minimize rainfall and/or stormwater runoff effects on water conditions. Moreover, one objective of this study was to capture CO2 dynamics in the near-shore shallow water area, a typical ecotone, whose distinct characteristics are often homogenized and overlooked in whole-lake scale studies. Therefore, the sampling site was selected 2 m (Fig. S4) from the shore primarily because this location represents a sensitive zone for CO2 exchange at the water-air interface and is also significantly influenced by anthropogenic activities (e.g., surface runoff and input from riparian vegetation). Further, our investigation revealed the presence of submerged aquatic vegetation, such as Ceratophyllum demersum and Potamogeton distinctus, distributed on the lakebed directly below the sampling point.Water transparency was assessed through Secchi disk measurements (TPC; cm), employing a standardized 20-cm diameter disk, with concurrent turbidity determinations (FNU; NTU) performed using a calibrated HACH-TSS turbidimeter (Danaher Corporation, Washington, DC, USA). Concurrently, a Hanna-HI9829/HI98186 multiparameter probe (Hanna Instruments, Italy) was deployed for synchronous in situ measurements of fundamental water characteristics: pH, DO (mg L−1), EC (µS cm−1), and Twater (°C). The quantification of pCO2 required precise determination of carbonate (CO32−; CB; mol mL−1) and bicarbonate (HCO3−; BCB; mol mL−1) concentrations via acid-base titration, utilizing phenolphthalein and methyl orange as dual-endpoint indicators in accordance with the standardized analytical procedure (Method S2).Additionally, during the dynamic monitoring phase, continuous in-situ measurements of water quality parameters were conducted only after the readings had stabilized and remained consistent for an additional 5 min. Moreover, the stabilization process typically required 15–30 min. For water samples intended for TOC and anion analysis, filtration through a 0.45 μm micropore membrane filter was performed prior to storage in pre-cleaned polyethylene bottles. All samples were stored in high-density polyethylene bottles that had been pre-acid washed, tightly sealed, and carefully inspected to prevent gas exchange. During transportation, samples were placed in coolers with sufficient wet-ice to maintain preservation conditions.Furthermore, comprehensive laboratory analyses were conducted on water samples to quantify multiple physicochemical parameters. The measured parameters encompassed: (i) Chla (mg m−3) quantified per the National Environmental Protection Agency (NEPA) standards106, (ii) nitrate (NT; mg L−1) analyzed following Method S3, and (iii) TP (mg L−1) assessed via Method S4. Additionally, TOC was fractionated into total carbon (TC) and inorganic carbon (IC) components (mg L−1) using validated methodologies4,5. Additional parameters assessed were TDN (mg L− 1; Method S5), CODMn (mg L− 1; Method S6). Anionic species (F−, Cl−, SO42−, NO3−; mg L− 1) were quantified following previously validated methodologies4,5,6. The trophic status of water bodies was evaluated through the Carlson’s TSI (trophic state index; Method S1).Besides, given its location within an urban setting, the concentrations and distribution patterns of metals in Baihuwan lake could serve as sensitive tracers of anthropogenic disturbances, such as discharges from urban wastewater and surface runoff. These indicators assist in evaluating the potential influence of allochthonous carbon inputs on the overall carbon balance of the system87. For instance, the speciation and solubility of metals such as Fe and Mn are strongly dependent on ambient redox conditions45,87. Variations in their concentrations and ratios could be used to infer dominant pathways of organic matter degradation (e.g., Fe and Mn reduction processes), which are closely linked to the production and consumption of greenhouse gases including CO2. Therefore, in this study, we also measured trace metal concentrations (Cu, Zn, Cr, Fe, Mn, K, Ca, Na, Mg; µg L− 1; see Method S7).To ensure sample integrity, water specimens were collected in amber glass bottles and maintained at 4 °C during transportation. All field sampling procedures were strategically conducted during periods of meteorological stability (clear weather conditions) to eliminate potential confounding effects from precipitation or surface runoff on analytical outcomes.Laboratory analysesCalculation of pCO2
    Extensive studies has demonstrated that the equilibrium distribution of aqueous carbonate species, encompassing bicarbonate, carbonate, carbonic acid, and dissolved CO2, is principally governed by a suite of physicochemical parameters, specifically hydrogen ion activity (pH), aqueous temperature (Twater), and ionic strength (IS) of the solution4,23,107,108. Building upon this empirical foundation, the present study employs a thermodynamic carbonate speciation model (Eqs. 1–4) to computationally derive the aqueous carbon dioxide partial pressure (pCO2, expressed in µatm), with rigorous implementation protocols detailed in Method S8.$$p{text{CO}}_{{text{2}}} = left[ {{text{H}}_{{text{2}}} {text{CO}}_{{text{3}}} } right]/K_{{{text{CO2}}}} , = ,aleft( {{text{H}}^{ + } } right) cdot a({text{HCO}}_{{text{3}}} ^{ – } )/left( {K_{{{text{CO2}}}} cdot K_{{text{1}}} } right)$$
    (1)
    $$alpha left( {{text{H}}^{ + } } right){mkern 1mu} = {mkern 1mu} 10^{{ – [{text{pH}}]}}$$
    (2)
    $${{alpha (HCO}}_{3}^{-} {text{)}} = {text{[HCO}}_{3}^{-} {text{]}} times mathop {10}nolimits^{{-0.5sqrt {text{I}} }}$$
    (3)
    $$begin{gathered} I, = ,0.{text{5}}(left[ {{text{K}}^{ + } } right], + ,{text{4}}left[ {{text{Ca}}^{{{text{2}} + }} } right]{text{ }} + {text{ }}left[ {{text{Na}}^{ + } } right], + ,{text{4}}left[ {{text{Mg}}^{{{text{2}} + }} } right] hfill \ {text{ }} + {text{ }}left[ {{text{Cl}}^{ – } } right], + ,{text{4}}left[ {{text{SO}}_{{text{4}}} ^{{{text{2}} – }} } right]{text{ }} + {text{ }}left[ {{text{NO}}_{{text{3}}} ^{ – } } right] + [{text{HCO}}_{{text{3}}} ^{ – } ])/{text{1}}000000 hfill \ end{gathered}$$
    (4)
    where the terms α(H+) and α(HCO3−) represent the chemically active fractions of hydrogen [H+] and bicarbonate [HCO3−] ions, accounting for non-ideal solution effects, whereas I quantifies the cumulative electrostatic environment through ionic strength.Furthermore, it should be noted that the concentration of CO2 in the lake water was calculated from bicarbonate alkalinity, using pH and temperature as the relevant thermodynamic parameters. However, this computational approach could lead to significant overestimation of CO2 under high DOC conditions (> 200 µmol L− 1). Therefore, prior to each measurement, pre-screening was conducted to ensure that the hydrochemical conditions of the lake, including pH, alkalinity, and DOC (or TOC in this study) concentration, remained within a “safe range” for reliable estimation.Estimation of fCO2
    Empirical studies have systematically demonstrated that the interfacial CO2 flux across the aquatic-atmospheric boundary layer is predominantly regulated by a complex interplay of environmental variables, including thermal conditions (temperature), solute concentration (salinity), atmospheric turbulence (wind speed), and the pCO269,109. In accordance with these established principles, we implemented a mechanistic transfer model (Eq. 2) to quantify the net fCO2 at the water-air interface, with flux density expressed in standardized units of mmol m− 2 h− 1, following the comprehensive methodological framework in Method S9.$$f {text{CO}}_{{text{2}}} , = ,K_{{text{T}}} K_{{text{H}}} left[ {p{text{CO}}_{{{text{2}}({text{water}})}} , – ,p{text{CO}}_{{{text{2}}({text{air}})}} } right]$$
    (5)
    where fCO2 quantifies the net exchange rate of CO2 per unit area at the water-air boundary interface. The parameters KH and KT correspond to the temperature-dependent Henry’s law constant (quantifying CO2 solubility) and the gas transfer velocity (characterizing the water-air exchange coefficient), respectively. The determination of KH, following established thermodynamic principles, incorporates a multivariate dependence on physicochemical conditions, specifically thermal regime (temperature), ionic strength (salinity), and hydrostatic pressure, as comprehensively characterized in the seminal work of Weiss110.$$mathop Knolimits_{H} = mathop {text{e}}nolimits^{{[mathop Anolimits_{1} + mathop Anolimits_{2} (100/T) + mathop Anolimits_{3} (T/100)]}}$$
    (6)
    Moreover, the normalized dimensionless Schmidt number (K600) was computationally transformed to the CO2-specific gas transfer velocity (KT) using the functional relationship expressed in Eq. (7). This transformation incorporates the well-established functional dependence between the Schmidt number (Sc) and gas exchange kinetics, thereby facilitating the precise estimation of KT across a range of environmental conditions, as originally demonstrated in the foundational work of Jahne et al.111$$:{text{}text{K}}_{text{T}}text{=}{text{K}}_{text{600}}times(frac{text{600}}{{text{Sc}}_{text{CO}text{2}}}text{)}{text{}}^{text{n}}text{}$$
    (7)
    where the exponent of the Schmidt number, denoted as n, exhibits variability contingent upon wind speed conditions. The exponent n adopts a value of 0.5 for wind speeds > 3.7 m s− 1, decreasing to 0.75 for calmer conditions (< 3.7 m s− 1), as established by Guérin et al.112 For the Schmidt number parameterization, we applied the widely accepted value of 0.67 following Cole and Caraco’s experimental determinations under reference conditions107. Furthermore, the computational algorithm for K600, as expressed in Eq. (8), accompanied by its comprehensive methodological elucidation, is presented in Method S9. In this study, U10 denotes wind speed values corrected to the conventional 10-m reference height over the water body at sampling time.$$K_{{{text{6}}00}} , = ,{text{2}}.0{text{7}}, + ,0.{text{215}}U_{{{text{1}}0}} ^{{{text{1}}.{text{7}}}}$$
    (8)
    Statistical analysisIn the present study, the statistical framework employed the IBM-SPSS Statistics 22 (IBM Corp., USA) for comprehensive data analysis, adopting a 95% confidence level (α = 0.05) to ensure analytical robustness. High-quality data visualization was achieved using SigmaPlot 14.0 (Systat Software Inc., USA), enabling precise graphical interpretation. This dual-platform approach complies with contemporary standards for quantitative research methodology, guaranteeing both statistical validity and visual clarity.

    Data availability

    All data generated or analysed during this study are included in this published article and its supplementary information files.
    ReferencesSaunois, M. et al. The global methane budget 2000–2017. Earth Syst. Sci. Data. 12, 1561–1623 (2020).Article 
    ADS 

    Google Scholar 
    Li, X., Shi, F., Ma, Y., Zhao, S. & Wei, J. Significant winter CO2 uptake by saline lakes on the Qinghai–Tibet plateau. Glob. Change Biol. 28, 2041–2052 (2022).Article 
    ADS 

    Google Scholar 
    Sun, H. et al. Drivers of Spatial and seasonal variations of CO2 and CH4 fluxes at the sediment water interface in a shallow eutrophic lake. Water Res. 222, 118916 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Yang, R. et al. Significant daily CO2 source–sink interchange in an urbanizing lake in Southwest China. Water 15, 3365 (2023).Article 
    CAS 

    Google Scholar 
    Yang, R. et al. Assessing the landscape ecological health (LEH) of wetlands: research content and evaluation methods (2000–2022). Water 15, 2410 (2023).Article 

    Google Scholar 
    Yang, R. et al. The shifting pattern of CO2 source sink in a subtropical urbanizing lightly eutrophic lake. Sci. Total Environ. 946, 174376 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Verpoorter, C., Kutser, T., Seekell, D. A. & Tranvik, L. J. A global inventory of lakes based on high–resolution satellite imagery. Geophys. Res. Lett. 41, 6396–6402 (2014).Article 
    ADS 

    Google Scholar 
    Tranvik, L. J. et al. Lakes and reservoirs as regulators of carbon cycling and climate. Limnol. Oceanogr. 54, 2298–2314 (2009).Article 
    ADS 
    CAS 

    Google Scholar 
    Yang, R., Xu, Z., Liu, S. & Xu, Y. J. Daily pCO2 and CO2 flux variations in a subtropical mesotrophic shallow lake. Water Res. 153, 29–38 (2019).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Yang, R. et al. Daily variations in pCO2 and fCO2 in a subtropical urbanizing lake. Front. Earth Sci. 9, 805276 (2022).Article 
    ADS 

    Google Scholar 
    Cole, J. J., Caraco, N. F., Kling, G. W. & Kratz, T. K. Carbon dioxide supersaturation in the surface waters of lakes. Science 265, 1568–1570 (1994).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Wen, Z. et al. Re-estimating china’s lake CO2 flux considering Spatiotemporal variability. Environ. Sci. Ecotech. 19, 100337 (2024).CAS 

    Google Scholar 
    Holgerson, M. A. & Raymond, P. A. Large contribution to inland water CO2 and CH4 emissions from very small ponds. Nat. Geosci. 9, 222–226 (2016).Article 
    ADS 
    CAS 

    Google Scholar 
    Cole, J. J. et al. Plumbing the global carbon cycle: integrating inland waters into the terrestrial carbon Budge. Ecosystems 10, 171–184 (2007).Article 
    CAS 

    Google Scholar 
    Zhang, L., Xu, Y. J. & Li, S. Changes in CO2 concentration and degassing of eutrophic urban lakes associated with algal growth and decline. Environ. Res. 237, 117031 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Li, S. et al. Large greenhouse gases emissions from china’s lakes and reservoirs. Water Res. 147, 13–24 (2018).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Ran, L. et al. Substantial decrease in CO2 emissions from Chinese inland waters due to global change. Nat. Commun. 12, 1730 (2021).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zhang, L., Xu, Y. J. & Li, S. Source and quality of dissolved organic matter in streams are reflective to land use/land cover, climate seasonality and pCO2. Environ Res. 216, 114608 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Downing, J. A. et al. The global abundance and size distribution of lakes, ponds, and impoundments. Limnol. Oceanogr. 51, 2388–2397 (2006).Article 
    ADS 

    Google Scholar 
    Lin, P. et al. Hotspots of riverine greenhouse gas (CH4, CO2, N2O) emissions from Qinghai Lake Basin on the northeast Tibetan Plateau. Sci. Total Environ. 857, 159373 (2023).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Karim, A., Dubois, K. & Veize, J. Carbon and oxygen dynamics in the Laurentian great lakes: implications for the CO2 flux from terrestrial aquatic systems to the atmosphere. Chem. Geol. 281, 133–141 (2011).Article 
    ADS 
    CAS 

    Google Scholar 
    Raymond, P. A. et al. Global carbon dioxide emissions from inland waters. Nature 503, 355–359 (2013).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Abril, G. et al. Amazon river carbon dioxide outgassing fuelled by wetlands. Nature 505, 395–398 (2014).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Wang, G., Liu, S., Sun, S. & Xia, X. Unexpected low CO2 emission from highly disturbed urban inland waters. Environ. Res. 235, 116689 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Foley, E. & Steinman, A. D. Urban lake water quality responses to elevated road salt concentrations. Sci. Total Environ. 905, 167139 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Xiao, Q. et al. Management actions mitigate the risk of carbon dioxide emissions from urban lakes. J. Environ. Manag. 344, 118626 (2023).Article 
    CAS 

    Google Scholar 
    Li, S., Luo, J., Xu, Y. J., Zhang, L. & Ye, C. Hydrological seasonality and nutrient stoichiometry control dissolved organic matter characterization in a headwater stream. Sci. Total Environ. 807, 150843 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Zhu, Z. et al. Land–water transport and sources of nitrogen pollution affecting the structure and function of riverine microbial communities. Environ. Sci. Technol. 57, 2726–2738 (2023).Article 
    ADS 
    PubMed 

    Google Scholar 
    Pacheco, F. S., Roland, F. & Downing, J. A. Eutrophication reverses whole-lake carbon budgets. Inland. Waters. 4, 41–48 (2013).Article 

    Google Scholar 
    Finlay, K. et al. Decrease in CO2 efflux from Northern Hardwater lakes with increasing atmospheric warming. Nature 519, 215–218 (2015).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Tao, Y. et al. Evolution of CO2 flux over 60 years: identifying source and sink changes caused by eutrophication of Hulun Lake. Sci. Total Environ. 953, 176052 (2024).Article 
    CAS 

    Google Scholar 
    Xiao, Q. et al. Lakes shifted from a carbon dioxide source to a sink over past two decades in China. Sci. Bull. 69, 1857–1861 (2024).Article 

    Google Scholar 
    Tang, W., Xu, X. J. & Li, S. Rapid urbanization effects on partial pressure and emission of CO2 in three rivers with different urban intensities. Ecol. Indic. 125, 107515 (2021).Article 
    CAS 

    Google Scholar 
    Wang, J. et al. pCO2 and CO2 evasion from two small suburban rivers: implications of the watershed urbanization process. Sci. Total Environ. 788, 147787 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Zhou, T. et al. Characteristics and influencing factors of CO2 emission from inland waters in China. Sci. China Earth Sci. 67, 2034–2055 (2024).Article 
    ADS 
    CAS 

    Google Scholar 
    Zhou, C. et al. Cyanobacteria decay alters CH4 and CO2 produced hotspots along vertical sediment profiles in eutrophic lakes. Water Res. 265, 122319 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    van Bergen, T. J. H. M. et al. Seasonal and daytime variation in greenhouse gas emissions from an urban pond and its major drivers. Limnol. Oceanogr. 64, 2129–2139 (2019).Article 
    ADS 

    Google Scholar 
    Rentschler, J. et al. Global evidence of rapid urban growth in flood zones since 1985. Nature 622, 87–92 (2023).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Waajen, G., van Oosterhout, F., Douglas, G. & Lurling, M. Geo-engineering experiments in two urban ponds to control eutrophication. Water Res. 97, 69–82 (2016).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Xiao, Q. et al. Eutrophic lake Taihu as a significant CO2 source during 2000–2015. Water Res. 170, 115331 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Xiao, Q. T. et al. Environmental investments decreased partial pressure of CO2 in a small eutrophic urban lake: evidence from long-term measurements. Environ. Pollut. 263, 114433 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Pickard, A. et al. Greenhouse gas budgets of severely polluted urban lakes in India. Sci. Total Environ. 798, 149019 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Ran, L. et al. Seasonal and diel variability of CO2 emissions from a semiarid hard-water reservoir. J. Hydrol. 608, 127652 (2022).Article 
    CAS 

    Google Scholar 
    Zhao, F. et al. Seasonal pattern of diel variability of CO2 efflux from a large eutrophic lake. J. Hydrol. 645, 132259 (2024).Article 
    CAS 

    Google Scholar 
    Sobek, S. et al. The catchment and climate regulation of pCO2 in boreal lakes. Glob. Change Biol. 9, 630–641 (2003).Article 
    ADS 

    Google Scholar 
    Marotta, H. et al. Long-term CO2 variability in two shallow tropical lakes experiencing episodic eutrophication and acidification events. Ecosystems 13, 382–392 (2010).Article 
    CAS 

    Google Scholar 
    Kosten, S. et al. Climate–dependent CO2 emissions from lakes. Glob. Biogeochem. Cycles. 24, GB2007 (2010).Article 
    ADS 

    Google Scholar 
    Xu, Y. J., Xu, Z. & Yang, R. Rapid daily change in surface water pCO2 and CO2 evasion: a case study in a subtropical eutrophic lake in Southern USA. J. Hydrol. 570, 486–494 (2019).Article 
    ADS 
    CAS 

    Google Scholar 
    Katkov, E. & Fussmann, G. F. The effect of increasing temperature and pCO2 on experimental pelagic freshwater communities. Limnol. Oceanogr. 68, S202–S216 (2023).Article 
    ADS 

    Google Scholar 
    Pu, J. et al. Varying thermal structure controls the dynamics of CO2 emissions from a subtropical reservoir, South China. Water Res. 178, 115831 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Attermeyer, K. et al. Carbon dioxide fluxes increase from day to night across European streams. Commun. Earth Environ. 2, 118 (2021).Article 
    ADS 

    Google Scholar 
    Couturier, M. et al. Long–term trends in pCO2 in lake surface water following rebrowning. Geophys. Res. Lett. 49, e2022GL097973 (2022).Article 
    ADS 
    CAS 

    Google Scholar 
    Jiang, B. et al. Wetland CH4 and CO2 emissions show opposite temperature dependencies along global climate gradients. Catena. 248, 108557 (2025).Article 
    CAS 

    Google Scholar 
    MacIntyre, S. et al. Buoyancy flux, turbulence, and the gas transfer coefficient in a stratified lake. Geophys. Res. Lett. 37, L24604 (2010).Article 
    ADS 

    Google Scholar 
    Wang, X. et al. Spatial dynamics of pCO2 and CO2 emissions from eutrophic lakes. Ecol. Indic. 166, 112529 (2024).Article 
    CAS 

    Google Scholar 
    Wang, Y. et al. Diel variability of carbon dioxide concentrations and emissions in a largest urban lake, central china: insights from continuous measurements. Sci. Total Environ. 912, 168987 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Sun, H. et al. Eutrophication decreased CO2 but increased CH4 emissions from lake: a case study of a shallow lake Ulansuhai. Water Res. 201, 117363 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Potter, L. & Xu, Y. J. Can a eutrophic lake function as a carbon sink? Case study of a subtropical eutrophic lake in Southern USA. J. Hydrol. 625, 130071 (2023).Article 
    CAS 

    Google Scholar 
    Yang, P. et al. Variation characteristics and influencing mechanism of CO2 flux from lakes in the Badain Jaran desert: a case study of Yindeer lake. Ecol. Indic. 127, 107731 (2021).Article 
    CAS 

    Google Scholar 
    Li, S. Y., Bush, R. T., Ward, N. J., Sullivan, L. A. & Dong, F. Y. Air–water CO2 outgassing in the lower lakes (Alexandrina and Albert, Australia) following a millennium drought. Sci. Total Environ. 542, 453–468 (2016).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Ritchie, H., Samborska, V., Roser, M. & Urbanization Our world in data. https://ourworldindata.org/urbanization?source=content_type:react|first_level_url:article|section:main_content|button:body_link (2024).Peacock, M. et al. Small artificial waterbodies are widespread and persistent emitters of methane and carbon dioxide. Glob. Change Biol. 27, 5109–5123 (2021).Article 
    ADS 
    CAS 

    Google Scholar 
    Fan, L. et al. Spatiotemporal patterns and drivers of CH4 and CO2 fluxes from rivers and lakes in highly urbanized areas. Sci. Total Environ. 918, 170689 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Wang, L. et al. Utilization patterns strongly dominated the dynamics of CO2 and CH4 emissions from small artificial lakes. J. Environ. Manag. 373, 123613 (2025).Article 
    CAS 

    Google Scholar 
    Nimick, D. A., Gammons, C. H. & Parker, S. R. Diel biogeochemical processes andtheir effect on the aqueous chemistry of streams: a review. Chem. Geol. 283, 3–17 (2011).Article 
    ADS 
    CAS 

    Google Scholar 
    Shao, C. L. et al. Diurnal to annual changes in latent, sensible heat, and CO2 fluxes over a Laurentian great lake: a case study in Western lake Erie. J. Geophys. Res. Biogeosci. 120, 1587–1604 (2015).Article 
    CAS 

    Google Scholar 
    Jia, J. et al. Determining whether Qinghai–Tibet plateau waterbodies have acted like carbon sinks or sources over the past 20 years. Sci. Bull. 67, 2345–2357 (2022).Article 

    Google Scholar 
    Li, S., Luo, J., Wu, D. & Xu, Y. J. Carbon and nutrients as indictors of daily fluctuations of pCO2 and CO2 flux in a river draining a rapidly urbanizing area. Ecol. Indic. 109, 105821 (2020).Article 
    CAS 

    Google Scholar 
    Cole, J. J. & Caraco, N. F. Carbon in catchments: connecting terrestrial carbon losses with aquatic metabolism. Mar. Freshw. Res. 52, 101–110 (2001).Article 
    CAS 

    Google Scholar 
    Cole, J. J. et al. Plumbing the global carbon cycle: integrating inland waters into the terrestrial carbon budget. Ecosystems 10, 172–185 (2007).Article 

    Google Scholar 
    Spafford, L. & Risk, D. Spatiotemporal variability in lake-atmosphere net CO2 exchange in the Littoral zone of an oligotrophic lake. J. Geophys. Res. Biogeosci. 123, 1260–1276 (2018).Article 
    CAS 

    Google Scholar 
    Rõõm, E. et al. Years are not brothers: Two-year comparison of greenhouse gas fluxes in large shallow lake Võrtsjärv, Estonia. J. Hydrol. 519, 1594–1606 (2015).Article 

    Google Scholar 
    Cole, J. J., Pace, M. L., Carpenter, S. R. & Kitchell, J. F. Persistence of net heterotrophy in lakes during nutrient addition and food web manipulations. Limnol. Oceanogr. 45, 1718–1730 (2000).Article 
    ADS 

    Google Scholar 
    Morales-Pineda, M. et al. Daily, biweekly, and seasonal temporal scales of pCO2 variability in two stratified Mediterranean reservoirs. J. Geophys. Res. Biogeosci. 119, 509–520 (2014).Article 
    CAS 

    Google Scholar 
    Sabine, C. L. et al. The oceanic sink for anthropogenic CO2. Science 305, 367–371 (2004).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Lai, C. et al. Control of carbon dioxide exchange fluxes by rainfall and biological carbon pump in karst river–lake systems. Sci. Total Environ. 937, 173486 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Zavoruev, V. V. et al. Daily course of CO2 fluxes in the atmosphere-water system and variable fluorescence of phytoplankton during the open–water period for lake Baikal according to long-term measurements. Dokl. Earth Sci. 479, 507–510 (2018).Article 
    ADS 
    CAS 

    Google Scholar 
    Du, Q. et al. Factors controlling evaporation and the CO2 flux over an open water lake in Southwest of China on multiple Temporal scales. Int. J. Climatol. 38, 4723–4739 (2018).Article 

    Google Scholar 
    Ni, M., Li, S., Luo, J. & Lu, X. CO2 partial pressure and CO2 degassing in the Daning river of the upper Yangtze river. China J. Hydrol. 569, 483–494 (2019).Article 
    ADS 
    CAS 

    Google Scholar 
    Wilson-McNeal, A. et al. Fluctuating seawater pCO2/pH induces opposing interactions with copper toxicity for two intertidal invertebrates. Sci. Total Environ. 748, 141370 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Minor, E. C. & Brinkley, G. Alkalinity, pH, and pCO2 in the Laurentian great lakes: an initial view of seasonal and inter-annual trends. J. Great Lakes Res. 48, 502–511 (2022).Article 
    CAS 

    Google Scholar 
    Schrier-Uijl, A. P. et al. Release of CO2 and CH4 from lakes and drainage ditches in temperate wetlands. Biogeochemistry 102, 265–279 (2011).Article 
    CAS 

    Google Scholar 
    Wu, Y. et al. Ocean acidification enhances the growth rate of larger diatoms. Limnol. Oceanogr. 59, 1027–1034 (2014).Article 
    ADS 
    CAS 

    Google Scholar 
    Duarte, C. M. et al. CO2 emissions from saline lakes: A global estimate of a surprisingly large flux. J. Geophys. Res. Biogeosci. 113, 7 (2008).Article 

    Google Scholar 
    Finlay, K., Leavitt, P. R., Wissel, B. & Prairie, Y. T. Regulation of Spatial and Temporal variability of carbon flux in six hard–water lakes of the Northern great plains. Limnol. Oceanogr. 54, 2553–2564 (2009).Article 
    ADS 
    CAS 

    Google Scholar 
    Carvalho, A. C. O. et al. Phytoplankton strengthen CO2 uptake in the South Atlantic ocean. Prog. Oceanogr. 190, 102476 (2021).Article 

    Google Scholar 
    Li, D. et al. Metal-algae interaction contributes to the water environment heterogeneity in an urbanized river. Ecol. Indic. 139, e108875 (2022).Article 

    Google Scholar 
    Wang, P., Ma, J., Wang, X. & Tan, Q. Rising atmospheric CO2 levels result in an earlier cyanobacterial bloom-maintenance phase with higher algal biomass. Water Res. 185, 116267 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Xu, Z. & Xu, Y. J. Rapid field Estimation of biochemical oxygen demand in a subtropical eutrophic urban lake with chlorophyll a fluorescence. Environ. Monit. Assess. 187, 4171 (2015).Article 
    PubMed 

    Google Scholar 
    Ngochera, M. J. & Bootsma, H. A. Spatial and Temporal dynamics of pCO2 and CO2 flux in tropical lake Malawi. Limnol. Oceanogr. 65, 1594–1607 (2020).Article 
    ADS 
    CAS 

    Google Scholar 
    Weyhenmeyer, G. A. et al. Significant fraction of CO2 emissions from boreal lakes derived from hydrologic inorganic carbon inputs. Nat. Geosci. 8, 933–936 (2015).Article 
    ADS 
    CAS 

    Google Scholar 
    Zagarese, H. E. et al. Patterns of CO2 concentration and inorganic carbon limitation of phytoplankton biomass in agriculturally eutrophic lakes. Water Res. 190, 116715 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    He, H. et al. Lake metabolic processes and their effects on the carbonate weathering CO2 sink: insights from diel variations in the hydrochemistry of a typical karst lake in SW China. Water Res. 222, 118907 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Rantakari, M. & Kortelainen, P. Interannual variation and Climatic regulation of the CO2 emission from large boreal lakes. Glob. Change Biol. 11, 1368–1380 (2005).Article 
    ADS 

    Google Scholar 
    Benassi, R. F. et al. Eutrophication effects on CH4 and CO2 fluxes in a highly urbanized tropical reservoir (Southeast, Brazil). Environ. Sci. Pollut. Res. 28, 42261–42274 (2021).Article 
    CAS 

    Google Scholar 
    Engel, F., Attermeyer, K. & Weyhenmeyer, G. A. A simplified approach to detect a significant carbon dioxide reduction by phytoplankton in lakes and rivers on a regional and global scale. Sci. Nat. 107, 29 (2020).Article 
    CAS 

    Google Scholar 
    Zhou, M. et al. Space-for-time substitution leads to carbon emission overestimation in eutrophic lakes. Environ. Res. 219, 115175 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Shi, W., Du, M., Ye, C. & Zhang, Q. Divergent effects of hydrological alteration and nutrient addition on greenhouse gas emissions in the water level fluctuation zone of the three Gorges Reservoir, China. Water Res. 201, 117308 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Kortelainen, P. et al. Sediment respiration and lake trophic state are important predictors of large CO2 evasion from small boreal lakes. Glob. Change Biol. 12, 1554–1567 (2006).Article 
    ADS 

    Google Scholar 
    Loken, L. C. et al. Large spatial and temporal variability of carbon dioxide and methane in a eutrophic lake. J. Geophys. Res. Biogeosci. 124, 2248–2266 (2019).Article 
    ADS 
    CAS 

    Google Scholar 
    Natchimuthu, S., Sundgren, I., Gålfalk, M., Klemedtsson, L. & Bastviken, D. Spatiotemporal variability of lake pCO2 and CO2 fluxes in a hemiboreal catchment. J. Geophys. Res. Biogeo. 122, 30–49 (2017).Article 
    CAS 

    Google Scholar 
    Gu, B., Schelske, C. L. & Coveney, M. F. Low carbon dioxide partial pressure in a productive subtropical lake. Aquat. Sci. 73, 317–330 (2011).Article 
    CAS 

    Google Scholar 
    Reis, P. C. J. & Barbosa, F. A. R. Diurnal sampling reveals significant variation in CO2 emission from a tropical productive lake. Braz. J. Biol. 74, 113–119 (2014).Article 

    Google Scholar 
    Reiman, J. H. & Xu, Y. J. Daytime variability of pCO2 and CO2 outgassing fromthe lower Mississippi river: implications for riverine CO2 outgassing Estimation. Water 11, 43 (2019).Article 
    CAS 

    Google Scholar 
    Liu, S. et al. Spatiotemporal dynamics of constructed wetland landscape patterns during rapid urbanization in Chengdu, China. Land 13, 806 (2024).Article 

    Google Scholar 
    The National Environmental Protection Agency (NEPA) & The Editorial Board of Water and Wastewater Monitoring/Analysis Methods. The Monitoring and Analysis Methods of Water and Wastewater, 4th edn (China Environmental Science, 2002).Cole, J. J. & Caraco, N. F. Atmospheric exchange of carbon dioxide in a low-wind oligotrophic lake measured by the addition of SF6. Limnol. Oceanogr 43, 647–656 (1998).Article 
    ADS 
    CAS 

    Google Scholar 
    Abril, G. et al. Technical note: large overestimation of pCO2 calculated from pH and alkalinity in acidic, organic-rich freshwaters. Biogeosciences 12, 67–78 (2015).Article 
    ADS 

    Google Scholar 
    Cai, W. J. & Wang, Y. The chemistry, flux, and sources of carbon dioxide in the estuarine waters of the Satilla and Altamaha rivers. Ga. Limnol. Oceanogr. 43, 657–668 (1998).Article 
    ADS 
    CAS 

    Google Scholar 
    Weiss, R. F. The solubility of nitrogen, oxygen and argon in water and seawater. Deep Sea Res. Oceanogr. Abstr. 17, 721–735 (1970).Article 
    ADS 
    CAS 

    Google Scholar 
    Jahne, B., Heinz, G. & Dietrich, W. Measurement of the diffusion coefficients of sparingly soluble gases in water. J. Geophys. Res. Oceans. 92, 10767–10776 (1987).Article 
    ADS 

    Google Scholar 
    Guérin, F. et al. Gas transfer velocities of CO2 and CH4 in a tropical reservoir and its river downstream. J. Mar. Syst. 66, 161–172 (2007).Article 

    Google Scholar 
    Download referencesAcknowledgementsWe sincerely thank all of assistants at Sichuan Agricultural University for their assistance, and Dr. Yijun Xu at Louisiana State University, LA, U.S.A., for their guidance in early experimental design. Sincere thanks to all of institutions for their supports in testing, and to the Management Center of Bailuwan Lake for their permission and facilitation.FundingThis work was partially supported by the Hai-Ju Program for the Introduction of High-end Talents in Sichuan Provincial Science and Technology Programs (Grant no., 2024JDHJ0017), the Project Supported by Sichuan Landscape and Recreation Research Center (Grant no., JGYQ2024011), the Provincial Innovation Training Program of Sichuan College Students (Grant no., S202510626056), and the Undergraduate Scientific Research Interest Cultivation and Entrepreneurship Training Program Projects at Sichuan Agricultural University (Grant no., 20252046).Author informationAuthors and AffiliationsCollege of Landscape Architecture, Sichuan Agricultural University, Chengdu, 611130, ChinaShiliang Liu, Yingying Chen, Rongjie Yang, Yuling Qiu, Aamir Mehmood Shah, Kezhu Lu, Xinyu Wang, Di Li, Xinhao Cao & Qibing ChenSchool of Tourism and Culture Industry, Chengdu University, Chengdu, 610106, ChinaRongjie YangGeophysical Exploration Brigade, Hubei Geological Bureau, Wuhan, 430100, ChinaDi LiKey Laboratory of Forest and Wetland Conservation in Sichuan Province, Sichuan Academy of Forestry, Chengdu, 610081, ChinaWenbao MaAuthorsShiliang LiuView author publicationsSearch author on:PubMed Google ScholarYingying ChenView author publicationsSearch author on:PubMed Google ScholarRongjie YangView author publicationsSearch author on:PubMed Google ScholarYuling QiuView author publicationsSearch author on:PubMed Google ScholarAamir Mehmood ShahView author publicationsSearch author on:PubMed Google ScholarKezhu LuView author publicationsSearch author on:PubMed Google ScholarXinyu WangView author publicationsSearch author on:PubMed Google ScholarDi LiView author publicationsSearch author on:PubMed Google ScholarWenbao MaView author publicationsSearch author on:PubMed Google ScholarXinhao CaoView author publicationsSearch author on:PubMed Google ScholarQibing ChenView author publicationsSearch author on:PubMed Google ScholarContributionsS.L.: Conceptualization, data curation, formal analysis, funding acquisition, investigation, methodology, resources, software, validation, visualization, writing–original draft, writing–review and editing. Y.C.: Data curation, formal analysis, investigation, software, validation, visualization. R.Y.: Conceptualization, data curation, formal analysis, investigation. Y.Q.: Data curation, formal analysis, investigation. A.M.S.: Writing-review and editing. K.L.: Data curation, Formal analysis, investigation. X.W.: Data curation. D.L.: Data curation, investigation. W.M.: Data curation, writing-review and editing. X.C.: Data curation. Q.C.: Conceptualization, funding acquisition, project administration, resources, supervision, validation.Corresponding authorsCorrespondence to
    Shiliang Liu or Qibing Chen.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary InformationBelow is the link to the electronic supplementary material.Supplementary Material 1Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleLiu, S., Chen, Y., Yang, R. et al. High-frequency monitoring reveals a CO2 source-sink shift in a subtropical eutrophic urban lake.
    Sci Rep 15, 43212 (2025). https://doi.org/10.1038/s41598-025-27331-zDownload citationReceived: 25 July 2025Accepted: 03 November 2025Published: 05 December 2025Version of record: 05 December 2025DOI: https://doi.org/10.1038/s41598-025-27331-zShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative
    KeywordsCO2 source‒sink
    pCO2/fCO2
    Eutrophic urban lakeWater–air CO2 exchangeHigh-frequency monitoring More

  • in

    Combined RP-HILIC for suspect screening of persistent, mobile, and toxic substances in surface water: A case study

    AbstractPersistent, Mobile, and Toxic (PMT) and very Persistent and very Mobile (vPvM) substances pose a significant environmental threat due to their widespread distribution, persistence, and potential toxicity. While previous studies have used reversed-phase (RP) and hydrophilic interaction liquid chromatography (HILIC) to extend analyte coverage, this study uniquely integrates a dual-chromatographic approach with multi-sorbent solid phase etration (SPE) and machine-learning-based retention time indices to rigorously minimize false positives and enhance confidence in suspect screening of PMT/vPvM substances. Two independently optimized LC methods (RP and HILIC) were applied in parallel to the same SPE extract to identify potential chromatographic blind spots. We applied this robust framework to systematically screen surface water and wastewater treatment plant (WWTP) effluent across a water catchment in Uppsala, Sweden. The RP and HILIC workflows, applied as independent chromatographic runs, identified 84 compounds after multi-stage filtering (with retention time indices (RTI) correction applied only to the RP data), with 27% and 48% uniquely detected by HILIC and RP, respectively, highlighting analytical blind spots overcome by this integrated approach. Strategic site selection, encompassing upstream rural zones, on-site sewage facilities, and major WWTP discharge points, demonstrated how local land use impacts PMT/vPvM profiles. This comprehensive method provides a powerful tool for environmental monitoring and regulatory surveillance of PMT/vPvM compounds.

    Similar content being viewed by others

    Contamination of private water supplies after a well communication event (frac-out) in southwest Pennsylvania

    Article
    Open access
    17 September 2025

    Development and validation of a green/blue UHPLC-MS/MS method for trace pharmaceutical monitoring

    Article
    Open access
    22 August 2025

    Surface waters in socially vulnerable areas are disproportionately under-monitored for nutrients in the US South Atlantic-Gulf region

    Article

    15 July 2025

    IntroductionPersistent, Mobile, and Toxic (PMT) and very Persistent, very Mobile (vPvM) substances have emerged as significant environmental contaminants due to their ability to persist and travel long distances in aquatic systems1,2. These substances, characterized by their resistance to degradation and polar nature, present substantial challenges for environmental monitoring and risk assessment. Their high mobility in water systems and limited removal during wastewater treatment make them a concern, as they can accumulate in surface and groundwater, posing risks to ecosystems and human health3.Unlike Persistent, Bioaccumulative, and Toxic (PBT) substances, PMT/vPvM compounds do not accumulate in sediments or biota, making their environmental behavior and risks distinct4. Their mobility and persistence in water bodies necessitate dedicated identification strategies to prevent long-term contamination of drinking water sources.Wastewater treatment plants (WWTPs) and on-site sewage facilities (OSSFs) have been reported to influence the distribution and behaviour of contaminants such as pharmaceuticals, personal care products, pesticides, and per- and polyfluoroalkyl substances (PFASs) in aquatic environments5. While WWTPs typically maintain stable contaminant profiles, OSSFs display greater variability in micropollutant release, highlighting their diverse effects on water quality5. Recognizing the environmental significance of PMT/vPvM substances, the European Commission has emphasized the need for systematic investigations to develop robust methodologies for their detection and management6.Advancements in analytical chemistry, particularly suspect screening methodologies combined with high-resolution mass spectrometry (HRMS), have revolutionized the detection of emerging contaminants7. Unlike targeted analyses, which require prior knowledge of analytes, suspect screening enables the identification of unknown or poorly characterized compounds. This approach uses suspect lists derived from regulatory databases, market data, and environmental studies to screen for potential contaminants8.Detecting and identifying PMT/vPvM substances remain challenging due to their polar and hydrophilic characteristics, which often fall outside the scope of conventional analytical techniques6. Reverse-phase liquid chromatography (RP-LC), commonly used in environmental analysis, is optimized for non-polar or moderately polar compounds but performs poorly with highly polar analytes2. Reemtsma and Jekel (2006) provide a foundational overview of polar organic pollutants in water, highlighting analytical challenges and the need for polarity-extended methods in environmental screening9. This limitation leaves many PMT/vPvM substances undetected, creating an analytical gap10. Hydrophilic interaction liquid chromatography (HILIC) offers a promising alternative for such analytes, providing improved retention and separation of polar compounds by leveraging interactions with their stationary phases11. When combined, HILIC and RP-LC can serve as complementary techniques, providing a broader analytical scope for studying complex environmental matrices12,13. However, most previous studies have applied these techniques separately or in complementary workflows, which limits the ability to systematically quantify the analytical blind spots of each method.Serial RP–HILIC chromatography, often referred to as polarity-extended chromatography, has been extensively developed and utilized to broaden polarity coverage in single-run environmental screening and omics analyses14,15. These studies have demonstrated the method’s capability to comprehensively detect analytes ranging from highly polar to non-polar compounds. For example, Haggarty et al. (2015) demonstrated that serially coupled RP-HILIC with simultaneous gradients enhances metabolomic coverage by enabling the separation of non-polar, polar, and highly polar compounds in one workflow16.Although RP-HILIC methods offer greater flexibility through independent optimization of retention mechanisms and often provide superior resolution for structurally diverse compounds, Montes et al. (2022) recommended combining mixed-mode LC (MMLC) and RP chromatographic modes to maximize analyte coverage—particularly when supercritical fluid chromatography (SFC) is unavailable and HILIC is not included. Their suspect screening approach, applied to surface and wastewater samples across Northern Portugal and Galicia, led to the tentative identification of 343 contaminants of emerging concern (CECs), including 153 classified as persistent, mobile, and toxic (PMT) and 23 as very persistent and very mobile (vPvM), highlighting the effectiveness of this dual-mode strategy for broad environmental monitoring17. However, previous studies did not apply RP and HILIC in parallel, limiting their ability to directly assess the selectivity gaps or retention limitations of each method.In this study, two independently optimized LC methods (RP and HILIC) were applied in parallel to the same solid phase extraction (SPE) extract to quantify chromatographic blind spots. By combining this dual-mode approach with multi-sorbent SPE, high-resolution mass spectrometry (HRMS), and machine-learning-based retention time indices (RTI), we enhance compound confirmation and minimize false positives. This comprehensive strategy addresses critical analytical gaps in PMT/vPvM monitoring and supports regulatory efforts to improve surveillance of these substances. Additionally, our sampling design—covering upstream rural zones, small-scale OSSFs, and large-scale WWTP discharge points—provides insights into how land use influences PMT/vPvM profiles, offering transferable knowledge for other mixed-use catchments across Europe and globally. This workflow enables environmental laboratories to benchmark and reduce RP–HILIC blind spots in suspect screening.Materials and methodsSampling and sample PreparationSurface water samples (n = 8) were collected from diverse locations along the Fyris River catchment and Lake Ekoln, as well as WWTP effluent at Uppsala, Sweden, including surface water and wastewater systems, representing urban, industrial, and rural influences (Fig. 1, Figure S1, Table S1 in Supporting Information (SI1)). These sites were selected to capture a range of contamination profiles influenced by urban, industrial, agricultural, and semi-natural land uses, providing a representative cross-section of typical European catchments impacted by full-scale WWTPs, small-scale OSSFs and diffuse sources.Fig. 1Sampling locations across surface water and wastewater systems impacted by wastewater treatment plants (WWTP) and on-site sewage facilities (OSSFs) in Uppsala, Sweden.Full size imageSamples (1 L) were collected in triplicate in pre-cleaned high-density polyethylene (HDPE) bottles, filtered using 0.7 μm glass fiber filters (Whatman GF/F) to remove particulate matter, and stored at 4 °C in pre-cleaned polypropylene containers to minimize adsorption and degradation of polar and hydrophilic PMT/vPvM substances18. Procedural blanks and field blanks were prepared to monitor potential contamination during sample handling and processing. Each sample, procedural blank, and pooled quality control (QC) was spiked prior to extraction with 50 µL of a mixed isotopically-labelled PFAS internal-standard solution (0.05 µg/mL, Table S2 in SI1). In this study the IS served as process and QC markers for ESI(−) (RP and HILIC) and used to anchor retention times (RT), monitor within-/cross-batch drift, and check carryover.Solid-phase extraction (SPE) was carried out for analyte enrichment following a previously described protocol8. In brief, SPE was performed on 500 mL of each sample in triplicate using four different SPE materials: Oasis HLB, Strata-X-AW, and Strata-X-CV (Phenomenex, Torrance, USA) and Isolute ENV+ (Biotage, Ystrad Mynach, UK). These sorbents were packed into 6 mL polypropylene SPE tubes fitted with 20 μm frits to create an in-house cartridge configuration, allowing simultaneous extraction of a broad range of polar and non-polar compounds. Samples were then eluted with 4 mL of methanol/ethyl acetate (50:50, v/v) containing 2% ammonia, selected to promote desorption of both acidic and basic compounds from the sorbents and enhance recovery across a broad polarity range. The eluates were evaporated under a gentle nitrogen stream and reconstituted in methanol/ethyl acetate (50:50, v/v) containing 1.7% formic acid to stabilize analytes and improve ionization efficiency during LC-HRMS analysis. Importantly, the same extraction protocol was applied for both RP and HILIC to ensure comparability between the two separation techniques. Using different extraction methods would compromise the ability to directly compare their performance. The reconstituted extracts were filtered through 0.45 μm syringe filters and stored at − 20 °C prior to instrumental analysis.Instrumental analysisLiquid chromatography analysis was performed using a Vanquish Horizon ultra-performance liquid chromatography (UPLC) system (Thermo Fisher Scientific, Bremen, Germany). Two complementary chromatographic methods (i.e. RP and HILIC) were employed to provide a comprehensive analytical scope.For RP-LC, a Waters CORTECS C18 + column (90Å, 2.7 μm, 2.1 mm × 100 mm) was used to target non-polar and moderately polar compounds. The mobile phases consisted of water with 0.1% formic acid as solvent A and methanol with 0.1% formic acid as solvent B in positive ionization mode. On the other hand, water with 5 mM ammonium acetate (solvent A) and methanol with 5 mM ammonium acetate (solvent B) were used in negative ionization mode. A gradient elution method was applied, starting at 10% solvent B, ramping to 90% over 16 min, holding for 2 min, and returning to 10% for re-equilibration within a 20 min runtime. The flow rate was maintained at 0.3 mL/min, with an injection volume of 10 µL.For HILIC, a Waters CORTECS HILIC column (2.7 μm, 3 mm × 100 mm) was employed to target polar compounds. This high-efficiency, solid-core column operates via hydrophilic interactions, primarily partitioning and hydrogen bonding, offering orthogonal selectivity compared to RP and enhanced retention and sensitivity for highly polar analytes. The mobile phases consisted of water: acetonitrile (95:5) with 5 mM ammonium formate as solvent A, and acetonitrile-water (95:5) with 5 mM ammonium formate as solvent B were used for analyzing samples in both positive and negative ionization modes. A gradient method started at 100% solvent B, decreased to 5% over 20 min, held for 5 min, and returned to 100% over a total runtime of 30 min. The flow rate was set to 0.7 mL/min with an injection volume of 10 µL.Mass spectrometric analysis was conducted using the QExactive Focus Orbitrap mass spectrometer equipped with a heated electrospray ionization (HESI-II) source. The system operated in both positive and negative ionization modes to maximize the coverage of analytes. Full MS scans were performed across a mass range of 80–1000 m/z for RP-LC and 100–1000 m/z for HILIC at a resolution of 70,000 (at m/z 200). Data-dependent acquisition (DDA) mode was used to collect MS/MS spectra for the most intense ions, employing stepped collision energies of 10, 20, and 40 eV. The MS/MS spectra were acquired at a resolution of 17,500. The ion source settings were optimized for sensitivity and included a sheath gas flow rate of 35 a.u., auxiliary gas flow rate of 10 a.u., spray voltage of 3.0 kV, capillary temperature of 350 °C, and auxiliary gas heater temperature of 300 °C.Data processing and identification workflowSuspect list, compound discoverer workflow, and data analysisThis study focused on a list of 318 prioritized PMT/vPvM substances from the NORMAN Network dataset (NORMAN-SLE-S36.0.3.0, updated 2022) registered under REACH19. These substances were classified into five main categories: 206 industrial chemicals, 42 pharmaceuticals, 39 pesticides, 23 PFAS, and 8 personal care products. The distribution of compound classes reflects the composition of this curated list rather than general environmental prevalence. All information regarding the compounds’ SMILES, molecular formula, molecular weights, categories, applications, and their physicochemical and toxicological data are shown in Table S1 in SI2. Toxicological data, including acute toxicity thresholds for fish, daphnids, mysids, green algae, and log n-octanol-water partition coefficient (Kow) and water solubility properties, were calculated using the Ecological Structure-Activity Relationships (ECOSAR) program 2.2.20.Data processing was performed using Compound Discoverer 3.3 (Thermo Fisher Scientific), with the specific steps of the workflow summarized in Table S3 in SI1. In brief, this workflow facilitated the detection, annotation, and comparison of unknown compounds across samples. RT alignment was performed separately for each chromatographic mode (RP and HILIC) using a conservative ± 2.0 min cross-batch matching window with ± 5 ppm mass tolerance. These parameters were chosen after following QC evaluation: internal standards (IS), RTI calibrants, and system-suitability test (SST) injections showed tight within-batch RT reproducibility but small, systematic cross-batch/matrix drifts. Thus a narrower RT filter risked excluding true features across batches. RT was therefore used only as a prefilter and final feature annotations additionally required accurate mass, consistent isotopic patterns, and MS/MS spectral agreement. Low signal(S)/noice (N) or aberrant isotope-pattern features were removed. We also applied an RTI workflow to normalize sample-to-sample RT differences and to imporve candidate ranking. In ESI(−), PFAS IS acted as RT anchors and drift checks prior to spectral matching; in ESI(+), alignment relied on pooled-QC/SST injections, blanks, and RTI calibrants. Accordingly, positive-mode signals were reported as qualitative or relative unless confirmed by MS/MS data.Suspect screening was performed to identify PMT compounds from the prioritized list, using multiple databases including mzCloud, mzVault, ChemSpider, ChEMBL, ECHA, EPA DSSTox, EPA ToxCast, MassBank, and PubMed. Compound annotation was supported by the mzLogic algorithm, which ranks candidates based on structural similarity. Background signals were subtracted, and gap filling was used. We did not apply log D filters for HILIC separation; instead log D (pH 7) was used only as a plausibility check, and confidence was based on ± 5 ppm mass error tolerance, mode-specific RT windows (RP 0–20 min; HILIC 0–35 min), isotopic pattern matching, MS² library/in-silico matching, and RTI for RP, and reported using Schymanski levels.Integration of RP-LC and HILIC data was central to this approach, allowing cross-validation of compounds identified by both methods to enhance detection confidence. Additionally, the unique detections from each method highlighted their complementary roles, addressing analytical blind spots and ensuring comprehensive coverage of PMT/vPvM substances.Retention time indices (RTI)The RTI for LC-HRMS (version 2.5.0), developed by the National and Kapodistrian University of Athens, Greece, was utilized to enhance the reliability and accuracy of compound identification by minimizing residual errors between experimental and predicted retention time (tR) via removing false positive features21. The advanced machine learning algorithm OTrAMS (Ordered Traits and Machine Learning System) model was employed to address this22. Calibration was achieved using RTI calibrants, encompassing four mixtures of 18 compounds each, tested under + ESI and -ESI modes (for details, see Figures S2 and S3, Tables S4 and S5 in SI1). The calibration method employs the RTI formula as shown in the Eqs. 1 and 2.$$RTI = :{frac{{tR}_{x}-{tR}_{min}}{{tR}_{max-}{tR}_{min}}}_{}x:1000:$$
    (1)
    $$RTI = α :left({tR}_{c}right)+C$$
    (2)
    where, α is defined as calibration coefficient, (:{tR}_{x}) and (::{tR}_{c}) as RT observed for the calibrants and a compound, respectively, and (:{tR}_{min}) and (:{tR}_{max}) are the minimum and maximum tR observed for the calibrants, respectively. This standardized approach confirms stability across analytical runs in the RP separation technique, enhances confidence in tR predictions, and supports more robust compound annotation in our study. Based on the RTI workflow, compounds are categorized into four confidence levels. The highest tR matches are classified from Level 1 to Level 4 (box 1–4, described at Table S6 in SI1), while Level 4 is considered as false-positive. Compounds at Level 2 are also acceptable, though with lower confidence than Level 1 (Table S6 in SI1).For compound identification, the in silico fragmentation tool in Compound Discoverer 3.3, FISh (Fragment Ion Search), was utilized, along with matching the scores within different mass spectral libraries. To further enhance confidence, all MS² spectra were verified using CFM-ID 4.0 (Competitive Fragment Modeling Identification)23 and the open mass library MassBank (https://www.massbank.eu, Last Updated version on 2024.11)24, ensuring precise structural annotation and improved reliability in compound identification. For the compound confidence level assessment, we adapted criteria from the Schymanski classification25. Accordingly, Level 1 compounds were confirmed with reference standards (RT and MS²). When MS² data were unavailable, features were assigned to Level 5 (accurate mass only), Level 4 (accurate mass and an unequivocal molecular formula), or Level 3.3 (accurate mass and molecular formula plus a match to a plausible candidate in mass/suspect libraries). When MS² data were available, stricter criteria were applied: Level 3.2 (low-quality MS²; low matching scores in Compound Discoverer; FISh < 50, less diagnostic fragments), Level 3.1 (higher CD matching scores, FISh ≥ 50, more diagnostic fragments, but not fully concordant with MassBank and the in-silico fragmentation tool CFM-ID), Level 2.2 (high-similarity library MS² scores in CD with full agreement from MassBank and CFM-ID, but isomeric uncertainty remains), and Level 2.1 (high-similarity library MS² with MassBank/CFM-ID support and no isomeric alternatives).Statistical analysisSpatial similarity among sites were quantified by computing pairwise Spearman correlations (ρ) on log 10-transformed intensities (RP and HILIC data collapsed per compound as the site-wise maximum). The resulting correlation matrix was visualized with hierarchical clustering (average linkage). For community-level composition, non-metric multidimensional scaling (NMDS) was applied using Bray–Curtis on Hellinger-transformed data. The statistical analysis were performed using RStudio (V 2024.12.0 + 467), with the corresponding R script provided (Table S2 in SI2).Results and discussionIdentification of compounds using RP and HILICThe dataset reveals the distribution and refinement of chromatographic features in RP and HILIC techniques in both + ESI (POS) and -ESI (NEG) ionization modes (Tables S7 in SI1 and S3─S5 in SI2). Figure 2 and Table S7 (SI1) demonstrate the stepwise reduction of features through quality assurance (QA)/QC filtering and the subsequent pre-annotation prioritization based on suspect and library matching. Applying filtering, including background exclusion, peak rating, mass tolerance, and RT detection windows, resulted in a significant reduction of features to 355 (RP-POS), 187 (RP-NEG), 285 (HILIC-POS), and 32 (HILIC-NEG) (Fig. 2). Subsequent application of criteria selecting features with the highest match scores (mzCloud, mzVault, and FISh) and eliminating false positives (Fig. 2, Table S7 in SI1) reduced the features finally to 58 (RP-POS), 35 (RP-NEG), 40 (HILIC-POS), and 7 (HILIC-NEG) (Fig. 2). All features detected as having MS2 were checked through MassBank and CFM-ID 4.023. Finally, incorporating the RTI filter improved specificity beyond fixed RT windows. In RP mode, RTI removed 33 of 93 pre-RTI candidates (− 35%), predominantly at tentative levels (Level 3.2 reduced 41→28; Level 3.3 reduced 32→17), while preserving all compounds at Level 1 and 2 (Tables S3 and S4 in SI2). Of the 60 RP assignments retained after RTI, 35 fell within the strict RTI acceptance domain (box 1, described at Table S6 in SI1) and 25 were retained as box 2 (described at Table S6 in SI1) due to consistent MS/MS and mass accuracy despite modest RT deviations (Table S4 in SI2).Fig. 2Number of features treated using (A) RP and (B) HILIC application. For details see Tables S8-S10 in SI.Full size imageThe separate applications of HILIC and RP provided complementary insights into compound identification by leveraging the distinct selectivity of each chromatographic mode. In total, 84 unique compounds were determined after combining both ionization modes within each chromatographic technique. HILIC identified 23 unique polar compounds (Table S6 in SI2), typically challenging for RP to retain, demonstrating its strength in analyzing hydrophilic analytes with functional groups such as hydroxyl and amine. Examples include dimorpholinodiethyl ether, DMH (5,5-dimethylhydantoin), hexamethyldisiloxane, losartan, methylpiperazine, m-phenylenediamine, pymetrozine, skatole, sulfadimidine, and triflumezopyrim. In contrast, RP identified 40 unique compounds (Table S4 in SI2), such as oxcarbazepine, perfluorohexanoic acid, and galaxolide, which are generally more hydrophobic and exhibit higher logP values. Both methods confirmed 21 overlapping compounds (Table S5 in SI2), including erythromycin, 1,2,3-benzotriazole, 2,6-xylidine, acid red 337, azobisisobutylonitrile, candesartan, ibuprofen, imazapyr, and isophorone diisocyanate. These overlapping compounds typically possess intermediate polarity or amphiphilic properties, allowing retention in both chromatographic systems. These findings reinforce the complementary nature of RP and HILIC and support their parallel use to achieve broader chemical coverage and reduce analytical blind spots, further enhanced by spectral tools and RTI for greater confidence. To quantify polarity differences, log P values were compiled for the identified compounds (dimensionless). Log P values were available for 77 compounds (8 entries were salts or dyes and lacked values), and were summarized by detection category. HILIC-only identifications (n = 22 with log P) had a median log P of 1.7 (IQR: −0.15–3.00), whereas RP-only identifications (n = 35) had a median log P of 2.6 (IQR: 1.05–3.50). Compounds detected by both techniques (n = 20) showed intermediate polarity with a median log P of 1.4 (IQR: 0.70–2.15). On average, the RP-only set was ~ 0.9 log P units more hydrophobic than the HILIC-only set, supporting the expected and complementary selectivity of the two chromatographic modes. For transparency, a “log P” column has been added to Table S6 in SI2.Unlike previous studies that applied RP or HILIC separately, we used both in parallel as independent runs to directly quantify each method’s analytical blind spots. This highlights the critical need for complementary workflows to avoid underestimating PMT/vPvM substances. For example, Schulze et al. (2019) combined multiple chromatographic (MMLC, HILIC, RP, and SFC) with various SPE techniques to detect 43 of 64 targeted persistent and mobile organic chemicals (PMOCs), but their methods were complementary rather than parallel, limiting direct assessment of method-specific limitations26. Similarly, Castro et al. (2021) showed that combining different sampling (POCIS, SPE) and chromatographic modes (RP, MMLC) significantly improves detection coverage, identifying 343 compounds, including 153 PMT and 23 vPvM substances4. However, their workflow also lacked a systematic quantification of blind spots for each chromatographic mode. In contrast, our parallel RP-HILIC strategy enables a more rigorous evaluation of selectivity, offering clearer insights for improving analytical coverage and reliability in PMT/vPvM monitoring. Because our suspect list is intentionally enriched for industrial chemicals (206 of 318 entries; Table S1 in SI2), raw across-class frequencies should not be interpreted as indicators of environmental prevalence. Accordingly, we focus on site-to-site contrasts and the complementary coverage provided by parallel RP–HILIC rather than absolute cross-class comparisons.Identified compounds using RPLevel 1Level 1–2 identifications are highlighted in the main text, whereas Levels 3–5 are provided in the SI1 with corresponding evidence. Of the 93 identified compounds using the RP technique (Tables S3 and S4 in SI2), metformin (C4H11N5, m/z 130.1086 [M + H]+1) was identified at Level 1 by confirming with the reference standard and MS² match. This antidiabetic exhibited persistence in aquatic environments, highlighting its prevalence in WWTP and the aquatic environment27. Its short tR (0.78 min) using RP agreed with the RTI prediction, demonstrating effective capture of highly polar analytes in RP (Figure S4a in SI1).Level 2At Level 2, 5 compounds were identified. Two Level 2.1 compounds—isophorone diisocyanate (IPDI) and piperazine—were supported by concordant RTI and MS² (FISh/CFM-ID; Figure S4b–c in SI1). IPDI, (C12H18N2O2, m/z 223.1439 [M + H]+1), used in polyurethane production28 and classified as a respiratory sensitizer by the U.S. Environmental Protection Agency (EPA)29, was detected in both RP and HILIC modes, with higher intensity in RP and maximal intensity in wastewater effluent (Figure S4b in SI1), indicating incomplete removal during treatment. Piperazine (C4H10N2, m/z 87.09155 [M + H]+1), an industrial and pharmaceutical intermediate with known antinematodal activity30 showed diagnostic fragments (m/z 70.065, 56.049) with RTI agreement. Three Level 2.2 compounds, N,N′-diphenylguanidine (a rubber accelerator previously detected in industrial wastewater31, N-2-ethylhexyl bicycloheptenedicarboximide (an insect repellent32, and N-butyl benzenesulfonamide (a plasticizer frequently reported in aquatic environments33, had high MS² matches with acceptable RTI (box 2, described at Table S6 in SI1) (Figure S4d–f in SI1).Level 3At Level 3.1 (n = 2), allyl 2,3-epoxypropyl ether (C6H10O2, m/z 115.0752 [M + H]+) and azithromycin (C38H72N2O12, m/z 375.2612 [M + H]+) exceeded FISh thresholds and showed strong agreement with both spectral library and in-silico matches, but structural ambiguity prevented higher confidence assignment. Allyl 2,3-epoxypropyl ether is a toxic and potentially carcinogenic epoxy ether34, while azithromycin—frequently detected in hospital effluents—indicates environmental persistence and potential ecotoxicity35,36. Level 3.2 (n = 28) comprised tentatives with lower FISh/library scores, yet acceptable experimental/predicted tR alignment (RTI Level 1) or within the applicability domain (RTI Level 2). Representative pharmaceuticals include candesartan (C24H20N6O3), climbazole (C15H17ClN2O2), diclofenac sodium (C14H11Cl2NO2), ibuprofen (C13H18O2), phenazone (C11H12N2O), phenytoin (C15H12N2O2), and iohexol (C19H26I3N3O9). Level 3.3 (n = 17) contained lower-confidence tentatives such as erythromycin (C37H67NO13), atrazine (C8H14ClN5), and diuron (C9H10Cl2N2O), all widely reported in aquatic environments37, and nonetheless captured by the combined RP–HILIC workflow, underscoring its value for broad contaminant screening.Levels 4 and 5Two compounds, including 4-(1-phenylmethyl)-1,3-benzenediol (C14H14O2) and 1-ethyl-2,3,3-trimethyl-3 H-indolium iodide (C13H19N), were identified at Level 4, as both cases matched the molecular weight and exact mass, and were validated using the RTI method. Lastly, Level 5 included five compounds characterized by their precise mass.
    Identified compounds using the HILIC separation technique
    Level 1Among 47 HILIC identifications (Table S5 in SI2), metformin was confirmed at Level 1 using a reference standard (Figure S5a in SI1). Notably, HILIC gave ~ 16× higher metformin intensity than RP, underscoring HILIC’s suitability for identification/quantification of this compound across matrices38.Level 2Three compounds, 5,5-dimethylhydantoin (DMH, C5H8N2O2, m/z 129.06581 [M + H]+1), (C12H24N2O3, m/z 245.18612 [M + H]+1), and 4-aminophenol (C6H7NO, m/z 110.0601 [M + H]+1), were assigned Level 2.1 based on top-ranked MS2 matches (libraries, FISh, CFM-ID) (Figure S5b–d in SI1). DMH, used as a preservative/disinfectant and pharmaceutical intermediate, has generally low reported toxicity but warrants qualitative dietary risk assessment (inclusing its degradate EMH) and can originate from industrial/biocidal uses or DBDMH hydrolysis39,40,41. Dimorpholinodiethyl ether, an industrial blowing agent associated with dermatotoxicity42, showed fragments m/z 158.11757 (C8H16NO2+), m/z 114.09135 (C6H12NO+), and m/z 96.08080 (C6H10N+)19 (Figure S5c in SI1). 4-Aminophenol, a toxic paracetamol by-product with reported aquatic releases, exhibited a characteristic MS2 spectrum showing fragmentations corresponding to the loss of hydroxyl group (C6H6N+, m/z 92.04947), and the opened anilin moiety (C5H8N+, m/z 82.06516), and 67.04165 (C4H5N+) (Figure S5d in SI1).Level 3Among 36 tentative HILIC identified compounds, skatole (C9H9N, RT: 8.15 min, m/z 132.08086 [M + H]+1) was the only Level 3.1 compound: Its MS2 spectrum matched in-silico predictions and libraries, but structural ambiguity with a near isomer prevented higher confidence. Skatole is widely used as a flavor/fragrance and is frequently reported in surface waters43. Most Level 3.2 compounds were pharmaceuticals, including ibuprofen (C13H18O2, m/z 205.12369 [M-H][- [1), iohexol (C19H26I3N3O9, m/z 821.8881 [M + H]+1), losartan (C22H23ClN6O, m/z 423.16956 [M + H]+1), phenazone (C11H12N2O, m/z 189.10229 [M + H]+1), phenytoin (C15H12N2O, m/z 253.0974 [M + H]+1), all previously observed in aquatic environments44. An additional 14 suspects were assigned at Level 3.3 based on accurate mass/formula and library matches but lacked confirmatory MS2 evidence.Levels 4 and 5Two PMT candidates were assigned at Level 4 (isobaric formulas with exact-mass agreement), and three additional contaminants were Level 5 (exact-mass only). These results support integrating dual-mode LC–HRMS (RP + HILIC) into regulatory monitoring to improve detection of PMT/vPvM substances that are often overlooked by conventional single-mode workflows.Spatial distributionThe identified compounds using RP and HILIC separation techniques were divided into 6 different categories, including industrial chemicals, pharmaceuticals, personal care products, pesticides, PFASs, and natural compounds (Fig. 3, Table S6 in SI2). Industrial chemicals account for 55% of all identifications; however, this proportion mirrors the composition of our suspect list (206 of 318 entries classified as industrial; Table S1 in SI2). Consequently, raw class fractions should not be interpreted as indicators of environmental prevalence. Instead, these summaries are used to frame site-to-site differences and highlight detections uniquely enabled by HILIC45. Pharmaceuticals (19%) and personal care products (11%) were frequently observed within the scope of the screened suspects (Table S1 in SI2), emphasizing their resistance to wastewater removal processes46. The occurrence of pesticides (9.5%) (e.g., atrazine, imazapyr, diuron) indicates impacts from agricultural runoff47. The presence of PFAS (4.8%), an emerging group of substances, has been shown in previous studies they are ubiquitously distributed in the aquatic environment due to their high persistence and mobility48. In contrast, natural products remain consistently low (1.2%), suggesting natural background levels.Fig. 3Categories of the identified compounds at the surface water samples (n = 8) and wastewater treatment plant (WWTP) effluent in Uppsala, Sweden.Full size imageThe observed presence and intensities of organic micropollutants varied largely across different sampling locations, influenced primarily by anthropogenic activities, wastewater treatment plant effluents, hydrological conditions, and physicochemical properties of the compounds49. As demonstrated in Figs. 4 and 5, the heatmaps of spatial distribution (cells show within-compound z-scored log₁₀ intensities: green = low, yellow = mid, red = high; gray = not detected) showed the highest intensities predominantly at locations directly impacted by WWTP effluent (site W). This site exhibited elevated contamination across all categories, specifically pharmaceuticals such as erythromycin, ibuprofen, and diclofenac, highlighting the limited removal efficiency of conventional wastewater treatment processes for pharmaceutical residues (Fig. 5)5,50. On the other hand, the reference site 1 was the cleanest location among all sampling sites, due to its location upstream of urban areas and upstream of WWTP effluents, which minimised exposure to contaminants commonly associated with wastewater discharge, industrial emissions, and urban runoff. Similarly, Lake Ekoln (site 7) exhibited lower contamination levels compared to the other sites, likely due to the dilution effect from diffuse sources as previously reported5.Fig. 4The heatmap graph of the spatial distribution of the industrial chemicals identified across diverse sampling locations.Full size imageFig. 5The heatmap graph of the spatial distribution of the pharmaceuticals, personal care products, pesticides, PFASs, and natural compounds identified across diverse sampling locations.Full size imageIn addition to the class-level heatmaps (Figs. 4 and 5), a site-to-site Spearman correlation heatmap with hierarchical clustering (Figure S6 in SI1) quantitatively supports the observed spatial patterns. Pairwise correlations coefficients ranged from approximately 0.4 to 1.0, with high within-cluster similarity (ρ ≈ 0.9–1.0) among several river and lake sites. In contrast, correlations between the WWTP effluent (W) and upstream reference sites were notably lower, for example, the W vs. Site 1 comparison showed the weakest correlation. The clustering dendrogram clearly separates WWTP-impacted locations from upstream or less-affected sites, reflecting the spatial gradient described earlier.This structure is further supported by the NMDS ordination (Figure S7 in SI1), based on Bray–Curtis dissimilarities of Hellinger-transformed, log10-scaled intensities. The ordination reveals a clear separation of W from upstream references, with downstream and urban-influenced sites clustering together, and tributary sites positioned intermediately. The low stress value (reported in the figure) indicates that the two-dimensional solution effectively captures the multivariate relationships. Together, the correlation heatmap and NMDS corroborate the spatial trends inferred from the class-level results.Locations downstream from OSSFs, such as the Husby tributary (site 3) and the Sävja River (site 4), demonstrated relatively high intensities of pharmaceuticals, including olmesartan and phenazone, as well as pesticides, such as propazine. These findings align with previous studies indicating that OSSFs often inadequately treat persistent pharmaceuticals, resulting in elevated contaminant concentrations downstream51.Comparatively, upstream sites such as sites 1 and 8 generally displayed lower intensities across all categories. However, pesticide residues, notably diuron, displayed measurable intensities even at the upstream sites, suggesting agricultural runoff as a consistent diffuse source. This aligns with the understanding that pesticides primarily enter waterways through surface runoff from agricultural activities, underscoring the diffuse nature of agricultural pollution52.Natural products, exemplified by hexadecatrienoic acid, maintained uniform intensities across various sites, reflecting a natural background level independent of wastewater treatment processes. In contrast, the intensities of cosmetic and personal care products were dominant at downstream urban-influenced sites 2 (e.g., oxybenzone and sunscreens) and 6 (e.g., galaxolide and skatol). This pattern correlates with human usage patterns and highlights their anthropogenic origin53.Persistent substances like PFAS (e.g., perfluorohexanoic acid and perfluorobutane sulfonic acid) showed no particular distribution, indicating their ubiquitous presence in the environment due to their persistent and mobile characteristics54. Notably, upstream sites of the military airport Ärna Air Base (sites 1, 2, and 8) showed no significant PFAS contamination55.Overall, the spatial distribution of identified organic micropollutants underscores the critical impact of WWTP discharge, industrial activities, and agricultural runoff on water quality. The significant site-specific variations align with compound properties and proximity to the source. This study demonstrates the critical role of HILIC as a valuable complementary technique to conventional RP separation methods in identifying organic micropollutants. These insights underscore the need for advanced treatment technologies and targeted pollution control measures to mitigate contaminant inputs and their environmental impacts effectively.Conclusion and prospectsThe current study established a comprehensive LC-HESI-HRMS method for the suspect screening of PMT/vPvM substances in surface water and wastewater systems in Uppsala, Sweden. The developed HILIC separation technique, combined with RP, enabled the identification of a broad range of these chemicals, possessing diverse polarities. In total, 84 unique compounds were identified at varying confidence levels, with 27% and 48% identified solely through HILIC and RP methods, respectively. Among the identified substances, industrial chemicals (55% of identified substances, e.g., hexamethyldisiloxane and para-cresidine) and pharmaceuticals (19%, e.g., metformin and tolytriazole) emerged as the primary contaminants, exhibiting significant intensities across the investigated aquatic matrices.This study highlights the importance of using complementary analytical methods, such as RP and HILIC, to address the limitations of single techniques, thereby ensuring a more comprehensive approach to monitoring environmental contaminants. This robust dual-separation framework can serve as a template for catchment-wide PMT/vPvM surveillance under evolving EU Water Framework Directive requirements, for stricter regulations and improved treatment technologies to manage the risks associated with PMT/vPvM substances. Additionally, ongoing environmental monitoring and effective screening methods are crucial for reducing the ecological and human health impacts of emerging aquatic pollutants.

    Data availability

    All data supporting the findings of this study are available within the paper and its Supplementary Information (SI), SI1 (the word file) and SI2 (spreadsheet excel file).
    ReferencesHale, S. E., Arp, H. P. H., Schliebner, I. & Neumann, M. What’s in a name: persistent, mobile, and toxic (PMT) and very persistent and very mobile (vPvM) substances. Environ. Sci. Technol. 54, 14790–14792. https://doi.org/10.1021/acs.est.0c05257 (2020).Article 
    CAS 
    PubMed 
    ADS 

    Google Scholar 
    Hale, S. E., Arp, H. P. H., Schliebner, I. & Neumann, M. Persistent, mobile and toxic (PMT) and very persistent and very mobile (vPvM) substances pose an equivalent level of concern to persistent, bioaccumulative and toxic (PBT) and very persistent and very bioaccumulative (vPvB) substances under REACH. Environ. Sci. Europe. 32 https://doi.org/10.1186/s12302-020-00440-4 (2020).Chirsir, P. et al. Grouping strategies for assessing and managing persistent and mobile substances. Environ. Sci. Eur. 36, 102. https://doi.org/10.1186/s12302-024-00919-4 (2024).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Castro, V. et al. Combination of different chromatographic and sampling modes for high-resolution mass spectrometric screening of organic microcontaminants in water. Anal. Bioanal Chem. 413, 5607–5618. https://doi.org/10.1007/s00216-021-03226-6 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Gago-Ferrero, P., Gros, M., Ahrens, L. & Wiberg, K. Impact of on-site, small and large scale wastewater treatment facilities on levels and fate of pharmaceuticals, personal care products, artificial sweeteners, pesticides, and perfluoroalkyl substances in recipient waters. Sci. Total Environ. 601–602, 1289–1297. https://doi.org/10.1016/j.scitotenv.2017.05.258 (2017).Article 
    CAS 
    PubMed 
    ADS 

    Google Scholar 
    Hale, S. E. et al. Getting in control of persistent, mobile and toxic (PMT) and very persistent and very mobile (vPvM) substances to protect water resources: strategies from diverse perspectives. Environ. Sci. Europe. 34 https://doi.org/10.1186/s12302-022-00604-4 (2022).Hollender, J. et al. NORMAN guidance on suspect and non-target screening in environmental monitoring. Environ. Sci. Europe. 35 https://doi.org/10.1186/s12302-023-00779-4 (2023).Gago-Ferrero, P. et al. Extended suspect and Non-Target strategies to characterize emerging Polar organic contaminants in Raw wastewater with LC-HRMS/MS. Environ. Sci. Technol. 49, 12333–12341. https://doi.org/10.1021/acs.est.5b03454 (2015).Article 
    CAS 
    PubMed 
    ADS 

    Google Scholar 
    Reemtsma, T. & Jekel, M. Organic Pollutants in the Water Cycle: Properties, Occurrence, Analysis and Environmental Relevance of Polar Compounds 1 edn (Wiley-VCH, 2006).Neuwald, I. J. et al. Occurrence, Distribution, and environmental behavior of persistent, mobile, and toxic (PMT) and very persistent and very mobile (vPvM) substances in the sources of German drinking water. Environ. Sci. Technol. 56, 10857–10867. https://doi.org/10.1021/acs.est.2c03659 (2022).Article 
    CAS 
    PubMed 
    ADS 

    Google Scholar 
    Buszewski, B. & Noga, S. Hydrophilic interaction liquid chromatography (HILIC)–a powerful separation technique. Anal. Bioanal Chem. 402, 231–247. https://doi.org/10.1007/s00216-011-5308-5 (2012).Article 
    CAS 
    PubMed 

    Google Scholar 
    Menger, F., Gago-Ferrero, P., Wiberg, K. & Ahrens, L. Wide-scope screening of Polar contaminants of concern in water: A critical review of liquid chromatography-high resolution mass spectrometry-based strategies. Trends Environ. Anal. Chem. 28 https://doi.org/10.1016/j.teac.2020.e00102 (2020).Neuwald, I. et al. Filling the knowledge gap: A suspect screening study for 1310 potentially persistent and mobile chemicals with SFC- and HILIC-HRMS in two German river systems. Water Res. 204, 117645. https://doi.org/10.1016/j.watres.2021.117645 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Bieber, S. et al. Polarity-Extended organic molecule screening in environmental (Water) samples. Anal. Chem. 89, 7907–7914. https://doi.org/10.1021/acs.analchem.7b00859 (2017).Article 
    CAS 
    PubMed 
    ADS 

    Google Scholar 
    Hemmler, D., Heinzmann, S. S., Wohr, K., Schmitt-Kopplin, P. & Witting, M. Tandem HILIC-RP liquid chromatography for increased Polarity coverage in food analysis. Electrophoresis 39, 1645–1653. https://doi.org/10.1002/elps.201800038 (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Haggarty, J. et al. Serially coupling hydrophobic interaction and reversed-phase chromatography with simultaneous gradients provides greater coverage of the metabolome. Metabolomics 11, 1465–1470. https://doi.org/10.1007/s11306-014-0770-7 (2015).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Montes, R. et al. Screening of contaminants of emerging concern in surface water and wastewater Effluents, assisted by the Persistency-Mobility-Toxicity criteria. Molecules 27 https://doi.org/10.3390/molecules27123915 (2022).Knoll, S., Rosch, T. & Huhn, C. Trends in sample Preparation and separation methods for the analysis of very Polar and ionic compounds in environmental water and biota samples. Anal. Bioanal Chem. 412, 6149–6165. https://doi.org/10.1007/s00216-020-02811-5 (2020).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Arp, H. P. H., Hale, S. E., Schliebner, I. & Neumann, M. (Zenodo, (2022).U.S.EPA. Ecological Structure Activity Relationships (ECOSAR) Predictive Model, (2024). https://www.epa.gov/tsca-screening-tools/ecological-structure-activity-relationships-ecosar-predictive-modelAalizadeh, R. et al. Development and application of liquid chromatographic retention time indices in HRMS-Based suspect and nontarget screening. Anal. Chem. 93, 11601–11611. https://doi.org/10.1021/acs.analchem.1c02348 (2021).Article 
    CAS 
    PubMed 
    ADS 

    Google Scholar 
    Aalizadeh, R., Thomaidis, N. S., Bletsou, A. A. & Gago-Ferrero, P. Quantitative Structure-Retention relationship models to support nontarget High-Resolution mass spectrometric screening of emerging contaminants in environmental samples. J. Chem. Inf. Model. 56, 1384–1398. https://doi.org/10.1021/acs.jcim.5b00752 (2016).Article 
    CAS 
    PubMed 

    Google Scholar 
    Wang, F. et al. CFM-ID 4.0: More accurate ESI-MS/MS spectral prediction and compound identification. Anal. Chem. 93, 11692–11700. https://doi.org/10.1021/acs.analchem.1c01465 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    MassBank. (2024).Schymanski, E. L. et al. Identifying small molecules via high resolution mass spectrometry: Communicating confidence. Environ. Sci. Technol. 48, 2097–2098. https://doi.org/10.1021/es5002105 (2014).Article 
    CAS 
    PubMed 
    ADS 

    Google Scholar 
    Schulze, S. et al. Occurrence of emerging persistent and mobile organic contaminants in European water samples. Water Res. 153, 80–90. https://doi.org/10.1016/j.watres.2019.01.008 (2019).Article 
    CAS 
    PubMed 
    ADS 

    Google Scholar 
    Golovko, O. et al. Occurrence and removal of chemicals of emerging concern in wastewater treatment plants and their impact on receiving water systems. Sci. Total Environ. 754, 142122. https://doi.org/10.1016/j.scitotenv.2020.142122 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Dearman, R. J., Spence, L. M. & Kimber, I. Characterization of murine immune responses to allergenic diisocyanates. Toxicol. Appl. Pharmacol. 112, 190–197. https://doi.org/10.1016/0041-008x(92)90187-w (1992).Article 
    CAS 
    PubMed 

    Google Scholar 
    Agency, U. S. E. P. Toxicological review of isocyanates, including isophorone diisocyanate (IPDI) (2015).Ren, K., Liu, S., Dong, Z. & Wang, J. Wet catalytic oxidation of a FeMnCe-activated semi-coke catalyst for treating piperazine wastewater. ACS Omega. 8, 8683–8694. https://doi.org/10.1021/acsomega.2c08074 (2023).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Challis, J. K. et al. Occurrences of tire rubber-derived contaminants in cold-climate urban runoff. Environ. Sci. Technol. Lett. 8, 961–967. https://doi.org/10.1021/acs.estlett.1c00682 (2021).Article 
    CAS 

    Google Scholar 
    USEPA/Office of Prevention & a., P. T. S. Reregistration Eligibility Decision (RED) Document – N-Octyl bicycloheptene dicarboximide (MGK-264), p.15. Report No. EPA 738-R-06-006, (2006).Klanovicz, N. & Pinto, C. A. Occurrence of pharmaceutically active compounds in groundwater and their effects to the human health. Environ. Sci. Pollut Res. Int. 31, 33223–33238. https://doi.org/10.1007/s11356-024-33423-6 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Program, N. T. N. T. P. & Toxicology and Carcinogenesis Studies of Allyl Glycidyl Ether (CAS No. 106-92-3) in Osborne-Mendel Rats and B6C3F1 Mice (Inhalation Studies). National Toxicology Program Technical Report Series 376, 1-219 (1990).Sorengard, M. et al. Mass loads, source apportionment, and risk estimation of organic micropollutants from hospital and municipal wastewater in recipient catchments. Chemosphere 234, 931–941. https://doi.org/10.1016/j.chemosphere.2019.06.041 (2019).Article 
    CAS 
    PubMed 
    ADS 

    Google Scholar 
    Malnes, D. et al. Hazard screening of contaminants of emerging concern (CECs) in sweden’s three largest lakes and their associated rivers. J. Hazard. Mater. 453, 131376. https://doi.org/10.1016/j.jhazmat.2023.131376 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Panthi, S. et al. Pharmaceuticals, herbicides, and disinfectants in agricultural water sources. Environ. Res. 174, 1–8. https://doi.org/10.1016/j.envres.2019.04.011 (2019).Article 
    CAS 
    PubMed 

    Google Scholar 
    Scherf-Clavel, O., Kinzig, M., Stoffel, M. S., Fuhr, U. & Sorgel, F. A HILIC-MS/MS assay for the quantification of Metformin and sitagliptin in human plasma and urine: A tool for studying drug transporter perturbation. J. Pharm. Biomed. Anal. 175, 112754. https://doi.org/10.1016/j.jpba.2019.07.002 (2019).Article 
    CAS 
    PubMed 

    Google Scholar 
    Lee, D. S., Shim, J. W. & Shin, J. W. Development and validation of an analytical method for the determination of 5,5-dimethylhydantoin in various processed food. Food Sci. Biotechnol. https://doi.org/10.1007/s10068-024-01747-1 (2024).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Agency, U. S. E. P. (Washington, DC: U.S. EPA (2020).Jiang, Z., Liu, Y., Li, R., Ren, X. & Huang, T. S. Preparation of antibacterial cellulose with a monochloro-s‐triazine‐based N‐halamine biocide. Polym. Adv. Technol. 27, 460–465. https://doi.org/10.1002/pat.3691 (2015).Article 
    CAS 

    Google Scholar 
    Karimi-Maleh, H. et al. State-of-art advances on removal, degradation and electrochemical monitoring of 4-aminophenol pollutants in real samples: A review. Environ. Res. 222, 115338. https://doi.org/10.1016/j.envres.2023.115338 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Coes, A. L., Paretti, N. V., Foreman, W. T., Iverson, J. L. & Alvarez, D. A. Sampling trace organic compounds in water: A comparison of a continuous active sampler to continuous passive and discrete sampling methods. Sci. Total Environ. 473–474, 731–741. https://doi.org/10.1016/j.scitotenv.2013.12.082 (2014).Article 
    CAS 
    PubMed 
    ADS 

    Google Scholar 
    Li, W. C. Occurrence, sources, and fate of pharmaceuticals in aquatic environment and soil. Environ. Pollut. 187, 193–201. https://doi.org/10.1016/j.envpol.2014.01.015 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Kato, S. & Kansha, Y. Comprehensive review of industrial wastewater treatment techniques. Environ. Sci. Pollut Res. Int. 31, 51064–51097. https://doi.org/10.1007/s11356-024-34584-0 (2024).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Oluwole, A. O., Omotola, E. O. & Olatunji, O. S. Pharmaceuticals and personal care products in water and wastewater: A review of treatment processes and use of photocatalyst immobilized on functionalized carbon in AOP degradation. BMC Chem. 14, 62. https://doi.org/10.1186/s13065-020-00714-1 (2020).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Li, S. et al. Occurrence and distribution of atrazine in groundwater from agricultural areas in China. Sci. Total Environ. 955, 177161. https://doi.org/10.1016/j.scitotenv.2024.177161 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Ahrens, L. & Bundschuh, M. Fate and effects of poly- and perfluoroalkyl substances in the aquatic environment: A review. Environ. Toxicol. Chem. 33, 1921–1929. https://doi.org/10.1002/etc.2663 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Posselt, M., Jaeger, A., Schaper, J. L., Radke, M. & Benskin, J. P. Determination of Polar organic micropollutants in surface and pore water by high-resolution sampling-direct injection-ultra high performance liquid chromatography-tandem mass spectrometry. Environ. Sci. Process. Impacts. 20, 1716–1727. https://doi.org/10.1039/c8em00390d (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Menger, F., Ahrens, L., Wiberg, K. & Gago-Ferrero, P. Suspect screening based on market data of Polar halogenated micropollutants in river water affected by wastewater. J. Hazard. Mater. 401, 123377. https://doi.org/10.1016/j.jhazmat.2020.123377 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Matamoros, V., Arias, C., Brix, H. & Bayona, J. M. Preliminary screening of small-scale domestic wastewater treatment systems for removal of pharmaceutical and personal care products. Water Res. 43, 55–62. https://doi.org/10.1016/j.watres.2008.10.005 (2009).Article 
    CAS 
    PubMed 
    ADS 

    Google Scholar 
    Lindahl, A. et al. A web-based pesticide risk assessment tool for drinking water protection zones in Sweden. J. Environ. Manage. 357, 120700. https://doi.org/10.1016/j.jenvman.2024.120700 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Dai, G. et al. Pharmaceuticals and personal care products (PPCPs) in urban and suburban rivers of Beijing, china: occurrence, source apportionment and potential ecological risk. Environ. Sci. Process. Impacts. 18, 445–455. https://doi.org/10.1039/c6em00018e (2016).Article 
    CAS 
    PubMed 

    Google Scholar 
    Moller, A. et al. Distribution and sources of polyfluoroalkyl substances (PFAS) in the river rhine watershed. Environ. Pollut. 158, 3243–3250. https://doi.org/10.1016/j.envpol.2010.07.019 (2010).Article 
    CAS 
    PubMed 

    Google Scholar 
    Sorengard, M., Bergstrom, S., McCleaf, P., Wiberg, K. & Ahrens, L. Long-distance transport of per- and polyfluoroalkyl substances (PFAS) in a Swedish drinking water aquifer. Environ. Pollut. 311, 119981. https://doi.org/10.1016/j.envpol.2022.119981 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Download referencesAcknowledgementsThis work was carried out in the framework of the European Partnership for the Assessment of Risks from Chemicals (PARC) and has received co-funding from the European Union’s Horizon Europe research and innovation programme under Grant Agreement No 101057014. Views and opinions expressed are, however, those of the author(s) only and do not necessarily reflect those of the European Union or the Health and Digital Executive Agency. Neither the European Union nor the granting authority can be held responsible for them. We also acknowledge financial support from the Swedish EPA (Naturvårdsverket).FundingOpen access funding provided by Swedish University of Agricultural Sciences. This work was carried out in the framework of the European Partnership for the Assessment of Risks from Chemicals (PARC) and has received co-funding from the European Union’s Horizon Europe research and innovation programme under Grant Agreement No 101,057,014. Views and opinions expressed are, however, those of the author(s) only and do not necessarily reflect those of the European Union or the Health and Digital Executive Agency. Neither the European Union nor the granting authority can be held responsible for them. We also acknowledge financial support from the Swedish EPA (Naturvårdsverket).Author informationAuthors and AffiliationsDepartment of Aquatic Sciences and Assessment, Swedish University of Agricultural Sciences (SLU), PO Box 7050, SE-75007, Uppsala, SwedenJavad Mottaghipisheh, Rajneesh Gautam & Lutz AhrensAuthorsJavad MottaghipishehView author publicationsSearch author on:PubMed Google ScholarRajneesh GautamView author publicationsSearch author on:PubMed Google ScholarLutz AhrensView author publicationsSearch author on:PubMed Google ScholarContributions**Javad Mottaghipisheh: ** Investigation, Writing – original draft, review and editing, Formal analysis, Data curation, Data analysis, Experiments; **Rajneesh Gautam: ** Writing – original draft, review and editing, Data curation, Visualization; **Lutz Ahrens: ** Writing– review & editing, Conceptualization, Methodology, Supervision, Funding acquisition.Corresponding authorCorrespondence to
    Javad Mottaghipisheh.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary InformationBelow is the link to the electronic supplementary material.Supplementary Material 2Supplementary Material 1Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
    Reprints and permissionsAbout this articleCite this articleMottaghipisheh, J., Gautam, R. & Ahrens, L. Combined RP-HILIC for suspect screening of persistent, mobile, and toxic substances in surface water: A case study.
    Sci Rep 15, 43272 (2025). https://doi.org/10.1038/s41598-025-29664-1Download citationReceived: 25 August 2025Accepted: 18 November 2025Published: 05 December 2025Version of record: 08 December 2025DOI: https://doi.org/10.1038/s41598-025-29664-1Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative
    KeywordsHigh-resolution mass spectrometryHydrophilic interaction liquid chromatography (HILIC)Reversed-phase (RP)Suspect screeningSurface waterWastewater water treatment plant (WWTP) More