AGU Journals

  • Search term Advanced Search Citation Search
  • Individual login
  • Institutional login

Log in with AGU

Geophysical Research Letters

Transition Zone Morphology Dynamics of Dissolved Oxygen (DO) in a Salinity-Impacted Hyporheic Zone

Qihao Jiang

  • orcid.org/0000-0003-0370-089X

State Key Laboratory of Hydrology-Water Resources and Hydraulic Engineering, Hohai University, Nanjing, China

College of Agricultural Science and Engineering, Hohai University, Nanjing, China

Contribution: Conceptualization, Methodology, Software, Validation, Formal analysis, ​Investigation, Resources, Writing - original draft, Writing - review & editing

Matthew H. Kaufman

  • orcid.org/0000-0003-1498-8219

Pacific Northwest National Laboratory, Earth and Biological Sciences Division, Richland, WA, USA

Contribution: Conceptualization, Methodology, Validation, Formal analysis, Data curation, Writing - review & editing

Corresponding Author

Guangqiu Jin

[email protected]

  • orcid.org/0000-0001-9021-1925

College of Water Conservancy and Hydropower Engineering, Hohai University, Nanjing, China

Correspondence to:

Contribution: Writing - review & editing, Visualization, Supervision, Funding acquisition

Hongwu Tang

  • orcid.org/0000-0003-2271-5288

Contribution: Resources, Writing - review & editing, Project administration

  • orcid.org/0000-0003-1467-7883

Contribution: Validation, ​Investigation, Writing - review & editing, Visualization, Supervision

Freshwater salinization is a common yet underappreciated environmental problem in rivers, yet how this process impacts the transport of dissolved oxygen (DO) in the hyporheic zone remains unknown. Using flume experiments and numerical simulations, we have demonstrated that the morphological dynamics of the DO transition zone are controlled by the dimensionless number Da·Ra −2 , a measure of the rate of aerobic respiration to free convection. At low Da·Ra −2 , the oxic front breaks up into fingers which subsequently grow, and leads to accelerating of DO transport and shrinking of the anaerobic zone in the underlying sediment. As Da·Ra −2 increases, the growth rate of the instability decreases, and the DO plume fingers are suppressed and delayed compared with the saltwater-freshwater interface, the mixing area between saltwater and freshwater. These results indicate that the freshwater salinization syndrome can have significant impacts on the functioning of aquifer ecosystems unless regulated and managed effectively.

Experiments and simulations are performed to investigate the morphology of dissolved oxygen (DO) transition zone in a salinity-impacted bedform

Asynchronous transport of salt and DO is controlled by a dimensionless number Da·Ra −2

Whether density-driven flow can be a DO pump depends on the intensity of aerobic respiration

Plain Language Summary

Increases in salinity, alkalinity, and major ions occur in a wide range of freshwater ecosystems around the world, and the freshwater salinization syndrome is becoming a serious water quality problem and a major chemical signature of the Anthropocene. The goal of this study is to understand how dissolved oxygen (DO), which is beneficial to the survival of benthic invertebrates, is transported in the sediment in response to freshwater salinity fluctuation. Flume experiments and numerical simulations are performed to investigate the development of DO plume in the riverbed. The experimental outcomes, obtained through high-resolution imagining techniques, show that the convection induced by saltwater density gradients in the hyporheic zone could influence DO plume development. Our numerical model is capable of describing the transport behaviors of DO before and after salinization in the surface water. A dimensionless number Da·Ra −2 is proposed to predict DO transport dynamics in the river with freshwater salinization syndrome. The implications and potential threats of freshwater salinization for the river ecosystem, including the biogeochemical processes and contaminant attenuation in the hyporheic zone, are also discussed.

1 Introduction

Dissolved oxygen (DO) plays an important role in a range of biogeochemical processes (e.g., carbon and nitrogen cycling) of aquatic systems (Peiffer et al.,  2021 ; Torgersen & Branco,  2007 ; Trimmer et al.,  2012 ). DO concentrations and consumption rates are primary indicators of anaerobic and aerobic respiration in the hyporheic zone (Haggerty et al.,  2014 ; Reeder et al.,  2018 ), which contribute substantially to greenhouse gas (e.g., CO 2 and N 2 O) emissions of fluvial systems (Jiang, Jin, Tang, Xu, & Chen,  2021 ; Quick et al.,  2016 ). In fact, it has been documented that 40%–90% of respiration occurs in the hyporheic zone (Battin et al.,  2003 ; Fellows et al.,  2001 ). The consumption rate of DO in the riverbed is controlled by the hyporheic exchange flux and the activities of microbial communities (Kaufman et al.,  2017 ; Reeder et al.,  2018 ). DO availability in the hyporheic zone is also a key environmental variable for the survival and development of benthic invertebrates (Chapman,  1988 ; Croijmans et al.,  2021 ). It exhibits pronounced concentration gradients inside fish redds, depending on whether hyporheic flow and transport occur in the streambed (Cardenas et al.,  2016 ).

Sediment–water interface (SWI) vertical DO fluxes have been obtained via in-situ measurement of drinking reservoirs, and the results determined the hydrodynamics affects the fluxes by changing the diffusive boundary layer thickness (Zhang et al.,  2021 ). In lotic systems, fluvial geomorphology has significant impacts on the hyporheic flow (Kasahara & Wondzell,  2003 ). In comparison with lateral exchange through riverbank, vertical exchange beneath submerged bedforms dominates hyporheic fluxes along the river corridor (Gomez-Velez & Harvey,  2014 ). Bedform-driven hyporheic exchange is caused by pressure gradients along the bed surface, which are dependent on the surface water flow velocity and sediment permeability (Cardenas & Wilson,  2007a ; Song et al.,  2017 ). As the flux of DO across the SWI is associated with the hydraulic gradient, it is primarily regulated by bed topography, stream water velocity and aerobic respiration rate (Kaufman et al.,  2017 ; Reeder et al.,  2018 ). Ziebis et al. ( 1996 ) and Kaufman et al. ( 2017 ) investigated advective oxygen transport in permeable sediments under different stream flow patterns. The former results revealed that oxygen could be rapidly transported into deeper layers and enhanced the mineralization of organic matter (Ziebis et al.,  1996 ). The latter ones focus attention on the morphology of the DO plume front, which was depends on surface flow acceleration or deceleration (Kaufman et al.,  2017 ). A large number of studies have also indicated that the spatial distributions of oxygen have significant impacts on the areas and rates of nitrification or iron precipitation (Bardini et al.,  2012 ; Dwivedi et al.,  2018 ).

In addition to stream hydrodynamics, the variation of river hydrochemistry induced by saltwater pollution, can also affect the hyporheic exchange in salinity-impacted rivers (Boano et al.,  2009 ; Jiang, Jin, Tang, Xu, & Chen,  2021 ; Jiang et al.,  2020 ; Jin et al.,  2011 ). Freshwater salinization due to human activities is an emerging global problem that impacts drinking water safety, ecosystem health, biodiversity, infrastructure corrosion, and food production (Kaushal et al.,  2018 ,  2021 ). For example, chloride concentrations are generally less than 20 mg L −1 in surface freshwater but in the range of 10–7,730 mg L −1 in rivers polluted by road salts (Hintz & Relyea,  2019 ). Many studies have investigated the porewater flow that dominates hyporheic exchange and salt transport driven by density gradients (Jiang et al.,  2020 ; Jin et al.,  2011 ,  2015 ). Jiang, Jin, Tang, Xu, and Chen ( 2021 ) have shown that greenhouse gas N 2 O emissions from the hyporheic zone are linked to the penetration of DO, which was carried by the saltwater in the river. However, the effect of river salinization on the DO dynamics in the hyporheic zone remains unclear.

In this study, laboratory flume experiments similar to those of Kaufman et al. ( 2017 ) were conducted to illustrate the development of DO plumes and transition zone (i.e., the area of the plume between 10% and 90% oxygen saturation) in the hyporheic zone under the influence of freshwater salinization using a planar optode imaging approach, which is capable of capturing DO dynamics in the hyporheic zone at high temporal frequency and spatial resolution. In numerical simulations, we employed both constant density, that is, the fluid density in the bedform are not changed with the infiltrating saltwater from stream and variable density models, that is, the fluid density in the bedform was assumed to be linearly increase with the infiltrating saltwater concentrations, coupled with DO transport and consumption behaviors. Comparison between simulated and experimental results allowed us to better understand the effect of freshwater salinization on the oxygen dynamics in the hyporheic zone.

2 Laboratory Experiments and Numerical Simulations

2.1 laboratory experiments.

Laboratory experiments were carried out in a water-recirculating flume (Figure S1a), which was 8 m long by 0.3 m wide by 1.3 m high following the suggestions of Kaufman et al. ( 2017 ,  2021 ). The water in the flume was collected from the Lower Colorado River near the Hornsby Bend in Austin, Texas, USA, to help provide realistic water quality conditions and to encourage the growth of microbial communities in the flume sediment. The flume had a sediment zone of 5 m long by 0.3 m wide by 0.7 m deep, and the sediment was clean and homogenous quartz sand ( D 50  = 1.5 mm) mixed with 0.05 wt.% of finely crushed walnut shells to provide a source of dissolved organic carbon. Seven stable bedforms were manually built and were 40–45 cm long and 4 cm tall, and the length of the lee face was approximately 1/3 of that of the stoss face. The water depth was about 5.25 cm above the peak bedform and the flow rate of flume was 50 L per minute, corresponding to a flow velocity of about 3.175 m per minute. Prior to measurement, the flume was incubated for approximately 5 months to ensure the growth and thriving of microbial communities.

The whole experimental procedure can be divided into two steps. The DO plume locations at the steady state (120 hr static discharge periods) without saltwater was determined (i.e., the background surface water conductivity was around 730 uS cm ‒1 ). After that, the addition of saltwater lasting 45 hr, which raised the surface water conductivity to a peak of 1,632 uS cm ‒1 , and DO transport behaviors were observed via the planar optode imaging (80 cm tall by 60 cm wide optode), which extended approximately 65 cm into the bed below the SWI (Figure S1b). Images were taken every 10 min, and the values should range from about 0 to 100 in percent of oxygen saturation. The design of planar optode DO sensor, planar optode imaging and calibration, analysis of DO fields were performed following the methods proposed by Larsen et al. ( 2011 ) and Kaufman et al. ( 2017 ).

2.2 Numerical Simulations

The conceptual model of the flow processes could be divided into two parts. First, the 2D stream flow over the fixed bedform was modeled using the Reynolds-Averaged Navier-Stokes (RANS) equations with k - ω turbulence closure scheme, which could be solved by the pisoFoam solver ( www.openfoam.org ) in OpenFOAM (Jiang et al.,  2020 ; Weller et al.,  1998 ). This method has been commonly used in 2D and 3D simulations of hyporheic exchange (Cardenas & Wilson,  2007b ; X. Chen et al.,  2015 ; Janssen et al.,  2012 ; Zheng et al.,  2019 ). The resulting pressure distribution along the bedform surface was used as the boundary conditions at the SWI for the pore water flow in the hyporheic zone. Second, the pore water flow in the streambed was simulated without considering the density effect (NDE; ρ  =  ρ f ) and considering density effect (DE; ρ  =  ρ f  +  γC salt ) due to concentration differences between infiltrating saltwater and pore water. The buoyancy effect in stream domain was ignored (i.e., solute transport was not included in the stream domain). The DO transport model with a first-order respiration reaction was coupled with the density-dependent flow to characterize DO transport behaviors in the streambed under the effect of river salinization.

urn:x-wiley:00948276:media:grl64786:grl64786-math-0001

The streambed permeability k [L 2 ], fluid viscosity μ [M L −1 T −1 ] and respiration rate k 1 [T −1 ] are constant, while the reduced pressure p  =  P − ρ f gz [M L −1  T −2 ] is obtained by eliminating the hydrostatic pressure from P [M L −1  T −2 ], the pressure in the liquid. The density ρ [M L −3 ] is assumed to depend linearly on C salt ( ρ  =  ρ f (1 +  βC salt ), where the freshwater density ρ f [M L −3 ] and the coefficient of density change because the concentration β [−] is a constant). The acceleration due to gravity is g [L T −2 ], θ [‒] is the porosity, D [L 2 T −1 ] is the hydrodynamic dispersion coefficient, which is calculated as described in previous studies (Jiang et al.,  2020 ; Jiang, Jin, Tang, Xu, & Chen,  2021 ; Jiang, Jin, Tang, Xu, Wei, & Li,  2021 ), and i is a vertical unit vector codirectional with the positive z axis.

urn:x-wiley:00948276:media:grl64786:grl64786-math-0005

Ra  =  k∆ρgL Z /( μDθ ) is the saltwater Rayleigh number, where ∆ ρ  =  ρ f βC s is the maximum density difference between freshwater and saltwater, and Da =  k 1 L z 2 /( Dθ ) is the Damkohler number. The vertical extent of the bedform L Z is used to calculate individual nominal values of Ra and Da, which allows for easier quantification. It is clear that the only parameter determining the flow and transport in this system is Da·Ra −2 when the penetration depth of the DO-rich layer is smaller than the sediment depth. The boundary conditions are formulated as described in the study of Jiang, Jin, Tang, Xu, and Chen ( 2021 ) and Jiang, Jin, Tang, Xu, Wei, & Li ( 2021 ).

Before saltwater addition (Figure  1a , t  = 0), the penetration depth of DO into the sediment at the steady state was determined by time-lapse photography in order to illustrate pore water flow patterns induced by current-bedform interactions. It is seen that the stream-borne DO enters the streambed on the stoss side and migrates along the flow path, forming a conchoidally shaped plume beneath the bedform (Figure  1b , t  = 0). The plume also expands laterally due to the horizontal ambient flow and an anoxic zone is formed between two oxic layers. The model without considering the density effect can accurately predict DO plume distribution and front location but fails to capture the anoxic zone (Figure  1c , t  = 0). The reason of generating anoxic zone in the DO plume may be attributed to the various of followings: First, the heterogeneous growth of microbial communities (Arnon et al.,  2007 ; De Falco et al.,  2018 ) in oxic areas of the streambed could trigger the dead zone (Deng et al.,  2013 ; Zhou et al.,  2020 ), which prevent the DO transport into these areas. Second, the potential role of small-scale heterogeneities in hydraulic conductivity also causes the uneven transport of DO.

Details are in the caption following the image

(a) The salinity fluctuation in the stream water after adding the saltwater, (b) time series photographs via flume experiments and model results (c) with variable-density flow and (d) without variable-density flow. Water flows from left to right at an overlying water velocity of 0.05 m s ‒1 . The red solid line indicates the boundary between hyporheic flow and ambient flow. The red dashed circle indicated the anoxic zone in the DO plume.

After saltwater is added to the stream (Figure  1a , t  > 0), the constant density model (Figure  1d ) is not capable of describing the DO plume development via planar optode (Figure  1b ). DO transport is clearly distorted after 15 hr (Figure  1b , t  = 15 hr), because the density gradients between infiltrating saltwater and ambient groundwater lead to gravitational instabilities (Figure  1c , t  = 15 hr). After 30 hr, it develops into streaming currents with increasing stream salinity, which profoundly accelerates the downward flow of pore water (Figure  1b ). Hence, the density gradients increase the respiration rate and the oxic area, which transports the DO-rich water downward to increasing depths and leads to the formation of DO plume fingers. Note that although the plume finger flow appears in a somewhat random fashion, the model with density-driven flow is capable of describing the development of DO plume (Figure  1c ). In contrast, without an additional engine or pump given by the salinity fluctuations, DO will be confined to the shallow regions of the benthic zone, as evident from the model without density-driven flow (Figure  1d ).

As the variable-density model can well reproduce experimentally observed patterns, we further investigate how density-driven flow and aerobic respiration influence DO plume propagation in the bed. The dimensionless number Da·Ra −2 , which is indicative of the ratio between density-driven flow induced by buoyancy and microbial aerobic respiration, is proposed in Equation  7 to determine the development of DO plume in the bed (Figure  2 ). A chemically inert (i.e., no aerobic respiration) case A (Da = 0, Ra = 32,600, Da·Ra −2  = 0), a weakly reactive case B (Da = 32,600, Ra = 32,600, Da·Ra −2  = 3.1 × 10 −5 ), and strong reaction cases C (Da = 326,000, Ra = 32,600, Da·Ra −2  = 3.1 × 10 −4 ), D (Da = 749,800, Ra = 32,600, Da·Ra −2  = 7.1 × 10 −4 ) and E (Da = 1,515,900, Ra = 32,600, Da·Ra −2  = 1.4 × 10 −3 ) are considered. It is seen that the transport of saltwater and DO is progressively delayed with increasing Da·Ra −2 (Figure  2 ). At the same dimensionless time (e.g., t * = 27,000), DO finger develops drastically with a thin transition zone and steep (sharp) concentration gradient in cases A and B, which is coincident with the development of saltwater plume. For cases C, D and E with higher Da·Ra −2 values (>3.1 × 10 −4 ), the finger-like oxic front is significantly inhibited and lagged by the aerobic respiration, and this reaction will attenuate the convective motion induced by the density-driven flow. The strong reaction in cases D and E suppresses the growth of fingers and the fluctuation of finger depth. Above the limit of Da·Ra −2 ≈ 1.4 × 10 −3 , the DO penetration depth stabilizes but convection evidently develops in the fluid below. In the end, the DO plume is distributed in a wider transition zone with border concentration gradients, which is dominated by dispersion/diffusion and respiration reaction.

Details are in the caption following the image

Effects of Da·Ra ‒2 on the dimensionless saltwater ( C saltwater / C saltwater , 0 ) and DO concentration ( C DO / C DO , 0 ) at dimensionless times (i) t * = 13,500, (ii) t * = 27,000, and (iii) t * = 39,600.

To further quantify the asynchronous transport of salt and DO, we focus on the morphologies of the oxygenated zone and the saltwater-fresh water mixing zone. The DO and saltwater plume front location d (Figure S1c) is calculated based on the 0.9 concentration contour line, and the width of the mixing area is calculated based on the vertical distance between 0.1 and 0.9 concentration contour lines, respectively (Figure  2 ). In Figure  3a , we plot the dimensionless DO and saltwater plume front location d * ( d / L z ), for a range of values of Da Ra −2 . For the low Da·Ra −2 , the d * versus Da·Ra −2 points overlap between saltwater and DO. Such overlap shows that the convection induced by saltwater density gradients simultaneously drag the DO plume development (Figures  2a and  2b ). However, the asynchronous transport of saltwater and DO becomes more pronounced with increasing Da·Ra −2 , especially at larger dimensionless times (e.g., t * = 39,600), and the location difference can reach 42.3% (Figure  3a ). Similar to front location, the width of the transition zone h * ( h / L z ) also shows a synchronous pattern at low Da·Ra −2 . However, our results indicate that the relationship between the width of the transition zone and Da·Ra −2 is rather complex, and it should not be expected to be always monotonic over a wide range of conditions. The maximum h * is expected to occur at an intermediate Da·Ra −2 (3.1 × 10 −4 ) because of the shift from convection to aerobic respiration (Figure  2c ).

Details are in the caption following the image

(a) The plume front location and (b) the transition zone of DO and saltwater under different dimensionless times.

4 Discussion and Conclusions

This work provides the first experimental look at the effects of dynamic river salinity fluctuation on the transport and distribution of stream-borne DO in the streambed. The high-resolution imaging results show that aerobic respiration is linked to the physical processes of hyporheic flow, mixing, and saltwater transport, forming salinity gradients and a biogeochemically active environment. This simple fact has potentially profound impacts on hyporheic exchange processes and DO dynamics of broad interest.

The results show that the density gradients between infiltrating saltwater and ambient groundwater led to gravitational instabilities and accelerated downward flow of pore water, which are expected to control the DO penetration depth and oxic area. Therefore, even a slight modification of hydrodynamic processes in the hyporheic zone may have a profound impact on benthic habitat and the capacity of streams to process biological and environmental compounds. For this reason, salinity is considered as an additional engine or pump that promote DO transport into the riverbed, which is beneficial for early stage redd development (Cardenas et al.,  2016 ) and the survival of invertebrates in the benthic zone where there is limited access to oxygen (Chambers et al.,  2000 ; Peralta-Maraver et al.,  2018 ).

The simulation results also suggest that DO transport is controlled by the dimensionless number Da·Ra −2 in salinized rivers. As previously discussed, the presence of strong respiration reaction inhibits the formation of oxic front fingers which may otherwise increase metabolic rates by channeling DO downward. Instead, reaction favors conchoidally shaped plume with roughly uniform lengths, resulting in mass flux values which become steadier and more constant with increasing Da·Ra −2 . When Da·Ra −2  ≈ O(1), respiration reaction dominates free convection: The DO is consumed instantaneously as it enters the benthic zone, leading to a constant inward flux without a boundary layer. However, the hyporheic zone system dominated by the low Da·Ra −2 values should receive enough attention. Freshwater salinization syndrome is a common yet underappreciated problem resulting from salt pollution such as road deicers, irrigation runoff, sewage, and potash (Cañedo-Argüelles et al.,  2016 ; Herbert et al.,  2015 ; Hintz et al.,  2022 ; Kaushal et al.,  2018 ). The chloride concentration may reach nearly 25% (5,000 mg L −1 ) of that of seawater in some streams (Hintz & Relyea,  2017 ). The convection induced by saltwater density gradients in the hyporheic zone could simultaneously drag the DO plume development in case of Da·Ra −2 < 3.1 × 10 −5 , and kept the synchronous transport pace (Figure  2 ). This result implies that the anaerobic area may shrink and even disappear because of saltwater-freshwater interactions. The evident invasion and expansion of the oxic area also is likely to alter the distributions of microbial communities, potentially promoting aerobic respiration and greenhouse gas CO 2 emissions.

The deepening of the oxic-anoxic interface and the reduction of the anoxic area due to the variable-density flow also could be detrimental to a cascade of redox reactions, such as nitrate reduction (Jiang, Jin, Tang, Xu, Wei, & Li,  2021 ,  2022 ), iron reduction and sulfate reduction (Dwivedi et al.,  2018 ; Ng et al.,  2020 ), that are responsible for the bioreactor ability of the hyporheic zone to transform nutrients and pollutants not only from stream water (Yang et al.,  2018 ) but also from groundwater (Hester et al.,  2014 ). The enhanced downwelling flow of low Da·Ra −2 systems decreases the anoxic area and provides additional hindrance for mixing-dependent reactions of upwelling groundwater pollutants, such as chlorinated ethenes (Weatherill et al.,  2018 ).

This study has focused on the flow-controlled effects of DO dynamics from a hydrodynamic perspective. However, the microbiological aspects were highly simplified. On one hand, the model ignored the bioclogging induced by biomass growth, and this factor could lead to an equilibrium between permeability reduction (Caruso et al.,  2017 ). The uneven DO transport in the bedform due to small-scale heterogeneity and resulting dead zones, which have been observed in the biofilm growth systems (Deng et al.,  2013 ; Zhou et al.,  2020 ). This could be used to explain the anoxic microzones in the observed DO plume and required to validated in future studies. In recent hyporheic studies (Briggs et al.,  2015 ; Roy Chowdhury et al.,  2020 ), microzones, that is, sediment pores depleted in oxygen (i.e., O2 concentration less than 2.0 mg/L), are also found and embedded within an oxygen-rich porous domain. On the other hand, the response of microbial communities to the river salinization were also ignored. The movement of the saltwater-freshwater transition zone in response to river salinization may also affect the types and activities of microbial communities in the benthic zone (L. Chen et al.,  2019 ; Dang et al.,  2019 ), and the microbe-mediated DO consumption reactions could be further changed.

Though anthropogenic activities have increased the rate of salinization of freshwater ecosystems around the world, the climate change in the salinized river is another key factor impacting the hyporheic DO fate. River temperature has been determined to be a primary driver of aquatic organism metabolism in both stream and hyporheic zone (Marzadri et al.,  2013 ). Though the object of this study was to demonstrate the sole dependence of hyporheic DO dynamics on stream salinity fluctuations and ignored temperature fluctuations, this still has some implications for future studies. First, temperature is an important control on groundwater flow and biogeochemical activity in aquifers due to its effect on fluid viscosity and microbial reaction kinetics. Previous studies have determined the temperature varied from 5°C to 35°C significantly changed the aerobic respiration rate (Cogswell & Heiss,  2021 ; Zheng & Cardenas,  2018 ; Zheng et al.,  2016 ). Assuming that the salinity of the surface water fluctuates very minor during the day, the spatial extent of DO in the hyporheic zone may vary due to temperature fluctuations, because the DO consumption rates are different during the day and night according the Arrhenius law (Dawson & Murphy,  1972 ). Second, the effect of temperature could have different contribution on DO consumption between rivers with different size, because large streams have a stable temperature that slightly fluctuate around the daily mean than small streams (Marzadri et al.,  2013 ). The nonlinear coupling between flow, temperature, salinity and reaction in the model is another issue to be aware of in future studies (Nguyen et al.,  2020 ). Both anthropogenic and climatic impacts are significant at the local scale, and they may cause spatially and temporally varying DO hot spots within the hyporheic zone. Overall, the freshwater salinization syndrome can have significant impacts on the functioning of ecosystems such as drinking water safety, contaminant retention, biodiversity, and climate change, unless regulated and managed effectively in the future.

Acknowledgments

This research was supported by the National Natural Science Foundation of China (U2040205 and 52109082), the Fundamental Research Funds for the Central Universities (B210202116), China Postdoctoral Science Foundation (2020M681476), the National Key R&D Program of China (2019YFE0109900), the National Science Foundation (EAR-0955750 and EAR-1344547) and the Geology Foundation at the University of Texas at Austin provided support of this study. We thank two anonymous reviewers and editor Harihar Rajaram for their helpful comments which led to substantial improvements of this study.

Open Research

Data availability statement.

The data that support the findings of this study are available online at https://doi.org/10.5281/zenodo.6649097 .

Supporting Information

Filename Description
1.1 MB Figure S1

Please note: The publisher is not responsible for the content or functionality of any supporting information supplied by the authors. Any queries (other than missing content) should be directed to the corresponding author for the article.

  • Arnon, S. , Gray, K. A. , & Packman, A. I. ( 2007 ). Biophysicochemical process coupling controls nitrogen use by benthic biofilms . Limnology and Oceanography , 52 , 1665 – 1671 . https://doi.org/10.4319/lo.2007.52.4.1665 10.4319/lo.2007.52.4.1665 CAS ADS Web of Science® Google Scholar
  • Bardini, L. , Boano, F. , Cardenas, M. B. , Revelli, R. , & Ridolfi, L. ( 2012 ). Nutrient cycling in bedform induced hyporheic zones . Geochimica et Cosmochimica Acta , 84 , 47 – 61 . https://doi.org/10.1016/j.gca.2012.01.025 10.1016/j.gca.2012.01.025 CAS ADS Web of Science® Google Scholar
  • Battin, T. J. , Kaplan, L. A. , Newbold, J. D. , & Hendricks, S. P. ( 2003 ). A mixing model analysis of stream solute dynamics and the contribution of a hyporheic zone to ecosystem function . Freshwater Biology , 48 ( 6 ), 995 – 1014 . https://doi.org/10.1046/j.1365-2427.2003.01062.x 10.1046/j.1365-2427.2003.01062.x Web of Science® Google Scholar
  • Boano, F. , Poggi, D. , Revelli, R. , & Ridolfi, L. ( 2009 ). Gravity-driven water exchange between streams and hyporheic zones . Geophysical Research Letters , 36 ( 20 ), L20402. https://doi.org/10.1029/2009GL040147 10.1029/2009GL040147 ADS Web of Science® Google Scholar
  • Briggs, M. A. , Day-Lewis, F. D. , Zarnetske, J. P. , & Harvey, J. W. ( 2015 ). A physical explanation for the development of redox microzones in hyporheic flow . Geophysical Research Letters , 42 ( 11 ), 4402 – 4410 . https://doi.org/10.1002/2015GL064200 10.1002/2015GL064200 ADS Web of Science® Google Scholar
  • Cañedo-Argüelles, M. , Hawkins, C. P. , Kefford, B. J. , Schäfer, R. B. , Dyack, B. J. , Brucet, S. , et al. ( 2016 ). Saving freshwater from salts . Science , 351 ( 6276 ), 914 – 916 . https://doi.org/10.1126/science.aad3488 10.1126/science.aad3488 CAS ADS PubMed Web of Science® Google Scholar
  • Cardenas, M. B. , Ford, A. E. , Kaufman, M. H. , Kessler, A. J. , & Cook, P. L. M. ( 2016 ). Hyporheic flow and dissolved oxygen distribution in fish nests: The effects of open channel velocity, permeability patterns, and groundwater upwelling . Journal of Geophysical Research: Biogeosciences , 121 ( 12 ), 3113 – 3130 . https://doi.org/10.1002/2016JG003381 10.1002/2016JG003381 CAS ADS Web of Science® Google Scholar
  • Cardenas, M. B. , & Wilson, J. L. ( 2007a ). Dunes, turbulent eddies, and interfacial exchange with permeable sediments . Water Resources Research , 43 ( 8 ), W08412. https://doi.org/10.1029/2006WR005787 10.1029/2006WR005787 ADS Web of Science® Google Scholar
  • Cardenas, M. B. , & Wilson, J. L. ( 2007b ). Hydrodynamics of coupled flow above and below a sediment–water interface with triangular bedforms . Advances in Water Resources , 30 ( 3 ), 301 – 313 . https://doi.org/10.1016/j.advwatres.2006.06.009 10.1016/j.advwatres.2006.06.009 ADS Web of Science® Google Scholar
  • Caruso, A. , Boano, F. , Ridolfi, L. , Chopp, D. L. , & Packman, A. ( 2017 ). Biofilm-induced bioclogging produces sharp interfaces in hyporheic flow, redox conditions, and microbial community structure . Geophysical Research Letters , 44 ( 10 ), 4917 – 4925 . https://doi.org/10.1002/2017GL073651 10.1002/2017GL073651 ADS Web of Science® Google Scholar
  • Chambers, P. A. , Brown, S. , Culp, J. M. , Lowell, R. B. , & Pietroniro, A. ( 2000 ). Dissolved oxygen decline in ice-covered rivers of northern Alberta and its effects on aquatic biota . Journal of Aquatic Ecosystem Stress and Recovery , 8 ( 1 ), 27 – 38 . https://doi.org/10.1023/A:1011491706666 10.1023/A:1011491706666 CAS Google Scholar
  • Chapman, D. W. ( 1988 ). Critical review of variables used to define effects of fines in redds of large salmonids . Transactions of the American Fisheries Society , 117 ( 1 ), 1 – 21 . https://doi.org/10.1577/1548-8659(1988)117<0001:crovut>2.3.co;2 10.1577/1548-8659(1988)117 2.3.CO;2 Web of Science® Google Scholar
  • Chen, L. , Hu, B. X. , Dai, H. , Zhang, X. , Xia, C.-A. , & Zhang, J. ( 2019 ). Characterizing microbial diversity and community composition of groundwater in a salt-freshwater transition zone . Science of the Total Environment , 678 , 574 – 584 . https://doi.org/10.1016/j.scitotenv.2019.05.017 10.1016/j.scitotenv.2019.05.017 CAS ADS PubMed Web of Science® Google Scholar
  • Chen, X. , Cardenas, M. B. , & Chen, L. ( 2015 ). Three-dimensional versus two-dimensional bed form-induced hyporheic exchange . Water Resources Research , 51 ( 4 ), 2923 – 2936 . https://doi.org/10.1002/2014WR016848 10.1002/2014WR016848 ADS Web of Science® Google Scholar
  • Cogswell, C. , & Heiss, J. W. ( 2021 ). Climate and seasonal temperature controls on biogeochemical transformations in unconfined coastal aquifers . Journal of Geophysical Research: Biogeosciences , 126 ( 12 ). e2021JG006605. https://doi.org/10.1029/2021JG006605 10.1029/2021JG006605 Google Scholar
  • Croijmans, L. , De Jong, J. F. , & Prins, H. H. T. ( 2021 ). Oxygen is a better predictor of macroinvertebrate richness than temperature—A systematic review . Environmental Research Letters , 16 ( 2 ), 023002. https://doi.org/10.1088/1748-9326/ab9b42 10.1088/1748-9326/ab9b42 Web of Science® Google Scholar
  • Dang, C. , Morrissey, E. M. , Neubauer, S. C. , & Franklin, R. B. ( 2019 ). Novel microbial community composition and carbon biogeochemistry emerge over time following saltwater intrusion in wetlands . Global Change Biology , 25 ( 2 ), 549 – 561 . https://doi.org/10.1111/gcb.14486 10.1111/gcb.14486 ADS PubMed Web of Science® Google Scholar
  • Dawson, R. N. , & Murphy, K. L. ( 1972 ). The temperature dependency of biological denitrification . Water Research , 6 ( 1 ), 71 – 83 . https://doi.org/10.1016/0043-1354(72)90174-1 10.1016/0043-1354(72)90174-1 CAS Web of Science® Google Scholar
  • De Falco, N. , Boano, F. , Bogler, A. , Bar-Zeev, E. , & Arnon, S. ( 2018 ). Influence of stream-subsurface exchange flux and bacterial biofilms on oxygen consumption under nutrient-rich conditions . Journal of Geophysical Research: Biogeosciences , 123 , 2021 – 2034 . https://doi.org/10.1029/2017JG004372 10.1029/2017JG004372 CAS ADS Web of Science® Google Scholar
  • Deng, W. , Cardenas, M. B. , Kirk, M. F. , Altman, S. J. , & Bennett, P. C. ( 2013 ). Effect of permeable biofilm on micro- and macro-scale flow and transport in bioclogged pores . Environmental Science & Technology , 47 ( 19 ), 11092 – 11098 . https://doi.org/10.1021/es402596v 10.1021/es402596v CAS ADS PubMed Web of Science® Google Scholar
  • Dwivedi, D. , Steefel, C. I. , Arora, B. , Newcomer, M. , Moulton, J. D. , Dafflon, B. , et al. ( 2018 ). Geochemical exports to river from the intrameander hyporheic zone under transient hydrologic conditions: East river mountainous watershed, Colorado . Water Resources Research , 54 ( 10 ), 8456 – 8477 . https://doi.org/10.1029/2018WR023377 10.1029/2018WR023377 ADS Web of Science® Google Scholar
  • Fellows, C. S. , Valett, M. H. , & Dahm, C. N. ( 2001 ). Whole-stream metabolism in two montane streams: Contribution of the hyporheic zone . Limnology & Oceanography , 46 ( 3 ), 523 – 531 . https://doi.org/10.4319/lo.2001.46.3.0523 10.4319/lo.2001.46.3.0523 ADS Web of Science® Google Scholar
  • Gomez-Velez, J. D. , & Harvey, J. W. ( 2014 ). A hydrogeomorphic river network model predicts where and why hyporheic exchange is important in large basins . Geophysical Research Letters , 41 ( 18 ), 6403 – 6412 . https://doi.org/10.1002/2014GL061099 10.1002/2014GL061099 ADS Web of Science® Google Scholar
  • Haggerty, R. , Ribot, M. , Singer, G. A. , Martí, E. , Argerich, A. , Agell, G. , & Battin, T. J. ( 2014 ). Ecosystem respiration increases with biofilm growth and bed forms: Flume measurements with resazurin . Journal of Geophysical Research: Biogeosciences , 119 ( 12 ), 2220 – 2230 . https://doi.org/10.1002/2013JG002498 10.1002/2013JG002498 ADS Web of Science® Google Scholar
  • Herbert, E. R. , Boon, P. , Burgin, A. J. , Neubauer, S. C. , Franklin, R. B. , Ardón, M. , et al. ( 2015 ). A global perspective on wetland salinization: Ecological consequences of a growing threat to freshwater wetlands . Ecosphere , 6 ( 10 ), art206. https://doi.org/10.1890/ES14-00534.1 10.1890/ES14-00534.1 Web of Science® Google Scholar
  • Hester, E. T. , Young, K. I. , & Widdowson, M. A. ( 2014 ). Controls on mixing-dependent denitrification in hyporheic zones induced by riverbed dunes: A steady state modeling study . Water Resources Research , 50 ( 11 ), 9048 – 9066 . https://doi.org/10.1002/2014WR015424 10.1002/2014WR015424 ADS Web of Science® Google Scholar
  • Hintz, W. D. , Arnott, S. E. , Symons, C. C. , Greco, D. A. , McClymont, A. , Brentrup, J. A. , et al. ( 2022 ). Current water quality guidelines across North America and Europe do not protect lakes from salinization . Proceedings of the National Academy of Sciences , 119 ( 9 ), e2115033119. https://doi.org/10.1073/pnas.2115033119 10.1073/pnas.2115033119 CAS PubMed Web of Science® Google Scholar
  • Hintz, W. D. , & Relyea, R. A. ( 2017 ). Impacts of road deicing salts on the early-life growth and development of a stream salmonid: Salt type matters . Environmental Pollution , 223 , 409 – 415 . https://doi.org/10.1016/j.envpol.2017.01.040 10.1016/j.envpol.2017.01.040 CAS PubMed Web of Science® Google Scholar
  • Hintz, W. D. , & Relyea, R. A. ( 2019 ). A review of the species, community, and ecosystem impacts of road salt salinisation in fresh waters . Freshwater Biology , 64 ( 6 ), 1081 – 1097 . https://doi.org/10.1111/fwb.13286 10.1111/fwb.13286 Web of Science® Google Scholar
  • Janssen, F. , Cardenas, M. B. , Sawyer, A. H. , Dammrich, T. , Krietsch, J. , & de Beer, D. ( 2012 ). A comparative experimental and multiphysics computational fluid dynamics study of coupled surface-subsurface flow in bed forms . Water Resources Research , 48 ( 8 ), W08514. https://doi.org/10.1029/2012WR011982 10.1029/2012WR011982 ADS Web of Science® Google Scholar
  • Jiang, Q. , Jin, G. , Tang, H. , Shen, C. , Cheraghi, M. , Xu, J. , et al. ( 2020 ). Density-dependent solute transport in a layered hyporheic zone . Advances in Water Resources , 142 , 103645. https://doi.org/10.1016/j.advwatres.2020.103645 10.1016/j.advwatres.2020.103645 Web of Science® Google Scholar
  • Jiang, Q. , Jin, G. , Tang, H. , Xu, J. , & Chen, Y. ( 2021 ). N 2 O production and consumption processes in a salinity-impacted hyporheic zone . Journal of Geophysical Research: Biogeosciences , 126 ( 10 ), e2021JG006512. https://doi.org/10.1029/2021JG006512 10.1029/2021JG006512 CAS ADS Web of Science® Google Scholar
  • Jiang, Q. , Jin, G. , Tang, H. , Xu, J. , Wei, Q. , & Li, L. ( 2021 ). Nitrate removal processes in the riverbed during a single-peak flood event . Hydrological Processes , 35 ( 2 ), e14041. https://doi.org/10.1002/hyp.14041 10.1002/hyp.14041 CAS Web of Science® Google Scholar
  • Jiang, Q. , Liu, D. , Jin, G. , Tang, H. , Wei, Q. , & Xu, J. ( 2022 ). N 2 O dynamics in the hyporheic zone due to ripple migration . Journal of Hydrology , 610 , 127891. https://doi.org/10.1016/j.jhydrol.2022.127891 10.1016/j.jhydrol.2022.127891 CAS Web of Science® Google Scholar
  • Jin, G. , Tang, H. , Li, L. , & Barry, D. A. ( 2011 ). Hyporheic flow under periodic bed forms influenced by low-density gradients . Geophysical Research Letters , 38 ( 22 ), L22401. https://doi.org/10.1029/2011GL049694 10.1029/2011GL049694 ADS Web of Science® Google Scholar
  • Jin, G. , Tang, H. , Li, L. , & Barry, D. A. ( 2015 ). Prolonged river water pollution due to variable-density flow and solute transport in the riverbed . Water Resources Research , 51 ( 4 ), 1898 – 1915 . https://doi.org/10.1002/2014WR016369 10.1002/2014WR016369 CAS ADS Web of Science® Google Scholar
  • Kasahara, T. , & Wondzell, S. M. ( 2003 ). Geomorphic controls on hyporheic exchange flow in mountain streams . Water Resources Research , 39 ( 1 ), SBH3-1 – SBH3-14 . https://doi.org/10.1029/2002WR001386 10.1029/2002WR001386 ADS Web of Science® Google Scholar
  • Kaufman, M. H. , Cardenas, M. B. , Buttles, J. , Kessler, A. J. , & Cook, P. L. M. ( 2017 ). Hyporheic hot moments: Dissolved oxygen dynamics in the hyporheic zone in response to surface flow perturbations . Water Resources Research , 53 ( 8 ), 6642 – 6662 . https://doi.org/10.1002/2016WR020296 10.1002/2016WR020296 CAS ADS Web of Science® Google Scholar
  • Kaufman, M. H. , Warden, J. G. , Cardenas, M. B. , Stegen, J. C. , Graham, E. B. , & Brown, J. ( 2021 ). Evaluating a laboratory flume microbiome as a window into natural riverbed biogeochemistry . Frontiers in Water , 3 , 596260. https://doi.org/10.3389/frwa.2021.596260 10.3389/frwa.2021.596260 Google Scholar
  • Kaushal, S. S. , Likens, G. E. , Pace, M. L. , Reimer, J. E. , Maas, C. M. , Galella, J. G. , et al. ( 2021 ). Freshwater salinization syndrome: From emerging global problem to managing risks . Biogeochemistry , 154 ( 2 ), 255 – 292 . https://doi.org/10.1007/s10533-021-00784-w 10.1007/s10533-021-00784-w Web of Science® Google Scholar
  • Kaushal, S. S. , Likens, G. E. , Pace, M. L. , Utz, R. M. , Haq, S. , Gorman, J. , & Grese, M. ( 2018 ). Freshwater salinization syndrome on a continental scale . Proceedings of the National Academy of Sciences , 115 ( 4 ), E574 – E583 . https://doi.org/10.1073/pnas.1711234115 10.1073/pnas.1711234115 CAS PubMed Web of Science® Google Scholar
  • Larsen, M. , Borisov, S. M. , Grunwald, B. , Klimant, I. , & Glud, R. N. ( 2011 ). A simple and inexpensive high resolution color ratiometric planar optode imaging approach: Application to oxygen and pH sensing: A simple RGB based planar optode imaging approach . Limnology and Oceanography: Methods , 9 ( 9 ), 348 – 360 . https://doi.org/10.4319/lom.2011.9.348 10.4319/lom.2011.9.348 CAS Web of Science® Google Scholar
  • Marzadri, A. , Tonina, D. , & Bellin, A. ( 2013 ). Quantifying the importance of daily stream water temperature fluctuations on the hyporheic thermal regime: Implication for dissolved oxygen dynamics . Journal of Hydrology , 507 , 241 – 248 . https://doi.org/10.1016/j.jhydrol.2013.10.030 10.1016/j.jhydrol.2013.10.030 CAS ADS Web of Science® Google Scholar
  • Ng, G. C. , Rosenfeld, C. E. , Santelli, C. M. , Yourd, A. R. , Lange, J. , Duhn, K. , & Johnson, N. W. ( 2020 ). Microbial and reactive transport modeling evidence for hyporheic flux-driven cryptic sulfur cycling and anaerobic methane oxidation in a sulfate-impacted wetland-stream system . Journal of Geophysical Research: Biogeosciences , 125 ( 2 ), e2019JG005185. https://doi.org/10.1029/2019JG005185 10.1029/2019JG005185 CAS ADS Web of Science® Google Scholar
  • Nguyen, T. T. M. , Yu, X. , Pu, L. , Xin, P. , Zhang, C. , Barry, D. A. , & Li, L. ( 2020 ). Effects of temperature on tidally influenced coastal unconfined aquifers . Water Resources Research , 56 ( 4 ). e2019WR026660. https://doi.org/10.1029/2019WR026660 10.1029/2019WR026660 ADS Web of Science® Google Scholar
  • Peiffer, S. , Kappler, A. , Haderlein, S. B. , Schmidt, C. , Byrne, J. M. , Kleindienst, S. , et al. ( 2021 ). A biogeochemical–hydrological framework for the role of redox-active compounds in aquatic systems . Nature Geoscience , 14 ( 5 ), 264 – 272 . https://doi.org/10.1038/s41561-021-00742-z 10.1038/s41561-021-00742-z CAS ADS Web of Science® Google Scholar
  • Peralta-Maraver, I. , Reiss, J. , & Robertson, A. L. ( 2018 ). Interplay of hydrology, community ecology and pollutant attenuation in the hyporheic zone . Science of the Total Environment , 610–611 , 267 – 275 . https://doi.org/10.1016/j.scitotenv.2017.08.036 10.1016/j.scitotenv.2017.08.036 CAS ADS PubMed Web of Science® Google Scholar
  • Quick, A. M. , Reeder, W. J. , Farrell, T. B. , Tonina, D. , Feris, K. P. , & Benner, S. G. ( 2016 ). Controls on nitrous oxide emissions from the hyporheic zones of streams . Environmental Science and Technology , 50 ( 21 ), 11491 – 11500 . https://doi.org/10.1021/acs.est.6b02680 10.1021/acs.est.6b02680 CAS ADS PubMed Web of Science® Google Scholar
  • Reeder, W. J. , Quick, A. M. , Farrell, T. B. , Benner, S. G. , Feris, K. P. , & Tonina, D. ( 2018 ). Spatial and temporal dynamics of dissolved oxygen concentrations and bioactivity in the hyporheic zone . Water Resources Research , 54 ( 3 ), 2112 – 2128 . https://doi.org/10.1002/2017WR021388 10.1002/2017WR021388 CAS ADS Web of Science® Google Scholar
  • Roy Chowdhury, S. , Zarnetske, J. P. , Phanikumar, M. S. , Briggs, M. A. , Day-Lewis, F. D. , & Singha, K. ( 2020 ). Formation criteria for hyporheic anoxic microzones: Assessing interactions of hydraulics, nutrients, and biofilms . Water Resources Research , 56 ( 3 ). e2019WR025971. https://doi.org/10.1029/2019WR025971 10.1029/2019WR025971 Web of Science® Google Scholar
  • Song, J. , Zhang, G. , Wang, W. , Liu, Q. , Jiang, W. , Guo, W. , et al. ( 2017 ). Variability in the vertical hyporheic water exchange affected by hydraulic conductivity and river morphology at a natural confluent meander bend . Hydrological Processes , 31 ( 19 ), 3407 – 3420 . https://doi.org/10.1002/hyp.11265 10.1002/hyp.11265 ADS Web of Science® Google Scholar
  • Torgersen, T. , & Branco, B. ( 2007 ). Carbon and oxygen dynamics of shallow aquatic systems: Process vectors and bacterial productivity . Journal of Geophysical Research , 112 ( G3 ), G03016. https://doi.org/10.1029/2007JG000401 10.1029/2007JG000401 Web of Science® Google Scholar
  • Trimmer, M. , Grey, J. , Heppell, C. M. , Hildrew, A. G. , Lansdown, K. , Stahl, H. , & Yvon-Durocher, G. ( 2012 ). River bed carbon and nitrogen cycling: State of play and some new directions . Science of the Total Environment , 434 , 143 – 158 . https://doi.org/10.1016/j.scitotenv.2011.10.074 10.1016/j.scitotenv.2011.10.074 CAS ADS PubMed Web of Science® Google Scholar
  • Weatherill, J. J. , Atashgahi, S. , Schneidewind, U. , Krause, S. , Ullah, S. , Cassidy, N. , & Rivett, M. O. ( 2018 ). Natural attenuation of chlorinated ethenes in hyporheic zones: A review of key biogeochemical processes and in-situ transformation potential . Water Research , 128 , 362 – 382 . https://doi.org/10.1016/j.watres.2017.10.059 10.1016/j.watres.2017.10.059 CAS PubMed Web of Science® Google Scholar
  • Weller, H. G. , Tabor, G. , Jasak, H. , & Fureby, C. ( 1998 ). A tensorial approach to computational continuum mechanics using object-oriented techniques . Computers in Physics , 12 ( 6 ), 620 . https://doi.org/10.1063/1.168744 10.1063/1.168744 ADS Web of Science® Google Scholar
  • Yang, C. , Zhang, Y. , Liu, Y. , Yang, X. , & Liu, C. ( 2018 ). Model-based analysis of the effects of dam-induced river water and groundwater interactions on hydro-biogeochemical transformation of redox sensitive contaminants in a hyporheic zone . Water Resources Research , 54 ( 9 ), 5973 – 5985 . https://doi.org/10.1029/2018WR023286 10.1029/2018WR023286 ADS Web of Science® Google Scholar
  • Zhang, Y. , Sun, B. , Ju, W. , Gao, X. , Guo, X. , Shen, C. , et al. ( 2021 ). Eddy correlation measurements of benthic oxygen fluxes in a stratified and operated reservoir . Journal of Hydrology , 595 , 126007. https://doi.org/10.1016/j.jhydrol.2021.126007 10.1016/j.jhydrol.2021.126007 CAS Web of Science® Google Scholar
  • Zheng, L. , & Cardenas, M. B. ( 2018 ). Diel stream temperature effects on nitrogen cycling in hyporheic zones . Journal of Geophysical Research: Biogeosciences , 123 ( 9 ), 2743 – 2760 . https://doi.org/10.1029/2018JG004412 10.1029/2018JG004412 CAS ADS Web of Science® Google Scholar
  • Zheng, L. , Cardenas, M. B. , & Wang, L. ( 2016 ). Temperature effects on nitrogen cycling and nitrate removal-production efficiency in bed form-induced hyporheic zones . Journal of Geophysical Research: Biogeosciences , 121 ( 4 ), 1086 – 1103 . https://doi.org/10.1002/2015JG003162 10.1002/2015JG003162 CAS ADS Web of Science® Google Scholar
  • Zheng, L. , Cardenas, M. B. , Wang, L. , & Mohrig, D. ( 2019 ). Ripple effects: Bed form morphodynamics cascading into hyporheic zone biogeochemistry . Water Resources Research , 55 ( 8 ), 7320 – 7342 . https://doi.org/10.1029/2018WR023517 10.1029/2018WR023517 CAS ADS Web of Science® Google Scholar
  • Zhou, Y. , Weng, S. , Zhang, Y. , Zhu, D. Z. , Tang, P. , Xiang, Z. , & Zhang, T. ( 2020 ). Experimental study of seepage flow properties with biofilm development in porous media with different filter morphologies . Journal of Hydrology , 591 , 125596. https://doi.org/10.1016/j.jhydrol.2020.125596 10.1016/j.jhydrol.2020.125596 CAS Web of Science® Google Scholar
  • Ziebis, W. , Huettel, M. , & Forster, S. ( 1996 ). Impact of biogenic sediment topography on oxygen fluxes in permeable seabeds . Marine Ecology Progress Series , 140 , 227 – 237 . https://doi.org/10.3354/meps140227 10.3354/meps140227 CAS ADS Web of Science® Google Scholar

salinity and dissolved oxygen experiment

Volume 49 , Issue 18

28 September 2022

e2022GL099932

salinity and dissolved oxygen experiment

Information

salinity and dissolved oxygen experiment

Log in to Wiley Online Library

Change password, your password must have 10 characters or more:.

  • a lower case character, 
  • an upper case character, 
  • a special character 

Password Changed Successfully

Your password has been changed

Create a new account

Forgot your password.

Enter your email address below.

Please check your email for instructions on resetting your password. If you do not receive an email within 10 minutes, your email address may not be registered, and you may need to create a new Wiley Online Library account.

Request Username

Can't sign in? Forgot your username?

Enter your email address below and we will send you your username

If the address matches an existing account you will receive an email with instructions to retrieve your username

IMAGES

  1. What Is The Relationship Between Dissolved Oxygen And Salinity?

    salinity and dissolved oxygen experiment

  2. What Is The Relationship Between Dissolved Oxygen And Salinity?

    salinity and dissolved oxygen experiment

  3. The Effects of Salinity and Temperature on Dissolved Oxygen

    salinity and dissolved oxygen experiment

  4. Dissolved oxygen concentrations, salinity, and temperature measurements

    salinity and dissolved oxygen experiment

  5. What Is The Relationship Between Dissolved Oxygen And Salinity?

    salinity and dissolved oxygen experiment

  6. Ex 15 Salinity Experiment

    salinity and dissolved oxygen experiment