ArticlePDF Available

Abstract and Figures

River damming has brought great benefits to flood mitigation, energy and food production, and will continue to play a significant role in global energy supply, particularly in Asia, Africa, and South America. However, dams have extensively altered global river dynamics, including riverine connectivity, hydrological, thermal, sediment and solute regimes, and the channel morphology. These alterations have detrimental effects on the quality and quantity of fish habitat and associated impacts on aquatic life. Indeed, dams have been implicated in the decline of numerous fishes, emphasizing the need for effective conservation measures. Here, we present a global synthesis of critical issues concerning the impacts of river damming on physical fish habitats, with a particular focus on key fish species across continents. We also consider current fish conservation measures and their applicability in different contexts. Finally, we identify future research needs. The information presented herein will help support sustainable dam operation under the constraints of future climate change and human needs.
This content is subject to copyright. Terms and conditions apply.
Box 1. Tutorial
This paper contains five sections. The “Introduction” briefly describes the background of the review.
Section2 is about river damming impact on fish physical habitat, including river connectivity, hydrolog-
ical regime, sediment regime and morphology, water temperature regime, and dissolved gas. Section3
is about river damming impact on key fish species across different continents, including salmonids,
Chinese carps, sturgeon, eel and lamprey, and other typical fish species. Section4 focuses on major
conservation measures, which include fish passage facilities, artificial breeding and release, reservoir
ecological operation, and habitat compensation in tributaries, for fish in dammed rivers. The last section
highlights future research perspectives on impact assessments and mitigations, effect of climate and
land cover changes, and long-term systematic observations.
Abstract River damming has brought great benefits to flood mitigation, energy and food production, and
will continue to play a significant role in global energy supply, particularly in Asia, Africa, and South America.
However, dams have extensively altered global river dynamics, including riverine connectivity, hydrological,
thermal, sediment and solute regimes, and the channel morphology. These alterations have detrimental effects
on the quality and quantity of fish habitat and associated impacts on aquatic life. Indeed, dams have been
implicated in the decline of numerous fishes, emphasizing the need for effective conservation measures.
Here, we present a global synthesis of critical issues concerning the impacts of river damming on physical
fish habitats, with a particular focus on key fish species across continents. We also consider current fish
conservation measures and their applicability in different contexts. Finally, we identify future research needs.
The information presented herein will help support sustainable dam operation under the constraints of future
climate change and human needs.
Plain Language Summary River damming yields great social-economic benefits, but also causes
significant eco-environmental impacts, particularly on fish. Dams block fish migration routes, alter hydrological
and water temperature regimes, and modify channel morphology. These changes impact fish physical habitats
and associated communities. Here, we synthesize the impacts of river damming on fish physical habitats, and
review potential conservation measures that could be used to off-set or mitigate the impacts of dams on fish
habitats, populations and communities.
CHEN ETAL.
© 2023. The Authors.
This is an open access article under
the terms of the Creative Commons
Attribution-NonCommercial-NoDerivs
License, which permits use and
distribution in any medium, provided the
original work is properly cited, the use is
non-commercial and no modifications or
adaptations are made.
River Damming Impacts on Fish Habitat and Associated
Conservation Measures
Qiuwen Chen1,2,3 , Qinyuan Li3 , Yuqing Lin1,2,3, Jianyun Zhang1, Jun Xia4,
Jinren Ni5 , Steven J. Cooke6, Jim Best7 , Shufeng He3, Tao Feng3, Yuchen Chen3, Daniele Tonina8 ,
Rohan Benjankar9 , Sebastian Birk10 , Ayan Santos Fleischmann11 , Hanlu Yan3, and Lei Tang12
1Yangtze Institute for Conservation and Green Development, Nanjing Hydraulic Research Institute, Nanjing, China, 2State
Key Laboratory of Hydrology-Water Resources and Hydraulic Engineering, Nanjing Hydraulic Research Institute, Nanjing,
China, 3Center for Eco-Environment Research, Nanjing Hydraulic Research Institute, Nanjing, China, 4State Key Laboratory
of Water Resources and Hydropower Engineering Sciences, Wuhan University, Wuhan, China, 5The Key Laboratory of Water
and Sediment Sciences, Ministry of Education, Peking University, Beijing, China, 6Department of Biology and Institute of
Environmental and Interdisciplinary Science, Carleton University, Ottawa, ON, Canada, 7Departments of Earth Science and
Environmental Change, Geography and GIS and Mechanical Science and Engineering, and Ven Te Chow Hydrosystems
Laboratory, University of Illinois at Urbana-Champaign, Urbana, IL, USA, 8Center for Ecohydraulics Research, University of
Idaho, Boise, ID, USA, 9Civil Engineering Department, Southern Illinois University Edwardsville, Edwardsville, IL, USA,
10Faculty of Biology, Aquatic Ecology, and Centre for Water and Environmental Research, University of Duisburg-Essen,
Essen, Germany, 11Mamirauá Institute for Sustainable Development, Tefé, Brazil, 12College of Water Resources, North China
University of Water Resources and Electric Power, Zhengzhou, China
Key Points:
Dam construction alters river
hydro-geomorphological conditions
and hence influences fish habitat
quality and quantity
Knowledge of river
hydrogeomorphology and reservoir
properties can inform which
conservation measures may benefit
fish conservation
Long-term monitoring is needed
to understand causal effects and
synergies with climate change
Supporting Information:
Supporting Information may be found in
the online version of this article.
Correspondence to:
Q. Chen and J. Zhang,
qwchen@nhri.cn;
jyzhang@nhri.cn
Citation:
Chen, Q., Li, Q., Lin, Y., Zhang, J., Xia,
J., Ni, J., etal. (2023). River damming
impacts on fish habitat and associated
conservation measures. Reviews of
Geophysics, 61, e2023RG000819. https://
doi.org/10.1029/2023RG000819
Received 23 JUN 2023
Accepted 29 NOV 2023
Author Contributions:
Conceptualization: Qiuwen Chen
Data curation: Shufeng He
Funding acquisition: Qiuwen Chen
Project Administration: Yuqing Lin
Supervision: Qiuwen Chen, Jianyun
Zhang
Validation: Yuqing Lin
Visualization: Qinyuan Li, Yuchen Chen,
Hanlu Yan
Writing – original draft: Qiuwen Chen,
Qinyuan Li, Yuqing Lin, Shufeng He, Tao
Feng, Yuchen Chen, Lei Tang
10.1029/2023RG000819
REVIEW ARTICLE
1 of 64
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
2 of 64
Box 2. Glossary
ATT: accumulated temperature threshold for fish gonad maturity.
ATU: accumulated thermal units.
CI: connectivity index. For each river basin, CI is quantified for each fish species by combining its
occurrence range with a high-resolution hydrography and the locations of the dams (Barbarossa
etal.,2020).
CSI: connectivity state index. CSI of river reaches is determined with four dimensions, including
longitudinal connectivity between up- and downstream, lateral connectivity to floodplain and ripar-
ian areas, vertical connectivity to groundwater and atmosphere, and temporal connectivity based on
seasonality of flows (Grill etal.,2019).
CTT: critical temperature threshold for fish spawning.
FMCCs: four major Chinese carps.
GBD: gas bubble disease.
IHA: indicators of hydrologic alteration.
LHPs: large hydropower plants.
MPUE: mass per unit effort.
NbS: nature-based solutions.
NFR: natural flow regime.
RVA: the range of variability approach.
SHPs: small hydropower plants.
TDG: total dissolved gas.
TGR: Three Gorges Reservoir.
1. Introduction
Since the first masonry dam was constructed around 3000 BCE on the Wadi Rajil at Jawa, Jordan, river damming
has made significant positive contributions to the supply of water and energy, flood management, irrigation, and
navigation worldwide, especially in the last century (Best, 2019). The first global boom of dam construction
occurred after the Second World War, peaking in the 1960s and 1970s, mainly in Western Europe and North
America (Figures1a and1b; Lehner etal.,2011). Following increasing concerns about the social and ecological
impacts caused by river damming, the trend of global dam construction slowed in the 1990s (Moran etal.,2018).
However, to meet the rapidly growing demands for energy and water for socio-economic development, there has
been a second boom in dam construction, mainly in developing countries and emerging economies in Asia, Africa
and South America (Figures1c and1d), and particularly in large river basins such as the Amazon, Congo, and
Mekong (Winemiller etal.,2016; Zarfl etal.,2015). To date, approximately 3,700 major dams are either planned
or under construction, and the number is predicted to increase further (Zarfl etal.,2015), although removal of
aged dams has become an issue for engineering security and possible restoration of damaged river ecosystem in
some countries (Habel etal.,2020; O'Connor etal.,2015). According to the International Commission on Large
Dams (ICOLD, https://www.icold-cigb.org/), the number of registered large dams (dam height ≥15m, or 5–15m
with reservoir capacity ≥0.03km
3) worldwide reached 58,713 by April 2020. Dams can be categorized accord-
ing to their height or the associated reservoir regulation capacity. Large dams usually convert the upstream lotic
rivers into lentic reservoirs, and have an annual to seasonal regulation capacity that can significantly modify the
hydrological regime downstream; small dams (dam height <15m) and run-of-river dams typically have a small
reservoir, or even no reservoir, and are associated with weekly to sub-daily regulation capacities that impose rela-
tively low impacts on the hydrological regime downstream (D. Anderson etal.,2015; Timpe & Kaplan,2017).
River damming disrupts free flows, and hence results in habitat fragmentation (Grill etal.,2015,2019). Globally,
63% of the world's large rivers (>1,000km) are no longer free flowing (Grill etal.,2019). The construction of
over 3,700 expected hydropower dams is estimated to increase global river habitat fragmentation to 93% in the
future, which has already reached 48% due to the 6,374 existing large dams (Grill etal.,2015; Zarfl etal.,2015).
Writing – review & editing: Qiuwen
Chen, Qinyuan Li, Yuqing Lin, Jianyun
Zhang, Jun Xia, Jinren Ni, Steven J.
Cooke, Jim Best, Shufeng He, Daniele
Tonina, Rohan Benjankar, Sebastian Birk,
Ayan Santos Fleischmann
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
3 of 64
The seasonal and interannual dynamics of river flow and water temperature regimes have been greatly dampened
by river dam operation (Poff etal.,2007). As the reservoir operating time increases, sediment is deposited in
reservoirs, while the downstream riverbed tends to experience scour, eventually altering the river morphology
(Kondolf etal.,2014,2018; Schmitt etal.,2019). Moreover, the flood discharge and energy dissipation associated
with high dams can result in supersaturation of total dissolved gas in the downstream waters (Q. Ma etal.,2018;
Weitkamp & Katz,1980).
The alterations of river geophysical conditions caused by dams have significant effects on river ecology, among
which the impact on fish is of great concern because fish contribute to human welfare as key food resources.
Inland fisheries provide the equivalent of all the dietary animal protein for 158 million people globally (McIntyre
etal.,2016), particularly for poor and undernourished populations in food-insecure regions such as the Amazon
(Winemiller etal.,2016) and Mekong (Orr etal.,2012; Ziv etal.,2012) basins, where river fish are a key source
of protein. For example, in the Lower Mekong River basin, fisheries are the primary source of protein for 60
million residents (Ziv etal.,2012). Rivers also provide essential habitats for a diverse array of fishes. Alterations
in river geophysical conditions arising from dams have considerably impaired fish biodiversity and resources
worldwide (Reid etal., 2019; G. H. Su etal.,2021), as dams modify the natural flow, thermal and sediment
regimes, and decrease fish access to spawning and nursery habitats (Freeman etal.,2007). These alterations
cause dramatic reductions in fish productivity and lead to declines in fish populations, threatening global fishery
production and regional food security. The effects of river damming on fish populations have received consider-
able attention (Fullerton etal.,2010), with many studies investigating impacts on globally important commercial
and endangered fish species, such as salmonids (Hilborn,2013), eels (Atkinson etal.,2020), Chinese carp (Q.
Chen etal.,2021), and sturgeon (Z. Huang & Wang,2018).
To mitigate the negative impacts of river damming on fish, a variety of conservation measures, spanning off-setting
to mitigation, have been developed and applied. Reservoir operations have been optimized to mimic natural flow
regimes (NFRs) to satisfy fish living conditions (W. Chen & Olden,2017; Sabo etal.,2017), and fish passage
facilities have been installed at river dams to maintain biological connectivity (Katopodis & Williams,2012;
Noonan etal.,2012). Some small dams and barriers have been removed in tributary channels to rehabilitate and
Figure 1. Global dam construction and distribution. (a) Distribution of global dams. (b) Number of global total dams and number of dams constructed in each decade
from 1900 to 2017. (c) Distribution of global dams in planning and under construction. (d) Number of dams under planning and construction in each continent. Data
of panel (a) and (b) are from Global Reservoir and Dam Database (Lehner etal.,2011). Data of panel (c) and (d) are from Future Hydropower Reservoir and Dams
Database (Zarfl etal.,2015).
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
4 of 64
compensate for fish habitats lost in the dammed mainstems (Marques etal.,2018; L. Tang etal.,2021). Artificial
breeding programs have been implemented to restore fishery resources, which to some extent have succeeded
by increasing the target fish populations (Holsman etal.,2012; J. Yang etal.,2013). However, the efficiency
and effectiveness of these conservation measures have not been well investigated, leading to some controversial
perceptions concerning fish conservation measures in dammed rivers. Therefore, assessing the efficiency and
cost-effectiveness of different conservation measures could contribute significantly to identifying and promoting
efficient practices.
The number of studies documenting the impacts of river damming on fish habitats, fisheries resources, regional
food insecurity, and associated conservation measures has rapidly increased in the past decade. This review seeks
to analyze and synthesize state-of-the-art studies concerning the impacts of river damming on fish physical habi-
tats and associated conservation measures. Specifically, we focus on the geophysical aspects of hydrological and
water temperature regimes, dissolved gas supersaturation, sediment dynamics and biogeomorphology in relation
to fish habitats, with emphasis on several key fish species that occur on various continents. In addition, we assess
the cost-effectiveness of existing fish conservation measures. Finally, we formulate future key research directions
for fish conservation in dammed rivers, with the aim of helping foster sustainable hydropower development.
2. River Damming Impacts on Fish Physical Habitat
Dam construction alters river hydro-geophysical characteristics, including river connectivity, hydrological
regime, sediment regime, and morphology, water temperature regime, and dissolved gas, which have essential
impacts on fish physical habitat.
2.1. Impacts on River Connectivity
River longitudinal and lateral connectivity play a vital role in maintaining the structure and functioning of river-
ine ecosystems (Díaz etal.,2020). River systems are hierarchical tree-like networks, whose ecological function
is highly dependent on physical connectivity (Fuller etal.,2016; Grant etal.,2007). The needs of fish for diverse
habitats are strongly dependent on river connectivity and natural mobility (Arthington etal.,2016). Longitudinal
connectivity is essential for fish migration (Figures2a and2b), and lateral connectivity provides fish the access
to spawning and rearing grounds in floodplains, side channels, oxbows and floodplain lakes (Figure2c).
With the increasing number of dams, the longitudinal connectivity of global rivers is significantly under threat.
Currently, about half of global river reaches show diminished longitudinal connectivity, with the Connectivity
State Index (CSI) below 100%; nearly 10% of global river reaches have CSI below 95%, which is the minimum
value for a high level of connectivity (Grill etal.,2019). Large river networks with completely natural connectiv-
ity (CSI=100%) exist only in remote regions of the Arctic, Amazon and Congo basins (Grill etal.,2019). During
the second half of the 20th century, the impairment of river network connectivity became more severe and is now
widespread throughout the entire pan-European continent. There are at least 1.2 million river barriers (an average
density of 0.74 per km) in the 36 European countries (Belletti etal.,2020), and in general more than 50% of river
length is affected (Duarte etal.,2021). The highest barrier densities are found in rivers in central Europe where
river connectivity has been severely altered, whilst the lowest barrier densities are found in the most remote and
less populated alpine areas. Relatively undisturbed rivers exist only in parts of the Balkans, Baltic States, and
Scandinavia in Europe (Belletti etal.,2020). As far as large rivers (catchment area >10,000km
2) are concerned,
the Mediterranean and Western Atlantic regions are those most affected by fragmentation in terms of the number
of basins, while the Black Sea and Caspian Sea regions are the most affected by fragmentation in terms of river
length (Duarte etal.,2021). In the USA, large dams have increased river segmentation by 801% compared to
free-flowing streams in the absence of dams, and 79% of stream length is disconnected from their outlet of oceans
or Great Lakes (Cooper etal.,2017). In South America, more than a hundred hydropower dams have fragmented
the rivers in the Amazon basin. For the eight major river systems in the Andean region of the Amazon headwaters,
the 142 existing or under construction dams have fragmented the tributaries of six major river systems, and the
160 proposed dams could further result in a significant loss of river connectivity in the mainstreams of five major
river systems (E. Anderson etal.,2018; Flecker etal.,2022; Latrubesse etal.,2017,2020; Lees etal.,2016).
In general, the Connectivity Index (CI) of rivers is the lowest in the Europe, United States, South Africa, India
and China, and the completion of the dams currently under construction or planned will further reduce river
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
5 of 64
connectivity, particularly in countries such as India and China that are facing a boom in dam construction.
However, commensurate studies are insufficient in these developing countries (Barbarossa etal.,2020).
Dam construction also reduces lateral connectivity of rivers, which decreases the interactions between rivers
and floodplains/wetlands in river basins (Latrubesse etal.,2017; Stoffels etal.,2022), and thereby affects the
productivity of floodplain and wetland ecosystems (Palmer & Ruhi,2019). In the Atreyee River basin in India,
the active floodplain area was reduced by 66.2%, and 48.9% of total pre-dam wetland was completely obliterated,
due to the reduction in lateral connectivity caused by the Mohanpur Dam (Saha etal.,2022). In the wet-dry
tropics of Australia, dam construction reduces the average duration of the lateral connectivity between flood-
plain wetlands and their main river channels by 1% and 2% in the Flinders and Gilbert catchments, respectively
(Karim etal.,2015). In the upper Paraguay River basin in South America, the Manso dam decreased the lateral
connectivity between the Cuiaba River and Pantanal wetlands, one of the largest wetland systems in the world,
which weakened the exchange of sediments and nutrients (Jardim etal.,2020). Regulated releases also impact
on the vertical connectivity between surface and subsurface waters, which alters the exchange of solute, heat and
nutrient between surface and sediment waters (Figure2d) (Sawyer etal.,2009).
It is also necessary to emphasize river fragmentation caused by small dams, due to their large number and wide-
spread distribution, which are usually neglected (E. Anderson etal., 2018; Castello & Macedo, 2016; Fuller
etal.,2016; Rodeles etal.,2017). Couto & Olden(2018) estimate that 82,891 small hydropower plants (SHPs)
are operating or under construction across 150 countries, which is more than the number (58,713 by April 2020)
of large hydropower plants (LHPs) recorded by ICOLD. In addition, there are 181,976 SHPs planned, 10,569 of
which are to be implemented in the coming decades, indicating that the number of SHPs will continue to increase
rapidly. In developed countries, as the hydropower potential of large rivers has been mostly exploited, an increasing
number of SHPs are under planning in some of these nations, such as Austria, to meet energy demands (Wagner
etal., 2015). In water-rich countries in Asia, Africa, Latin America and southeastern Europe, the potential of
SHPs has gained particular interest from policymakers and more SHPs will be constructed (Harlan etal.,2021).
SHPs are conventionally considered as a type of clean energy resource with less environmental impact than LHPs
(Dursun & Gokcol,2011; Nautiyal etal.,2011). However, SHPs are usually set in high-gradient alpine streams
and a river basin usually contains a large number of SHPs, which lead to cumulative, and hence more severe,
impacts on river fragmentation than LHPs (Timpe & Kaplan,2017). For instance, the average loss of connectivity
Figure 2. River connectivity and the impacts of dams. (a) Migration and spawning grounds of fish in free-flowing river.
(b) Migration and spawning grounds of fish in dammed rivers, (c) Spawning and rearing grounds of fish in floodplain. (d)
Hyporheic exchange and redd habitats.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
6 of 64
due to SHPs is much higher than that due to LHPs in Brazil (Couto etal.,2021). Great concerns have been raised
about the large expansion of SHPs in rivers feeding the Pantanal wetlands, as these SHPs have affected the lateral
connectivity between rivers and wetlands (Figueiredo etal.,2021).
River fragmentation has significant impacts on freshwater fish (Barbarossa etal.,2020). Dams act as physical
barriers in spawning or foraging routes and limit the expansion of fish populations (Figure2b). Widely distributed
dams have impeded fish migration and generated population isolation, leading to declines in fish populations and
ultimately to local or total extinction, especially for migratory fish species (Duponchelle etal.,2021; Rodeles
etal.,2017; van Puijenbroek etal.,2018). Over the past 50 years, freshwater migratory fishes have declined by
96% globally, becoming the vertebrate group suffering the most severe decline (Deinet etal.,2020). Studies
concerning the impacts of river fragmentation on migratory fish have mainly analyzed the life history of certain
migratory species, measured their habitat characteristics at different life stages, and assessed the potential impacts
of damming on their habitats (Goodwin etal.,2014; Liermann etal.,2012; Wofford etal.,2005). River frag-
mentation also leads to fish populations becoming isolated and genetically fragmented, exposing them to severe
effects of genetic drift and inbreeding (Brinker etal.,2018; Cheng etal.,2015). Moreover, river fragmentation
changes the structure of riverine food webs, reducing the taxonomic diversity of fish (Freedman etal.,2014).
The beta diversity of native fishes in pools and non-native fishes in riffles decreased with an increase in the ratio
between the length of the longest non-fragmented sections in the river network and the total length of the river
network (Díaz etal.,2020). The decreased lateral connectivity due to dam construction causes loss of access
for some fish species to their spawning and rearing grounds in floodplain and wetlands (Figure2c) (O'Mara
etal.,2021). In the Amazon River, decreases in river-floodplain interactions due to dam construction have caused
significant reduction in catch-per-unit-effort and shifts in the functional composition of fisheries in the Madeira
River floodplain (C. C. Arantes etal.,2021). In addition to fragmentation effects of large dams, widespread small
dams also impose significant, particularly cumulative, impacts on fish distribution and diversity (Consuegra
etal.,2021). Studies show that two-thirds of 191 migratory fish species in Brazil would be affected by river
fragmentation due to the increasing number of SHPs, which is greater than the connectivity loss caused by LHPs
(Couto etal.,2021).
The impacts of river fragmentation on fish diversity and distributions at larger scales (e.g., continental and global
scales) may differ from that at local scales (e.g., basin and sub-basin scales), as fish beta diversity shows signifi-
cant differences among river networks with similar degrees of connectivity (Díaz etal.,2020). In recent years, the
effects of dam-induced river fragmentation on fish, and the mechanisms by which fish diversity responds to river
fragmentation at large scales, have become new areas of interest. Grill etal.(2015) assessed the effects of river
fragmentation on fish habitat at high spatial resolution from sub-basin to watershed scales. Given current progress
in river connectivity studies from watershed to continental scales (Belletti etal.,2020; Duarte etal.,2021; Grill
etal.,2019), it is anticipated that research concerning the impacts of river fragmentation on fish diversity and
distributions will quickly expand to global scales.
Most current studies concerning river fragmentation adopt physical connectivity indices (e.g., CSI, CI); however,
the hydrological and hydraulic disconnections of dammed rivers should receive sufficient attention as they
impose significant impacts on fish habitats. In addition, the impacts of cascade dams on river connectivity and
their cumulative effects on fish are typically much greater than a single dam, and thus deserve more investigation.
At present, indices of physical connectivity are mainly based on geo-spatial data, which may be insufficient in
resolution or possess inadequate records for small dams. As such, the impacts of small dams on river connectivity
worldwide are seriously underestimated, and demand urgent attention.
2.2. Alterations to River Hydrological Regimes
Hydrological regime, including variables such as discharge, flow velocity, water depth, and peak flow, plays an
essential role in riverine bio-habitats and ecosystems (Y. Chen etal.,2016). Fish spawning, rearing and wintering
are strongly related to hydrological conditions. For fish species spawning drifting eggs, continuous stimulation
of flow velocity is required, and thus flooding processes can provide favorable conditions for their reproduction
(Young etal.,2011). With sufficient flow velocity, drifting eggs are not susceptible to sinking, which improves
their survival rate (George etal.,2017). River flow can mediate the dispersal of fish eggs, extending their survival
range and promoting the stability of fish communities (Castello & Macedo,2016). The annual rise and fall of
river depth induces lateral exchange of materials between the river channel and floodplain, which extends fish
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
7 of 64
physical habitat and increases baits for them (Castello & Macedo,2016; Paillex etal.,2013). In cold regions,
fish require water bodies with sufficient depth that have warm water in the deeper layers for overwintering (Cott
etal.,2008).
River hydrological regime depends mainly on climatic conditions and possesses distinct features in different
regions (Fiseha etal.,2014). In tropical regions, most rivers have a large annual runoff and their discharge
possesses strong interannual fluctuations with little seasonality (J. P.Syvitski et al.,2014); some rivers have
well-defined high and low discharges in correspondence to a unimodal rainy period, such as the Ganges River in
south Asia and the Purus River in the Amazon basin; some rivers have two peak flows per year in correspond-
ence to bimodal rainy periods, such as the Magdalena River in South America and the Congo River in Africa
(Latrubesse etal.,2005; J.P.Syvitski etal.,2014). In subtropical and temperate regions, rivers usually have a
relatively smaller annual runoff than those in tropical regions, and their discharges have distinct seasonality that
is influenced by seasonal rainfall and snowmelt, such as the Yangtze River and Yellow River in eastern Asia (W.
Yang etal.,2020). In cold regions, rivers are characterized by low flow in winter due to ice cover and high flow
in spring due to ice-jam floods (Peters etal.,2014; Prowse,2001), such as the Mackenzie River in the North
America and Lena River in Russia. In some rivers, peak flows are dominated by snowmelt in spring (from March
to May), and these rivers are mainly distributed in southern North America, eastern Europe and westernmost
Asia; peak flows may appear near June due to relatively late snowmelt (Hansford etal.,2020), and these rivers
are located mainly in regions of low altitude and mid-high latitude (Q. Liu etal.,2022). Peak flows can also be
brief and intense as determined by monsoon rains, and these rivers are mainly distributed in eastern and southern
Asia, and eastern Australia (Dettinger & Diaz,2000). In some rivers, there are no significant peak flows due to
stable rainfall or mixed climates, such as the St. Lawrence River in central-eastern North America and rivers in
southern Finland (Haines etal.,1988).
Dam construction may significantly alter river hydrological regimes (Timpe & Kaplan,2017). The conversion
from lotic river to lentic reservoir upstream of a large dam fully modifies the original hydrological regime.
Dams directly decrease river flow velocity (Y. Yang etal.,2017). Stevaux etal.(2009) reported that the annual
mean flow velocity of the Paraná River in South America was 0.88m/s in the pre-dam period, and decreased
to 0.56m/s after dam construction. Reservoirs impound water in wet seasons and release it in dry seasons, thus
decreasing the seasonal variability of discharges in rivers (Figure3a). The discharge of the Mekong River in the
dry season is now 63% higher than that in the pre-dam period, while the discharge in the wet season has declined
by 22% (Chong etal.,2021). Although run-of-river reservoirs do not alter the seasonal pattern of discharge
(Figure3b), they can significantly increase the variability of diel or daily discharge through hydropeaking oper-
ations (R. M. Almeida etal.,2020). River damming decreases the number and duration of peak flows, and alters
the frequency of water level variability (Timpe & Kaplan,2017). After the construction of the Gezhouba Dam
and Three Gorges Dam, the number of flow pulses in the downstream Yangtze River decreased by 22%, and their
maximum duration decreased from 16days to 4–6days (Y. Wang etal.,2016). River damming decreases the
maximum discharge and increases the minimum discharge of river, resulting in reduced water level fluctuation
zone (Poff etal.,2007). River damming may also reduce the extent of active floodplain, the period and duration
of flooding, as well as the exchange of materials between main channel and floodplain (Jardim etal.,2020; Moi
etal.,2020). Dam construction in the Balonne River in Australia has resulted in a 23% loss of active floodplain
area and decreased the availability of nutrients from the floodplain (Thoms,2003). To quantify alterations to river
hydrological regime incurred by dam construction, Richter etal.(1996, 1998) have developed the Indicators of
Hydrologic Alteration (IHA) method (Table1), which is widely used to evaluate alterations to the hydrological
regime of dammed rivers. Based on the IHA, the range of variability approach (RVA) has been developed to
evaluate the degree of alteration within specific ranges (Richter etal.,1997).
Alterations to hydrological regime caused by river damming could detrimentally impact the spawning, migration
and feeding behavior of fish (Mims & Olden,2012). A decrease in flow velocity caused by river damming has
been found to reduce the stimulation for fish spawning, resulting in declines of fish reproduction (Figures3c
and3d). In dammed rivers, fish gamma diversity, which indicates the regional or total diversity of fish, is nega-
tively correlated to magnitude in flow velocity (Jarzyna & Jetz,2016; McGarvey & Terra,2016; Timpe &
Kaplan,2017). For fish that fertilize in vitro, the decelerated flow of dammed rivers can reduce the success
rate of fertilization (Campton,2004). Decreases in flow velocity may lead to sinking of drifting eggs or failure
to reach their destination for successful hatching (George etal.,2017). For fish species spawning sticky eggs,
artificial hydropeaking incurred by reservoir operation can cause unsuitable conditions for their reproduction
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
8 of 64
(Vilizzi,2012). The reduced seasonal variability of discharge in dammed rivers can decrease floodplain produc-
tivity, which leads to poor growth and low survival of the fry, resulting in reduction of fish populations (Ficke
etal.,2007; Reinfelds etal.,2013; Terra etal.,2010). Meanwhile, the degradation of plants in floodplain areas
reduces the matrix for sticky eggs to attach, leading to a decrease in the survival of the eggs (Perna etal.,2012).
The decrease in high and low flow pulses and their duration due to river damming can eliminate the hydrolog-
ical cues for fish migration, which reduces foraging opportunities and increase the risk of fish stranding (Bao
etal.,2022; Reid etal.,2019). Rapid reductions of river water surface elevation due to dam operation may result
in stranding of fish species (E. Bell etal.,2008; Irvine etal.,2015). Such fish stranding has been reported during
flow reductions downstream of dams, leading to mortality of salmonids and sturgeon (Johnston etal., 2020;
Nagrodski etal.,2012). Stranding is mainly caused by changes in flow magnitude, known as ramping rates in
dam operations (Le Coarer etal.,2022; Poff etal.,1997). However, the potential for fish stranding also depends
on factors such as fish species and their life stages, stream temperature, and the time of the day, and thus fish
stranding may be site specific (Auer etal.,2022; Benjankar etal.,2023; Glowa etal.,2022). In newly constructed
reservoirs, the decomposition of organic matter can provide more bait, and thus omnivorous fishes can increase
in the short term (Bunn & Arthington,2002; Junk etal.,2013). The increase in water depth and alteration of the
substrate textures in reservoirs reduce the suitable habitat for periphytic algae and benthic macroinvertebrates
(Holt etal.,2015; Taniwaki etal.,2013). This affects the fish feeding on periphytic algae and benthic macroin-
vertebrates, and increases the number of fish preferring phytoplankton and zooplankton, thereby altering the fish
community structure in dammed rivers (W. Zhang etal.,2020). In winter, the rise of water depth in dammed
rivers buffers the drop of water temperature, increasing the chances of fish survival during overwintering (Fuchs
Figure 3. (a) Hydrograph of the Pingshan hydrological station (averaged monthly discharge during 2007–2010, free-flowing)
and the Xiangjiaba hydrological station (replacement of Pingshan hydrological station, averaged monthly discharge during
2016–2020, regulated by a large reservoir) of the upper Yangtze River. (b) Hydrograph of the Ningnan hydrological station
of the Heishui River, a tributary of the upper Yangtze River, before (2015) and after (2019) the removal of the Laomuhe
Dam (a small hydropower dam with no reservoir regulation capacity). (c) Discharge from the Three Gorges Reservoir (TGR)
and fish egg density of four major Chinese carps (FMCCs) measured at the Yidu cross-section of the Yangtze River during
the ecological operation of the TGR in 2016. (d) The number of annual spawned eggs of FMCCs measured at the Yidu
cross-section of the Yangtze River before and after the operation of the TGR. The data of panel (a) and (b) are available from
Hydrological Data of Changjiang River Basin in Annual Hydrological ReportP.R. China, and the data of panel (c) and (d)
are from the authors.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
9 of 64
etal.,2021; Keefer etal.,2008). In summary, alterations to river hydrological regime caused by dam construction
impact significantly on fish habitats, further affecting the population and diversity of fish in dammed rivers.
Existing studies mainly focus on the impact of altered hydrological regime on the spawning aspects of fish
reproductive biology. However, adequate attention should be paid to the impact of altered hydrological regime on
the hatching processes of spawned eggs in dammed rivers, which also plays an essential role in early-stage fish
resources. The reduced water levels during the flood season diminish material exchanges between channels and
floodplain, which reduces feeding field and food abundance of fish. In addition, most available studies investigate
the behavioral response of fish to hydrological alterations, and it is imperative to better understand the physiolog-
ical mechanisms of fish reactions to dam-induced hydrological alterations.
IHA statistics group Regime characteristics Hydrologic parameters Examples of river ecosystem influences
Magnitude of monthly discharge
conditions
Magnitude Mean discharge for each calendar month Habitat availability for fish
Timing Influences water temperature, oxygen
levels, and photosynthesis in water
Magnitude and duration of annual
extreme discharge conditions
Magnitude Annual minima (1-day means; 3-day means;
7-day means; 30-day means; 90-day means)
Balance of competitive, ruderal, and
stress-tolerant fish
Structuring of river ecosystems by abiotic
versus biotic factors
Duration Annual maxima (1-day means; 3-day means;
7-day means; 30-day means; 90-day means)
Structuring of river channel morphology
and physical habitat conditions
Number of zero-flow days Volume of nutrient exchanges between
rivers and floodplains
Duration of stressful conditions (e.g., low
oxygen, concentrated chemicals) in
river ecosystem
Seven-day minimum flow divided by mean flow
for year (base flow)
Duration of high flows for aeration of
spawning beds in river sediments
Timing of annual extreme discharge
conditions
Timing Julian date of each annual 1-day maximum
discharge
Compatibility with life cycles of fish
Predictability/avoid ability of stress for fish
Access to special habitats during
reproduction or to avoid predation
Julian date of each annual 1-day minimum
discharge
Spawning cues of migratory fish
Evolution of life history strategies and
behavioral mechanisms
Frequency and duration of high/low flow
pulses
Magnitude Number of high pulses each year Availability of fish habitat in floodplain
Nutrient and organic matter exchange
between rivers and floodplains
Frequency Number of low pulses each year Soil mineral availability
Duration Mean duration of high pulses within each year Water birds enter foraging, resting, and
breeding places
Mean duration of low pulses within each year Bed load transport, channel sediment
structure, and substrate disturbance
duration (high pulses)
Rate/frequency of hydrograph changes Means of all positive differences between
consecutive daily values
Entrapment of fish on islands and
floodplains (rising levels)
Frequency Means of all negative differences between
consecutive daily values
Rate of change Number of flow reversals Desiccation stress on low-mobility stream
edge fish
Table 1
Summary of Hydrologic Indicators and Their Ecological Influences (Richter etal.,1998)
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
10 of 64
2.3. Changes to River Sediment Regimes and Morphology
Sediment is an essential component in rivers, and plays a pivotal role in maintaining the ecological status of global
river systems (Chapman & Wang,2001; Förstner etal.,2004,2008; Netzband etal.,2007; Ralph etal.,2009).
River sediment is entrained, transported and deposited, thus shaping river morphology and affecting fish habitat
(Nichols,2009). The balance between sediment supply and transport capacity of a river system is a fundamental
driver for river geomorphology, which not only dictates the aggradational or degradational state of a system,
but also controls channel morphology and substrate textures (Dietrich etal.,1989; Lisle etal.,1993; Pitlick &
Wilcock,2001). River planform can adopt a range types, form straight to sinuous, braided, anabranching or anas-
tomosing shapes (Latrubesse,2008; Leopold & Wolman,1957). These different river planforms induce different
hydrodynamic conditions, which in turn cause different patterns of sediment erosion and deposition, forming a
variety of geomorphic units such as riffles, pools, barforms and bedforms, and flood plains, which thus increase
the diversity of biological habitats, for example, those used for fish spawning and wintering grounds (Chapuis
etal.,2015; Namour etal.,2015). Rivers also transport large quantities of organic matter, providing food sources
for aquatic organisms (Karr,1991). The content of organic matter in sediment is affected by sediment character-
istics, including particle size and density, surface site density, and particle morphology (Y. Wu etal.,2020). In
most rivers, bed sediments show an overall downstream-fining trend (Luo etal.,2012), as illustrated in Figure4a.
Sediment from global rivers delivered to the oceans was estimated to be about 20 Bt/yr, prior to significant dam
interception (Milliman & Syvitski,1992). Dams alter the natural balance of sediment flux in rivers by trapping
sediment in reservoirs, discharging waters often free of sediment downstream (Morris & Fan, 2010). Global
dams decreased about 5Gt/yr of sediment flux to the oceans by the 21st century (Milliman & Syvitski,1992).
Upstream of a dam, the reduction in flow velocity may facilitate sediment deposition in the river channel and
floodplain (Fencl etal.,2015; X. Su etal.,2017; Walter & Merritts,2008). As shown in Figure4b, coarse parti-
cles, such as gravel and coarse sand, are the first to settle, forming a delta at the point where the backwater effect
ends; fine sediment particles enter the reservoir and are transported by turbid density currents or non-stratified
flow, and may be deposited near the dam (Fan & Morris,1998; Garde & Raju,1979). In addition, the peak
sediment load in dammed rivers is separated in time from the maximum water flow (Dang etal.,2010; Topping
etal.,2000), which also promotes the deposition of sediment. In the river downstream of a dam, the reduction
in sediment content usually leads to channel incision, chronic erosion of the bed and banks, and even the loss of
delta plain (Bittencourt etal.,2007; Graf,2006; Magilligan & Nislow,2005; Petts & Gurnell,2005). Intensified
scour and floodplain incision are often observed in sediment-starved rivers, as the flows entrain bed material
equaling to their transport capacity (Csiki & Rhoads,2010). In some cases, dams reduce the number of intense
floods and lead to sedimentation downstream, which raises the river bed (Kotti etal.,2016; Słowik etal.,2018).
After damming, pools may occur more frequently and individual pools are longer in the lower reach than that in
Figure 4. Impact of river damming on river bed sediment. (a) Bed sediment in free-flowing rivers. (b) Bed sediment in dammed rivers.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
11 of 64
the upper reach; in contrast, individual riffles in the upper reach are longer than that in the lower reach (Kobayashi
et al., 2012). Low sediment loads downstream after dam construction also leads to reductions in associated
nutrient transport, and thus affects fish feeding grounds (C. Guo etal., 2020). Moreover, the balance between
inorganic and organic sediment is disrupted by damming, with mineral particles deposited predominately in the
reservoirs, and the increased biological production causing the suspended load of the reservoir outflow to be
largely composed of organic matter from aquatic organisms (Sokolov etal.,2020).
Alterations to sediment regimes due to river damming have potential impacts on fish habitat. The changes to
erosion and deposition patterns, and hence the river morphology, may impede suitability for fish spawning and
overwintering grounds (Kruk & Penczak,2003; McLaughlin etal.,2006). For resident or potamodromous fish,
river damming changes the number and distribution of fish habitats for spawning, feeding and overwintering
downstream, leading to increases in competition for spawning and wintering sites as well as food resources
(Cambray etal.,1997). The changes of mesoscale riverbed morphology, such as riffles and pools in the upstream
and downstream, after river damming can indirectly lead to changes in the diversity and distribution of fish
communities (Calderon & An,2016; Langeani etal.,2005). In addition, different fish species may have a differ-
ent preference for sediment properties. For example, larvae of white sturgeon (Acipenser transmontanus) prefer
substrates of clean gravel and cobbles (Nguyen & Crocker,2006), while larvae of Schizothorax wangchiachii
change their preference for substrate at different life stages (Chai etal., 2019). Typically, lamprey spawn in
riverbeds covered with a mixture of sand, gravel, and pebbles (N. S. Johnson etal.,2015). Fine sediment with
relatively high organic matter is a primary source of food and energy for some species, and can even be an inte-
gral requirement within the lifecycles of species such as lamprey ammocete and psammophilous fish. Sediment
coarsening downstream of a dam would result in less bait for these fish (Maitland,2003). Moreover, complex
habitat structure, such as pore space between coarse sediment, can increase predator-free space and thus reduce
predation efficiency, an effect that is most pronounced at low prey densities (Barrios-O’Neill etal.,2015,2016;
Toscano & Griffen,2013). Therefore, the altered river morphology and sediment gradation due to dam construc-
tion may seriously impact fish habitat. The alteration of suspended sediment content can also cause a variety of
effects on fish habitat in dammed rivers. Some fish species prefer turbid over clear water, presumably benefiting
from a reduced risk of predation and increased opportunity of feeding (Cyrus & Blaber,1987,1992). As reser-
voirs convert flowing water to relatively still water, the upstream water changes from turbid to limpid (C. Guo
etal.,2020), and the increase in water clarity can directly affect the habitat of these fish species. In summary,
the alterations of both suspended and bed load sediment regimes have important implications for the whole life
history of fish in dammed rivers.
Despite extensive studies concerning the impacts of sediment regime alterations induced by river damming on
fish, many challenges remain. The effects of changes in transparency and bed color on fish communities after
river damming have not been explored sufficiently. Quantification of the contributions from direct effects, such
as turbidity changes, and indirect effects such as predation behavior changes, also demand investigation. To date,
most studies have focused on the response of fish to alterations in sediment regimes, but neglected the effects
of alterations in fish activity on river sediment and morphology. It has been reported that spawning mucus of
some fish species can modify substrate characteristics and thus affect sediment transport, which highlights the
significance of spawning as a zoogeomorphic activity (Roberts etal.,2020). The effect of changes in spawning
grounds of these fish species on river morphology after dam construction is a topic that demands further study.
2.4. Alterations in River Water Temperature Regimes
Water temperature is an important and highly sensitive factor in river ecosystems, and possesses distinct and
regular seasonality. The natural rhythm of water temperature affects the phenological functioning of aquatic
species, and the longitudinal variation of water temperature plays an essential role in generating spatial patterns
of species communities (Isaak etal.,2012). Water temperature regime can affect the whole life cycle of fish
(Servili et al.,2020), including migration timing, reproductive performance, embryo health, and growth rate
(Figure5c). For migratory fish, the change of water temperature within a suitable range is one of the key envi-
ronmental cues for migration (Harvey etal.,2020; Rijnsdorp etal.,2009). The Atlantic salmon (Salmo salar)
smolts in the three tributaries of the West River in Vermont, USA, begin migration when the water tempera-
ture rises to 5°C, and reach peak migration when water temperature rises to 8°C (Whalen etal.,1999). Fish
have specific reproduction strategies, and a suitable temperature promotes healthy egg development and ensures
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
12 of 64
successful spawning (Kurita etal.,2003; Sieiro etal.,2020). For temperate or tropical fishes that spawn in spring
and summer, warm winters could affect the accumulation of nutrients during egg development, resulting in
smaller eggs (Collingsworth etal.,2017; T. M. Farmer etal.,2015). For most cyprinids in temperate and tropical
climates, temperature acts as a time clue for reproduction events, such as ovulation and oviposition (Pankhurst &
King,2010; N. Wang etal.,2010). In addition, changes in plankton and benthos in response to variations in water
temperature directly affects fish food chains, and thus indirectly regulates the energy intake and growth of fish
(Pörtner & Farrell,2008; Prokešová etal.,2020).
The spatio-temporal variation of river water temperature is influenced by multiple factors, including water
flux, climate, latitude and human activity (Caissie, 2006; Collins, 2009; Markovic etal.,2013; B. W. Webb
etal.,2008). The study on a Cairngorm stream in northeast Scotland (Hannah etal.,2004) shows that in terms
of average energy flux affecting water temperature, the main heat sources are sensible heat (38.7%), bed heat
(37.0%) and friction at stream bed and banks (24.3%), while the main heat losses are latent heat (73.1%) and
longwave radiation (26.9%). Typically, the water temperature of headwater streams is often lower than that of
downstream reaches, because many rivers originate in plateaus or snow-covered mountains with water sourced
from ice melting (Cai etal.,2018; Dugdale etal.,2017; Moors etal.,2011). At the global scale, regional trends
of river water temperature are consistent with air temperature (J. Syvitski etal.,2019). In addition, river water
temperature is increasingly affected by climate change (Dugdale etal.,2017; van Vliet, Franssen, etal.,2013; van
Vliet, Ludwig, & Kabat,2013). The response of river water temperature to climate change varies with latitude.
Increases in water temperatures at low latitudes are generally greater than those in high and middle latitudes.
However, water temperature changes in rivers in high-altitude regions, such as the Qinghai-Tibet Plateau (S. Liu
etal.,2020), are different from this broad latitudinal pattern.
River damming can significantly alter water temperature regimes, as illustrated in Figure5a (D. Cheng etal.,2015;
Jung etal.,2022). The alteration of water temperature regime by reservoirs can be influenced by various factors,
including the shape, storage and depth of the reservoir, the inflow water temperature, the hydraulic residence
time, and the operation modes of the reservoir (Lessard & Hayes,2003; Prats etal.,2010; Wotton,1995). The
Figure 5. (a) Water temperature of the Pingshan hydrological station (averaged monthly water temperature during
2007–2010, free-flowing) and Xiangjiaba hydrological station (replacement of Pingshan hydrological station, averaged
monthly water temperature during 2016–2020, regulated by the large Xiangjiaba Reservoir) of the upper Yangtze River.
(b) Water temperature stratification in free-surface (left) and frozen-surface (right) reservoirs. (c) Water temperature in
downstream of the Xiangjiaba Dam on the upper Yangtze River before and after the impoundment of Xiangjiaba Reservoir,
and the phenological properties of fish (OW: overwintering, MG: migration, SP: spawning, GD: gonadal development,
GR: gonadal recovery). Data are available from T. Li etal.(2021). (d) Impact of altered water temperature regime on fish
spawning (CTT: critical temperature threshold for spawning, ATT: accumulated temperature threshold for gonad maturity).
The data of panel (a) are available from Hydrological Data of Changjiang River Basin in Annual Hydrological ReportP.R.
China, and Figure 5d is adapted from T. Li etal.(2021).
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
13 of 64
essence of water temperature alteration by reservoirs is that the impounded water is thermally stratified season-
ally and the outflow is mostly from deep layers. Although the impact of reservoirs on river water temperature
regime is complex, the degree of alteration depends principally on dam height and regional climate characteristics
(Maheu etal.,2016b), as summarized in Table S1 in Supporting InformationS1. Reservoirs in different climatic
regions have different effects on the water temperature regimes of dammed rivers (Pieters & Lawrence,2012; D.
Yang etal.,2005). In tropical, subtropical and temperate regions, large reservoirs usually form thermal stratifica-
tion in the spring, summer and autumn, with high temperatures in the surface layer and low temperatures in the
bottom layers (Figure5b). In cold regions, large reservoirs have thermal stratification in freezing periods, mostly
in winter and early spring, with low temperatures in surface layers but relatively high temperatures in the bottom
layers (Figure5b). The seasonal stratification of reservoir waters can, to some extent, alter the natural rhythm
of water temperature downstream (Figure5a), which are characterized by elevated water temperatures in winter,
decreased water temperatures in spring and summer, and reduced amplitude of maximum water temperatures
(Long etal.,2019; Soleimani etal.,2019). In particular, the largest alterations in downstream water temperature
occur in dry seasons (Maheu etal.,2016a). Because of the surface layer discharge mode and non-stratification,
the effects of small dams on river water temperature are different from that of large dams (Maheu etal.,2016b).
The release of surface water from small dams in spring and summer tends to elevate downstream water tempera-
ture (B. W. Webb etal.,2008), which is opposite to water temperature alterations in the summer period caused by
large dams (Skoglund etal.,2011). To quantify the alteration to water temperature in dammed rivers, a variety of
indicators have been proposed, such as changes in suitable water temperature range of target species, and changes
in maximum and minimum water temperatures (T. Li etal.,2021; Maheu etal.,2016a).
Alteration of water temperature regime by dams has serious impacts on river fish habitat (Ahmad etal.,2021;
Couto etal.,2021; Grill etal.,2019; Kuczynski etal.,2017; Prats etal.,2010), which directly affect the spawning,
migration and growth of fish (Figure5c). Fish reproduction is the most vulnerable and sensitive to water tempera-
ture alterations. Due to the operation of the Xiluodu and Three Gorges dams on the upper Yangtze River in China,
the elevated water temperatures in autumn inhibit the reproduction activities of Chinese sturgeon (Acipenser
sinensis) that usually spawn in autumn, causing the effective breeding quantity to reduce to 0%–4.5% (Z. Huang
& Wang,2018). The warming of water during winter due to reservoir operation in northern Europe and North
America leads to delayed timing of spawning and shortened time of egg incubation for winter-spawning Atlantic
salmon and brown trout (Salmo trutta) (Bohlin etal.,1993; Heggenes etal.,2018; Jonsson & Jonsson,2009). In
addition, a phenological knock-on effect could emerge as a consequence of delays in spawning time (Elliott &
Elliott,2010; Skoglund etal.,2011), resulting in low survival rates of Atlantic salmon and brown trout larvae.
In the upper Yangtze River, the ray-finned fish (Coreius guichenoti) spawns in late spring and early summer
when the critical water temperature exceeds 20°C. The timing of the arrival of this critical temperature is delayed
due to the operation of the Xiluodu Reservoir (T. Li etal.,2021). Meanwhile, the warming of water in winter
accelerated the development of gonads and advances the time of egg maturity. The joint effects of warmed water
in winter and cooled water in early summer thus cause ray-finned fish to miss the historical time window for
spawning, resulting in the sharp decline of its population (T. Li etal.,2021; Z. Yang etal.,2021).
The critical temperature threshold (CTT) for fish spawning and the accumulated temperature threshold (ATT) for
fish gonadal development have been used to quantify the impact of water temperature alteration on fish repro-
duction (Figure5d). The ATT is the minimum accumulated temperature required for gonadal maturity, which is
calculated by summing the temperatures higher than the ontogeny temperature in days during gonadal develop-
ment (Chezik etal.,2014; Honsey etal.,2018). For example, the CTT and ATT thresholds for bronze gudgeon
(Coreius heterodon) reproduction in the upper Yangtze River were about 18.4 and 1,324.9°C·day, respectively.
Operation of large reservoirs results in cooling in spring and summer, and warming in autumn and winter (Long
etal.,2019; Y. Tao etal.,2020), which delays the timing for reaching CTT and advances the timing for reaching
the ATT (T. Li etal.,2021). It has been shown that elevated temperatures of 2–3°C in winter can have a signif-
icant impact on the development of fish gonads (T. M. Farmer etal.,2015; T. Li etal., 2022). When the time
to reach the ATT is advanced, fish tend to consume energy in the yolk in order to maintain early exercise and
strengthen metabolism (Mcqueen & Marshall,2017; Wright etal.,2017). When the time to reach the CTT is
delayed and becomes behind the time for reaching the ATT, temperature-sensitive fish cannot spawn in a timely
manner, resulting in over-maturity of their gonads and thereby a decrease in reproduction (Kennedy etal.,2011;
J. King etal.,1998). At present, most studies use the time to reach the CTT as the indicator for reservoir operation
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
14 of 64
management (Y. Tao etal.,2020; Y. Wang etal.,2016); however, re-matching the time for reaching both the CTT
and ATT is key to mitigate the impact of water temperature alterations on fish in dammed rivers.
Despite many studies addressing the effects of river damming on water temperature regimes, there is a great
knowledge gap between alterations in water temperature and their impacts on fish, particularly on the whole
life cycle of fish migration, growth and food chain. The limited number studies that document the impacts of
water temperature alterations on fish physiology, such as sex determination and gonad development, have mostly
been conducted at extreme temperature conditions (C. Li etal.,2014; Ribas etal.,2017) instead of natural
temperature conditions (Dorts etal.,2012). In addition, understandings to the long-term impact of water temper-
ature alterations on fish phylogeny are insufficient. In the future, climate change may increase the frequency of
extreme weather events, which could intensify or offset the impacts of reservoir operations on water temperature.
Therefore, the combined effects of river damming and climate change upon water temperature regimes and fish
communities demand further investigations.
2.5. Dissolved Gas Supersaturation
Total dissolved gas (TDG) supersaturation refers to the physical condition that the sum of dissolved gas partial
pressures in water exceeds the sum of gas partial pressures in the air under local atmospheric pressure (Weitkamp
& Katz, 1980). In natural rivers, TDG supersaturation is often caused by a change of water temperature
(Bouck,1984) and water falling (Rowland & Jensen,1988). Dammed rivers can occur severe TDG supersatura-
tion caused by flood discharge during flooding seasons (Figure6a), resulting in serious impacts on fish down-
stream (Ji etal.,2019; Pulg etal.,2016). For instance, the Chief Joseph Dam released flow in excess of the 110%
TDG compliance criterion (EPA,1987) in the Columbia River, causing the downstream Wells Dam to be unable
to comply with water quality criteria for 125 of 133 compliance-mandated days in 2012 (Witt etal.,2017). During
the flood discharge of the Xiluodu Reservoir in the upper Yangtze River in July 2014, the average TDG saturation
of the downstream water was 135%, with a maximum value of 144%, resulting in massive fish mortality in the
downstream reservoir (Q. Ma etal.,2018; Zeng etal.,2020). The release of TDG is slowed with increases in
water depth and decreases in flow velocity. Thus, in a reservoir cascade system, water with supersaturated TDG
from the upstream reservoir is transported to the downstream reservoir and can cause cumulative effects (Q. Ma
etal.,2018). TDG supersaturation effects are thus particularly prominent in reservoir cascade systems.
The generation of TDG supersaturation is mainly affected by water pressure, water temperature, bubble pressure,
bubble retention time, and aeration concentration. The saturation solubility of gas is positively related to water
pressure (J. Feng etal.,2018). When a large volume of water with high head is released from a dam, the water jet
entrains a large amount of air to form aerated water, and drops deeply into the energy dissipation pool near the
dam (Bertola etal.,2018). Due to high water pressure in the pool, the air carried by the water jet dissolves rapidly,
forming TDG supersaturation in the water (Pulg etal.,2016; H. Xue etal.,2018). The degree of TDG super-
saturation in the stilling pool is positively correlated to the bubble retention time (Lu etal.,2019; Y. Peng, Lin,
etal.,2022). The initial saturation of high dam release has little impact on the generation of TDG supersaturation
in the stilling pool, as the water jet is largely saturated after undergoing the process of full gas mixing during the
downward discharge (S. Xue etal.,2019). The influence of aeration on the stable saturation of the water body
Figure 6. Impact of total dissolved gas (TDG) supersaturation on fish. (a) TDG supersaturation due to reservoir flood
dispatch causes fish gas bubble disease. (b) Relation between TDG saturation and half-lethal time of Ya-fish (Schizothorax
prenanti). The data are available from Y. Wang etal.(2015).
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
15 of 64
downstream of the dam is also marginal, since the amount of aeration carried by the water jet into the stilling pool
is usually sufficient (Qu etal.,2011).
TDG supersaturated water gradually releases the dissolved gas during its flow downstream, with the release rate
influenced principally by water depth, flow velocity, wind speed and water temperature (C. Cao etal.,2020; Ou
etal.,2016). The release rate of TDG decreases with increase in water depth and decrease in flow velocity. Obser-
vations downstream of the Three Gorges, Ertan and Zipingpu dams, China, showed that the release rate down-
stream of the Zipingpu Dam is the highest, which corresponds to its lowest water depth (J. Feng etal.,2010).
Compared to downstream, the upstream side of a dam has larger water depths and slower flow velocities, result-
ing in lower release rates, as observed in the Kootenay River, North America (Kamal etal.,2018). Increase in
wind speed enhances gas transfer at the gas-liquid interface, which accelerates the release of TDG in the water
(Chu & Jirka,2003; J. Huang etal.,2016). Elevated water temperatures promote the thermodynamic movement
of gas molecules, which can also increase the release of TDG (S. Liu,2013; Ou etal.,2016). Suspended particles
and water plants provide nucleation and aggregation sites for dissociative gas molecules, enhancing free gas
molecules in the water to concentrate and escape in the form of bubbles, and thus promote the release of TDG (J.
Feng etal.,2012; Y. Yuan, Huang, etal.,2018).
When TDG saturation exceeds the sum of atmospheric pressure and hydrostatic pressure, gases in the dissolved
state in fish tissues and body fluids will precipitate and accumulate to form gas bubbles (Figure6a). Gas bubbles
often appear in the tissue of the head, mouth, fins, and gill arches of fish, or in the capillaries of the gill plates
(Lemarie etal.,2011), resulting in gas bubble disease (GBD). GBD could affect the physiology and behavior of
fish, such as loss of balance and abnormal buoyancy, leading to fish death in severe cases (Bouck,1980; J. Huang
etal.,2021). Several fish abnormalities caused by GBD have been reported, such as fast swimming, rapid breath-
ing, back and forth movement, and mouth breathing with sticky bubbles (C. Feng etal.,2019; Y. Yuan, Wang,
etal.,2018). Bleeding may also appear in tissues of gills, fins, muscles, gonads, and intestinal epithelium (Meyers
etal.,2008). These symptoms may impose potential damage on fish, such as tissue necrosis, abnormal growth
(Geist etal.,2013), decreased immunity (Schisler etal.,2000), reduced swimming ability (Y. Wang etal.,2017),
increased risk of predation and increased buoyancy (Shrimpton etal.,1990a,1990b), as well as changes in phys-
iological characteristics (Yuan etal.,2021).
The impact of TDG supersaturation on fish is determined by TDG saturation, exposure time, age and species of
fish, swimming depth, and behavioral habits. High TDG saturation and long exposure time increases the mortal-
ity of fish. The half-lethal time of Ya-fish (Schizothorax prenanti) at 150% TDG saturation is 8.5 times shorter
than that at 120% TDG saturation (Figure6b). Repeated exposure of fish to TDG supersaturated water can lead
to reduced feeding ability and increased susceptibility to fungal and bacterial infections, thereby further reducing
fish tolerance to TDG supersaturated water and affecting fish survival (Brosnan etal.,2016; Huchzermeyer,2003;
Schisler etal., 2000). Fish tolerance to supersaturated TDG also varies between species and growth stages.
During the incubation phase of eggs, increased TDG saturation causes a gradual decrease in the hatching rate (N.
Li etal.,2019; R. Liang etal.,2013). The tolerance threshold of TDG saturation for the juvenile fish of Ya-fish
(Y. Wang etal.,2015), silver carp (Hypophthalmichthys molitrix) (Deng etal.,2020), and Chinese sucker (Myxo-
cyprinus asiaticus) (L. Cao etal.,2016) is higher than that for the juvenile fish of rock carp (Procypris rabaudi
Tchang) (X. Huang etal.,2010) and grass carp (Ctenopharyngodon idella) (F. Wu etal.,2020). However, when
TDG saturation exceeds 135%, the tolerance threshold of TDG saturation for fish is not significantly different
between species and body sizes (S. Xue etal.,2019).
With an increase in water depth, the saturation solubility of TDG increases, and the relative saturation of TDG
decreases. The release of TDG is also related to turbulence, and thus TDG supersaturation is unevenly distributed
vertically (P.Li etal.,2022). Therefore, the tolerance of fish to TDG supersaturation is improved with greater
water depths, which is known as the effect of depth compensation (Yuan etal.,2020). Some fish species can avoid
TDG supersaturation stress using the effect of depth compensation. For instance, when TDG saturation exceeds
120%, Ya-fish, Chinese sucker and elongate loach (Leptobotia elongata) have the ability to detect and avoid TDG
supersaturated water in horizontal and vertical directions (Deng etal.,2020). Rainbow trout (Oncorhynchus
mykiss) has a significantly higher TDG exposure risk than mountain whitefish (Prosopium williamsoni), but
refuge habitats with sufficient water depth can mitigate exposure risk and GBD. To alleviate the stress of TDG
supersaturation, rainbow trout uses depth compensation, with each 1.0m increase in swimming depth offsetting
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
16 of 64
9.7% TDG saturation (Pleizier etal.,2020). Therefore, TDG exposure risk and actual risk depend on the interplay
between species-specific ecology and dam-induced TDG patterns (Algera etal.,2022).
TDG supersaturation only occurs occasionally during the discharge of floodwaters, when it is usually risky and
difficult to conduct field surveys (Bertola etal.,2018; Pulg etal.,2016). Therefore, field observation data of
TDG supersaturation are relatively scarce. Also, limited by monitoring technologies, TDG concentration in the
near-field of large dams is difficult to measure well (Algera etal.,2022). Scale models in laboratories possess
substantial difficulties in simulating the processes involved in generation and release of supersaturated TDG in
the field, due to insufficient consideration of the Reynolds and Weber numbers as well as water pressures (P.Li
etal.,2022). Therefore, it remains challenging to model the mass transfer process accurately, and predict bubble
size as well as size distribution in order to support the development of effective mitigation measures. In addition,
the threshold of TDG saturation for fish injury in different tissues requires further investigation. Moreover, fish
adopt an avoidance behavior when TDG saturation exceeds their tolerance threshold, and the tolerance of fish to
TDG supersaturation is enhanced with an increase in water depth. Previous laboratory studies investigating the
response of fish to TDG only consider different TDG concentrations, lacking investigations on depth compensa-
tion effects. Therefore, it is essential to further study the effect of depth compensation, as well as fish physiolog-
ical and behavioral response, in the assessment of TDG supersaturation impacts on fish.
3. River Damming Impacts on Key Fish Species
The interests in the key fish species impacted by river damming vary between different continents. Herein, we
focus on salmonids, Chinese carps, sturgeon, eel and lamprey, which have been intensively studied.
3.1. Impact on Salmonids
Salmonids (family Salmonidae) are one of the most popular commercial fishes, providing high-quality protein to
people around the world, particularly in Europe and the United States (Phillips & Rab,2001). They mainly include
the genera Salmo and Oncorhynchus, which have received most research concerns (Klemetsen etal.,2003). The
genera Salmo contains two typical anadromous species, which are the Atlantic salmon and brown trout. Atlantic
salmon are mainly distributed along both the east and west coasts of the North Atlantic Ocean, while brown
trout is indigenous to Europe, North Africa and western Asia (MacCrimmon etal.,1970; MacCrimmon and
Gots,1979). The genera Oncorhynchus contains the Pacific salmon (Oncorhynchus tshawytscha, named Chinook
salmon; Oncorhynchus keta, named Chum salmon) and Pacific trout (Oncorhynchus mykiss, including rainbow
trout and steelhead trout subspecies), which mainly inhabit the north Pacific and the coastal rivers of both Amer-
ica and Asia (Figure7). The Chinook salmon and Atlantic salmon have received most research attentions due to
the threats of river dams (Harnish etal.,2014; Hilborn,2013; Potter & Crozier,2000).
Figure 7. Distribution of salmonids around the world and their migration route for spawning.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
17 of 64
The anadromous salmonids migrate to freshwater rivers for spawning and juvenile rearing, and to saltwater
oceans for feeding, growing and maturing (Crozier etal.,2021; Groot,1991). The main life cycle of anadromous
salmonids includes a spawning stage (egg, alevin), juvenile stage (fry, fingerling, parr), smolt stage, and adult
stage. In the autumn, female salmonids excavate nest pits (called redds) in river bed gravels and spawn eggs into
them. The optimal spawning temperature for salmonids is in the range of 6–14°C, and the spawning timing is
especially sensitive to temperature changes. Over the winter, the eggs develop into very small salmonids (alev-
ins). In the spring, the alevins swim out of the redds and become fry, with the fry growing into parr that can
protect themselves from predators. The parr grows in freshwater for 2–3years, and transforms into smolts. In
the early spring, silvery smolts swim to the ocean and spend 1–2years maturing into adults. Atlantic salmon
is a cool-water species, and the downstream migrating smolts must swim to the ocean before river tempera-
ture becomes too warm. River discharge plays an important role once migration is under way, and influences
the onset, duration and termination of smolting (Sykes etal.,2009). During the adult stage, salmonids feed
in the open ocean for a certain period of time, which differs between different species. In the summer, adult
salmonids migrate back to the freshwater, strongly following the migratory route they adopted when leaving the
river as smolts (Rivinoja etal.,2001). Increasing flows can facilitate downstream migration of smolts, and flow
augmentation during the upstream migration period can improve the survival of migrating salmonids (Connor
etal.,2003). The upstream swimming capability of salmonids would be reduced at low (below 10°C) and high
(>20°C) temperatures (Alabaster,1990; Johnsen & Jensen, 1994). Therefore, the connectivity between fresh-
water and ocean, water temperature, and river discharge are key factors affecting the physiological behaviors of
anadromous salmonid across their life cycle (Caudill, etal.,2013).
Dams impede both the downstream and upstream migration of salmonids, resulting in the decline of salmonid
populations around the world (Lawrence etal.,2016; Limburg & Waldman,2009). In North America, dams have
been one of the most important factors blamed for the decline of Chinook salmon populations in the Pacific coast
of Oregon and British Columbia (P.H. Wilson,2003). Dams in the upper Columbia River have caused a decline
in the abundance, survival and population of Chinook salmon (Levin & Tolimieri,2001). In South America,
both the catch rate and body size of introduced salmonids decreased in several dammed large rivers in Chile and
Argentina (Arismendi etal.,2019).
Budy etal.(2002) demonstrated that the survival rates of salmonids in dammed rivers decreased mostly in the
smolt-to-adult life stage, rather than in the spawner-to-smolt life stage. When smolts migrate downstream from
the dammed river to the ocean, most of them pass the dams through the spillway, the juvenile bypass system, or
the hydropower turbines. Physiological or behavioral stresses caused by this passage experience on smolts would
result in direct death or a chronic influence, including injury, trauma, diminished physical abilities, and increased
susceptibility to predation and disease, which may eventually lead to death at a later life stage (Budy etal.,2002).
Compared to rivers with a single dam, rivers with cascade dams make juvenile salmonids experience multiple
dam passages, and these cumulative stresses can result in significantly higher mortality (Molina-Moctezuma
etal.,2021). For instance, in rivers with three or four dams, the decline of fish populations can increase to exceed
30% (Lawrence etal.,2016).
When adult salmonids migrate upstream to their natal spawning and rearing habitat, their instinct and imprint of
the downstream passage experience during their juvenile stage leads them to the bypass channel or the turbines,
which increases mortality or results in the situation that they cannot find the correct route, and delays their
migration until they determine the correct route (Rivinoja etal.,2001). On average, there is a 70% loss of poten-
tial Atlantic salmon spawners during their upstream passage at dams in Sweden (Lundqvist et al., 2008). In
Europe, many rivers of the Baltic Sea have lost the natural juvenile reproduction of Atlantic salmon, because river
damming has blocked or reduced the access of adult salmon to their spawning grounds (Rivinoja etal.,2001).
Over several generations, some of the white-spotted charr (Salvelinus leucomaenis) in Japan in dammed-off areas
no longer migrate to the sea and become resident fish due to habitat fragmentation, which leads to a decrease in
their spawning and populations (Morita etal.,2009). Therefore, river fragmentation caused by dams can lead to
the increasingly direct or delayed mortality of salmonids, and thus may severely reduce their population.
Hydrological alterations caused by dam operations may lead to the decline of salmon populations (E. J. Ward
etal.,2015). During the spawning period, the female and male Chum salmon that remain on their redds would
increase swimming activity, reduce digging activity and leave their redds when flow velocities are above a thresh-
old of 0.8m/s. Therefore, artificial hydro-peaks would lead to a decline in spawning rates (Tiffan etal.,2010).
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
18 of 64
A significant reduction in discharge under the dam could lead to the stranding of salmonid fry and increase their
mortality (E. Bell et al.,2008). For example, dam operations result in dewatering of Chinook salmon redds,
causing the mortality of eggs and larval fishes (Harnish etal.,2014; Ostberg & Chase,2022; Young etal.,2011).
Regulated flows, even if they do not dewater the redds, can still impact on redd habitats for embryos survival and
development by altering hyporheic exchange of heat and dissolved oxygen (Figure2d) (Bhattarai et al.,2023;
Martin etal.,2020). The upstream migration of adult salmonids may be disrupted by the seasonal and weak
variation in spillway discharge from the dam (Rivinoja etal.,2001). Therefore, alterations to hydrological regime
caused by dams may result in declines in spawning rate, survival of fish eggs and juveniles, and migration of adult
salmonids, thus decreasing the populations of salmonids.
Dam construction can cause the historical spawning grounds of salmonids to become warmer, which makes the
hatchery juveniles progress their life history at an earlier time and thus swim to areas with relatively cool water
temperatures and lower growth opportunities, leading to a decline in their survival and populations (Connor
etal.,2002). However, rising water temperatures in springtime caused by dam operation can negatively affect
the rate of salmonid egg incubation, such as Chinook salmon (Dusek Jennings & Hendrix,2020). Compared to
the daily mean temperature or threshold temperature, temperature experience (accumulated thermal units, ATU)
has more effects on the onset, duration and termination of downstream migration of Chinook salmon smolts
(Sykes etal.,2009). Stich etal.(2015) found that dams decreased the ATU for smolts, thus delaying the time to
initiate their downstream migration. In addition, dams as barriers along migration routes can delay or prolong
the migration of smolts, which may result in a mismatch with migration-timing adaptations. For instance, water
tempera tures in downstream reaches become warmer in the spring due to dam operations, which delays the down-
stream migration of smolts of Atlantic salmon. Consequently, the delayed smolts may face a situation that the
water temperature further increases to a lethal or near-lethal level for smolts (Marschall etal.,2011). Compared
to fish in free-flowing rivers, upstream-migrating fish in impounded systems could encounter potential thermal
barriers in fishways. For instance, fishways offer opportunities for upstream migration of adult Chinook salmon,
but the unfavorable water temperature gradient in the fishway caused by thermal stratification in reservoirs
presents an obstacle to the upstream migration of individuals (Caudill etal.,2013). Therefore, the altered water
temperature regime caused by dams negatively affects egg hatching, the downstream migration of smolts, and the
upstream migration of adult salmonids in fish passages.
The TDG supersaturation caused by flood discharge of dams has little impact on juvenile and adult salmonids
(Geist etal.,2013; Muir et al., 2001). There is little potential of negative effects of TDG supersaturation on
populations of adult Chinook salmon, although fish tissues are probably damaged by the dissolved gases (E. L.
Johnson etal.,2005,2007). Among the fish that pass through dam spillways, the survival of juvenile salmonids
is the highest (Beeman & Maule,2006; Muir etal.,2001).
Overall, most studies focus on the impact of dam construction on the spawning process and migration of salmo-
nid smolts. However, sufficient attention should also be devoted to investigating the direct impact of altered
flow regimes and water temperatures on the upstream migration of adult salmonids, which can provide specific
evidence to improve conservation measures for adult salmonids during upstream migration. Meanwhile, natural
factors such as climate and oceanic conditions also affect the physiological behavior of salmonids. Dam-induced
and natural factors could interact with each other, resulting in nonlinear interdependence (Goodwell &
Bassiouni,2022; Ye etal.,2015). This complex interaction highlights the necessity to better understand how
co-varying dam-induced and natural factors influence the physiological behavior of salmonids across their whole
life cycle. Engineering mitigation measures have been used to reduce the adverse influence of dams on salmo-
nids. However, it is controversial whether these mitigation efforts are effective if differential mortality rates of
salmonids occur for reasons unrelated to river damming (Rechisky etal.,2013; Welch etal.,2008). In the future,
long-term observations of upstream and downstream salmonid are needed to determine whether dams are the
major cause for the decline of salmonids in impounded rivers.
3.2. Impact on Chinese Carps
The four major Chinese carps (FMCCs) are grass carp (C. idellus), black carp (Mylopharyngodon piceus), silver
carp (H. molitrix) and bighead carp (Aristichthys nobilis). FMCCs are freshwater fish whose adults migrate
upstream to spawning sites during flood seasons and spawn semi-buoyant drifting eggs (Figure8). The fertilized
eggs drift a long distance for development, and then juvenile fish swim into riparian lakes, which serve as nursery
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
19 of 64
habitats (Q. Chen etal.,2021; W. Xu etal.,2020; B. Yi etal.,1988). The spawning activity of FMCCs requires a
minimum flow velocity to trigger the release of eggs (Q. Chen etal.,2021), and the drifting eggs require a suitable
flow velocity to keep them suspended (Garcia etal.,2013; M. Li etal.,2013; Q. H. Yang etal.,2014;P.Zhang
etal.,2021). High flow events are the most visible physical phenomena associated with the reproductive success
of FMCCs (Coulter etal.,2018; Embke etal.,2016). Under natural conditions, adult fish usually initiate spawn-
ing activity when the water level continues to rise for 0.5–2days, and reduce or even cease spawning when water
levels begin to recede (B. Yi etal.,1988). In China, the Yangtze River basin provides the principal habitats and
spawning sites for the FMCCs, although they are widely distributed across many parts of China. The FMCCs in
the Yangtze River are the dominant natural germplasm resource, which sustains the gene diversity and maintains
freshwater fish aquaculture (J. Wang, Li, Duan, Chen, etal.,2014). Prior to the construction of the Three Gorges
Dam, there were 30 spawning sites for FMCCs in the mainstream Yangtze River (C. Tang etal.,2022), with the
middle reach of Yangtze River being the main reproduction area. There are 12 spawning sites scattered along the
380km long reach from Yichang to Chenglingji, producing about 43% of the total eggs in the Yangtze River (Y.
Yi etal.,2010).
FMCCs play a vital role in providing high-quality animal protein, ensuring national food security, as well as
promoting rural economic development in China (Ban etal.,2019; D. Li, Prinyawiwatkul etal.,2021). However,
FMCCs are an aggressive invasive species in other countries (D. Li, Prinyawiwatkul, etal.,2021), as shown in
Figure8. The silver carp and bighead carp, jointly known as Asian carp, were first imported to North America
in the 1970s as aquaculture fishes. Due to their rapid growth rates and lack of natural predators, they have estab-
lished dense populations in the Mississippi, Ohio, Missouri and Illinois Rivers, and may pose a threat to the Great
Lakes (Heer etal., 2019; Wittmann etal., 2014; Zhu etal.,2018). In Europe, bighead carp are found to have
been widely recruited in northeastern Italy, demanding serious efforts to limit the spread and establishment of
reproducing populations (Milardi etal.,2017). In Australia, Asian carp impose significant impacts throughout the
Murray-Darling basin and other freshwater systems (Marshall etal.,2018).
River damming can block or delay reproductive migration, affect fish assemblages, impair spawning habitat
conditions, and thus reduce the recruitment of FMCCs. After the impoundment of the Three Gorges Reservoir
(TGR) in 2003, the larval abundance of FMCCs in the middle reach of the Yangtze River declined sharply to
less than 20% of the pre-dam abundance (P.Zhang etal.,2021), and the eggs and larvae in the lower reaches of
the Yangtze River declined to 0.34 billion, accounting for only 13% of the pre-dam amount (M. Li etal.,2016).
The Three Gorges Dam reduces river connectivity, preventing FMCCs migration (P.Zhang etal.,2021). Loss
of connectivity in dammed rivers also restricts the gene flow between fish populations located upstream and
downstream of a dam, affecting population genetic structure. Since the construction of the Gezhouba Dam and
Three Gorges Dam, significant genetic differentiations have appeared in the grass carp populations between the
upper and middle reaches of the Yangtze River, which might stimulate population divergence (Zhao etal.,2011).
Although spawning grounds for FMCCs downstream of the dam remain after impoundment, eight spawning
Figure 8. Native and invading distribution of Chinese carps around the world and their migration route for spawning.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
20 of 64
grounds in the reservoir have been lost due to establishment of the lentic ecosystem (Y. Yi etal.,2010). Reservoir
operations cause direct changes in the annual hydrograph and thus water depth and flow velocity, which affect
the spawning activity of FMCCs (Duan etal.,2009; W. Jiang etal.,2010; Z. Wang etal.,2013). Downstream of
the Three Gorges Dam, the suitable flows required for the reproduction of FMCCs are suggested to start from
12,500m
3/s on the day when a spawning event begins, and gradually increase to 18,600m
3/s at day 4 when
spawning actions reach their peak, and then quickly drop to support hatching and larvae survival (Q. Chen
etal.,2021). However, this flow process is mostly not met during the spawning period of FMCCs.
Water temperature has strong impacts on the reproduction of FMCCs. Laboratory studies and field measurements
have shown that the minimum water temperature threshold for gonad development of FMCCs is 18°C, and the
ideal water temperature range for spawning is between 21 and 24°C (Embke etal.,2016; M. Li etal.,2013; Y. Yi
etal.,2010). There is also a significant correlation between spawning time and the arrival date of the cumulative
temperature needed for gonad development. The initial date of FMCCs spawning in the Yangtze River has been
delayed from early May to middle June since operation of the TGR in 2003. The main reason for this delay is that
the water temperature downstream of the dam drops by 2–4°C from March to May, which is the critical period for
gonad development of FMCCs (J. Wang, Li, Duan, Luo, etal.,2014), and thus postpones the arrival date of their
gonad maturity. In addition, flood release could cause the TDG saturation to exceed 120% in the downstream
reach of the Three Gorges Dam during the spawning periods of FMCCs, resulting in the dispersing larvae being
highly vulnerable to the effect of TDG supersaturation. Dead larvae were found due to GBD during the early
operation stage of the Gezhouba Dam between 1981 and 1984 (L. Liu etal.,1986).
However, as invasive species in some regions such as the North America, a viable strategy to prevent the spread
of FMCCs and reduce their population is to manage the existing dams and barriers to limit their dispersion (Fritts
etal.,2021; Whitledge etal.,2019). In the upper Illinois River, reducing gate openings of the locks and dams
during the late spring and summer could provide opportunities to avoid the upstream migration of bighead carp,
and thus limit their recruitment into the upper section of the river (Lubejko etal.,2017). Rather than attempting
to directly block the migration route through dams, another option would be to encourage adult carp to spawn in
the reaches that have suitable spawning sites but lack adequate hydrological conditions to further support embryo
development. Such action could be implemented by adaptive management of hydraulic engineering structures in
rivers (Coulter etal.,2018; Cupp etal.,2021; Prada etal.,2020).
Previous studies attribute the primary cause of decreases in the population of FMCCs in the Yangtze River to the
construction of dams. However, it has also been argued that intensive fishing contributes greatly to their decline
(D. Chen etal.,2009). However, there are few published data on long-term trends in FMCCs catches, making
it difficult to assess whether dam construction or overfishing plays the major role in the degradation of natural
FMCCs resources. Meanwhile, a series of physical barriers, including dams, behavioral deterrents and electric
dispersal barriers, have been applied to block migratory pathways of FMCCs in the rivers in which they have
become invasive, although in practice the effectiveness of this approach is limited. There remains a knowledge
gap in understanding the associated hydrological mechanisms that affect the spawning behavior and early life
development of FMCCs. Studies have indicated that a warm winter-spring period advances the expansion of
Asian carp by altering thermal characteristics to be more favorable for their growth and lessening the time for
competitive interaction with invasive mussels (Alsip etal.,2020; M. Li etal.,2013). It has been observed that the
surface water temperature of the Great Lakes is warming faster than the global rate (Collingsworth etal.,2017;
O’Reilly etal.,2015), and there is thus a great value to investigate whether climate change will increase the risk
of invasion by Asian carp in this region.
3.3. Impact on Sturgeon
Sturgeons are one of the earliest extant vertebrates, and play an important role in the evolution of fishes and even
all vertebrates (Shen etal.,2020). There are 27 species of sturgeons that belong to the Acipenseridae and Poly-
odontidae families. The family Acipenseridae mainly includes Chinese sturgeon (A. sinensis), Russian sturgeon
(Acipenser gueldenstaedtii), Dabry's sturgeon (Acipenser dabryanus), Atlantic sturgeon (Acipenser oxyrinchus),
ship sturgeon (Acipenser nudiventris), sterlet (Acipenser ruthenus), and starry sturgeon (Acipenser stellatus).
The family Polyodontidae, which is commonly called paddlefish, includes the American paddlefish (Polyodon
spathula) and Chinese paddlefish (Psephurus gladius). The number of wild sturgeons in the world has decreased
significantly in recent years, due to the degradation of natural habitats, dam construction, and overfishing (Billard
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
21 of 64
& Lecointre,2000; T. Webb & Meyer,2019). Sturgeons are typically anadromous fish that spend most of their
lives in the ocean and return to freshwater to spawn (Figure9). They lay adhesive eggs in rivers with fast currents
and gravel beds. They are subcooled fishes, which prefer to live in water with relatively low temperature (Billard
& Lecointre,2000; McDowall,1997; Siddique etal.,2016), and are mainly found in Eurasia and North America
(Du etal., 2020). According to the International Union for the Conservation of Nature and Natural Resources
(IUCN) Red List of Threatened Species, Dabry's sturgeon has been considered extinct in the wild, Chinese
paddlefish has been considered completely extinct, and 85% of the extant 25 species, such as Chinese stur-
geon, Russian sturgeon, starry sturgeon and sterlet, are endangered (Brevé etal.,2022; IUCN,2022; Lenhardt
etal.,2006; H. Zhang, Jarić, etal.,2020).
Chinese sturgeon has lived in the Yangtze River for about 140 million years, and is often considered a living
fossil. It is a demersal fish species that usually breeds in river reaches with fast currents and gravel beds (Z.
Huang & Wang,2018). Every year, this migratory fish travels 2,000 miles from the East China Sea to its spawn-
ing grounds in the Yangtze River (D. Cheng etal.,2015). The spawning season of Chinese sturgeon is from
October to November, and spawning occurs at water temperatures between 15.3 and 20.0°C, with the optimal
water temperature for spawning being 18.0–20.0°C (Y. Wang etal.,2020). Atlantic sturgeon is one of the seven
sturgeon species found in North America, and extends from New Brunswick, Canada, to the eastern coast of
Florida, USA. Atlantic sturgeon spawns in the spring, and the water temperature requirement for spawning is
13.3–23.0°C (Hager etal.,2014). It spawns in running waters with rocks, pebbles and other hard objects on the
bed, or in pits and pools under waterfalls (Popov,2017). Russian sturgeon is widely distributed in the Caspian
Sea, the Sea of Azov and the Black Sea, as well as in the rivers flowing into these waters (H. Song etal.,2022).
The water temperature requirement for the spawning of Russian sturgeon is 12–14°C (Elhetawy etal.,2020).
Most of the spawning activity of Russian sturgeon takes place in the sloughs of main channels, with a few indi-
viduals spawning in the high tide zone. Sterlet is distributed in the Black Sea, Caspian Sea, Yenisei River and Ob
River in Russia. The optimal water temperature range for sterlet spawning is 13–16°C (Ponomareva etal.,2020),
and the typical spawning grounds of sterlet include riverbed and roaming beach formed by spring water (Lenhardt
etal.,2006). Juvenile sterlets are often found in groups in shallow water, while individual adults are scattered
in deeper water for feeding. White sturgeon is the largest of the eight sturgeon species found in North America.
White sturgeon inhabits the Pacific Ocean from northern Baja California of Mexico to the Aleutian Islands in
Alaska, and the large rivers flowing into the Pacific Ocean between Monterey, California and Alaska. White
sturgeon exhibits freshwater amphidromy, although evidence suggests that only small proportions of their popu-
lations live in marine environments. White sturgeon is an iteroparous broadcast spawner, spawning in the spring
and early summer when water temperatures are between 7 and 18°C (Counihan & Chapman,2018). Therefore,
water temperature and riverbed substrate are major factors that affect sturgeon life cycles, including spawning
migration and reproductive processes.
Figure 9. Distribution of sturgeon around the world and their migration route for spawning.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
22 of 64
Global dam construction has blocked the migration routes and damaged the spawning habitats of sturgeons,
leading to a significant reduction in their populations. The completion of Gezhouba Dam in 1981 and the Three
Gorges Dam in 2003 on the Yangtze River blocked the migration routes, cut off the spawning grounds and
affected the spawning time, spawning size as well as spawning frequency of the Chinese sturgeon, Dabry's stur-
geon and Chinese paddlefish, leading to significant reductions in their populations (Boscari et al., 2022; H.
Zhang, Jarić, etal.,2020). The Gezhouba Dam shortened the migration distance and reduced the capacity of
new spawning grounds of Chinese sturgeon, resulting in the number of hatched larvae being only 25% of that
before dam construction (Z. Huang,2019; J. Tao etal.,2009). The construction of the Volgograd Dam on the
Volga River, Russia, has destroyed the natural spawning area of the Russian sturgeon, starry sturgeon and Beluga
sturgeon, which caused mass mortality of these species. The Volgograd Dam has reduced the area of Russian stur-
geon spawning grounds by 80% (Secor etal.,2000), and the limited spawning area led to high spawning densities,
resulting in egg mortality rates of up to 60% (Popov,2017). In the Don River, Russia, dams cut off the main path
of Russian sturgeon migrating to their spawning grounds, resulting in a sharp reduction in their spawning areas
that are only located downstream of the dam (Boldyrev,2018). At present, Russian sturgeon with natural repro-
duction only exists in the undammed Ural River in Russia and Kazakhstan. In North America, the Eddyville Dam
on the Hudson River has blocked the migratory routes of shortnose sturgeon (Acipenser brevirostrum), leading
to a significant reduction in its population and hence its inclusion in the Endangered Species Act (T. Webb &
Meyer,2019). In the Great Lakes region, dam induced sedimentation in many tributaries has led to a reduction
in spawning grounds and juvenile habitats of lake sturgeon (Acipenser fulvescens), posing great challenges to
the restoration of their population (C. C. Wilson etal.,2022). In Europe, the construction of the Djerdap I and
Djerdap II dams blocked the migration of sturgeon in the Danube River, which seriously reduced populations of
the ship sturgeon and Atlantic sturgeon (Lenhardt etal.,2006). Therefore, river damming seriously compromises
the connectivity of sturgeon habitats worldwide, leading to significant declines in successful spawning and egg
hatching rates, and thus severely decreasing the population of sturgeons.
Water temperature is a primary factor affecting the migration of sturgeons, and determines their spawning time.
However, reservoir operations can alter water temperature regimes and thus affect the migration and spawning
of sturgeon (Y. Wang etal.,2020; H. Zhang etal.,2019). In the Yangtze River, the patterns of water tempera-
ture have been temporally and spatially altered by reservoir operation, which leads to the degradation of gonad
development and delays the spawning of the Chinese sturgeon (Z. Huang & Wang,2018; H. Zhang etal.,2019).
The TGR and Xiluodu Reservoir in the upper Yangtze River have reduced the effective breeding quantity down
to below 4.5% by elevating the water temperature that inhibits breeding activity during the spawning season. The
cumulative effect of the cascade dams, including Wudongde, Baihetan, Xiluodu, Xiangjiaba, Three Gorges and
Gezhouba, has led to an ongoing decline in the abundance of adult Chinese sturgeon in the Yangtze River and
the sea (Chang etal.,2017; Z. Huang & Wang,2018). In the Sacramento River, USA, the optimal water temper-
ature for Chinook salmon spawning is 12°C, while the optimal water temperature for green sturgeon (Acipenser
medirostris) growth is 19°C. At present, the Keswick Dam regulates the water temperature of the river in winter
to match the suitable spawning temperature for the endangered Chinook salmon, causing the cold water to extend
to the habitat of the green sturgeon, whose growth is thereby greatly affected (Zarri etal.,2019). In the Colum-
bia River, cascade dams have fragmented the spawning habitat of white sturgeon in the mainstream into short
sections connected by long-distance impoundments. Due to differences in dam operation and geographically
local environments, the water temperature in different river reaches varies greatly, which in turn leads to differ-
ences in the spawning temperature and time of white sturgeon in different river sections. In the lower Columbia
River, the spawning of white sturgeon downstream of the Bonneville Dam begins at a water temperature of 8°C,
but spawning in the three furthest downstream dam tailraces begins when water temperature reaches at least
10°C. In addition, spawning occurs earlier downstream than upstream in the Columbia River during the spring
season (Counihan & Chapman,2018; Péril,2004).
Dam construction alters the natural hydrological regimes of rivers, and thus affects the growth and reproduc-
tion of sturgeons. The reduced downstream flow results in limited lateral connectivity to the floodplain, which
adversely affects habitat availability and reproduction of sturgeons (F. He etal.,2021). In the Kootenai River, the
demand discharge during the spawning period of the white sturgeon is 1,416–2,832m
3/s; however, the peak flow
is reduced usually to 250–450m
3/s due to operation of the Libby Dam, which has seriously affected the reproduc-
tion of white sturgeon (Paragamian etal.,2001). River impoundment reduces flow velocity and possibly causes
hypoxia in the transition zone of the reservoir upstream, which exposes the pallid sturgeon (Scaphirhynchus
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
23 of 64
albus) to low oxygen conditions and thus results in their high mortality (Guy etal.,2015). However, dedi cated
flow regulations can in some cases provide suitable condition for migration and spawning of sturgeon. For exam-
ple, intermittent flood pulsing through dam operation to achieve key water temperature thresholds may have
better facilitated the upstream migration and spawning of shortnose sturgeon adults (Vine etal.,2019).
Previous studies claim that the barrier effect of dams is the primary cause for the decrease in Chinese stur-
geon populations. However, new spawning grounds for Chinese sturgeon may have formed downstream of the
Gezhouba Dam (P.Zhuang etal.,2016), indicating that blockage by the dam has not completely hindered the
natural reproduction of this species. Therefore, it is essential to investigate the impacts of other factors, such as
altered water temperature regimes, in addition to the blockage of migration routes, on the spawning of Chinese
sturgeon. To date, most studies focus on the impacts of river damming on the migration and spawning of sturgeon,
but it is also important to understand the impact on other life stages, such as foraging and overwintering.
3.4. Impact on Eel and Lamprey
Eels (genus Anguilla) and lamprey (genus Lampetra) are among the most valuable fishery species around the
world (P.R. Almeida etal.,2021; FAO,2015). There are 19 species or subspecies of eels being identified glob-
ally (Righton etal.,2021), of which 17 are scattered throughout the Indo-Pacific and the other two species are
distributed in the Atlantic Ocean (Figure10). At present, several species of eel have been listed as “endangered”
or “critically endangered” in the wild (Jacoby etal.,2015; Vié etal.,2009), including the European eel (Anguilla
anguilla L.), American eel (Anguilla rostrata) and Japanese eel (Anguilla japonica). Lampreys are often misi-
dentified as eels because of their similar appearance; however, lampreys belong to the Order Petromyzontiformes
whilst eels belong to the Order Anguillifomes (Renaud,2011). Most lampreys, such as the sea lamprey (Pet ro-
myzon marinus), European river lamprey (Lampetra fluviatilis), Pacific lamprey (Entosphenus tridentatus), and
Caspian lamprey (Caspiomyzon wagneri), are anadromous parasitic fish. The sea lamprey is a notorious invasive
species in the Laurentian Great Lakes (Figure10), and has devastated the fisheries of whitefish and lake trout
(McDonald & Kolar,2007; Zielinski etal.,2019).
The life stages of eels comprise eggs, leptocephalus (larva), glass eel (post-larva), elver (juvenile), yellow
eel (non-mature adult), and silver eel (migratory adult). At each life stage, eels exhibit distinct morphologies
(Tsukamoto etal.,2011). The larva of eels is notably larger than that of almost all the other fish species, and
their morphology is well equipped for both passive and active swimming in their oceanic migration (Righton
etal.,2021). Eels usually form their sex at the silver stage (Miller & Tsukamoto,2016). Unlike salmonid and
sturgeon, eels have reverse migration and spawning behaviors, as they grow up in freshwater and return to the
ocean for reproduction (Figure10). They typically dwell in seawater during the life stages of eggs, leptocephalus
and glass eel, and in brackish and freshwater during the life stages of elver, yellow eel, and silver eel (Haro,2014).
The spawning of temperate eels, such as European eel and American eel, occurs in the Sargasso Sea (Béguer-Pon
Figure 10. Native and invading distribution of eel and lamprey around the world and their migration route for spawning.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
24 of 64
etal.,2015; Schmidt,1923; Tucker,1959), which has been recently reconfirmed by genetic analyses (Barth
etal.,2020). The eggs and spawning adults of Japanese eel have only been discovered within a restricted area
along the seamount chain of the Pacific West Mariana Ridge (Tsukamoto etal.,2011). In contrast to the distribu-
tion of spawning zones, the habitats of eels during the entire growth stage are flexible, encompassing all saltwater
and freshwater areas, because eels can adapt to diverse environmental conditions and different dietary niches
(Tesch,2003). Lampreys have an anadromous migration habit and a wide habitat range (Moser etal., 2021).
Some lampreys are life-long freshwater dwellers, whilst others are seawater dwellers that migrate to freshwater
to spawn (Renaud,2011). The sea lamprey breeds anadromously in basins of Western Europe and eastern North
America, and swims into the North Atlantic and Western Mediterranean. The Korean lamprey (Eudontomyzon
morii) is a small freshwater species that is mostly distributed in the Yalu River as well as some mountain rivers
in northeast China, North Korea and the Russian Far East (Renaud,2011).
Although eels and lampreys have strong acceleration and high mobility (Tytell & Lauder,2004), they cannot
jump over obstacles and their burst swimming speed is relatively low (Kemp etal., 2011). The barrier effect
caused by dams is thus the most immediate impact on eels and lampreys, impeding their migration between
spawning and rearing grounds. The downstream passage through turbines and dams leads to the high disorienta-
tion and mortality of breeding silver eels, which results in significant declines in spawning rates and populations.
It has been reported that 75% of eels are delayed in their downstream migration, and up to 65% are definitively
halted by dams (Besson etal.,2016). Eels are especially vulnerable to screens and turbines due to their anguill-
iform morphotype (Kerr etal.,2015), and thus hydroelectric facilities in dams can cause sublethal injury and
direct mortality to migrating adult eels (Bruijs & Durif,2009). In rivers with cascade dams, the cumulative effect
may result in the level of overall escape of eels not reaching the required conservation criteria, which is 40% for
silver eel in European countries (Pedersen etal.,2012). Habitat fragmentation caused by dams also impedes the
upstream migration of eels (Jellyman,2022). More than 15,000dams have been constructed in the coastal drain-
ages of the North Atlantic, obstructing direct access to 87% of rivers and streams flowing into the Atlantic, and
hence drastically reducing the inland extension of American eels (Jellyman,2022; Miller & Casselman,2014).
The decline of the European eel (Bevacqua etal.,2015), American eel (Kwak etal.,2019) and Japanese eel (J.
Z. Chen etal.,2014) in turn affects the ontogenetic stage and physiological traits of eels (Righton etal.,2021).
The barrier effect of dams is also the main reason for the decline in the global population of most lampreys, such
as about 80% loss of sea lamprey in the Iberian Peninsula and Caspian lamprey in the Volga River, leading to the
collapse of associated commercial fisheries (Atkinson etal.,2020; Jellyman,2022), and strong decline of Pacific
lamprey in the Columbia and Snake Rivers (Moser & Close,2003).
Eels have specific habitat preferences and requirements, with their habitat selection involving water tempera-
ture, water depth, substrate, salinity, flow velocity, oxygen concentration, vegetation cover, prey availability,
and predation threat (Jellyman,2022; Righton etal.,2021). Water temperature appears to be an important factor
throughout the whole lifespan of eels, as they grow faster and mature earlier in warm waters (Tesch,2003). Water
depth can affect the habitat quality of eels, since small eels tend to favor shallow waters and large eels prefer
deep waters (Jellyman,2022; Jellyman & Arai,2016). In particular, large eels spend the daytime in deep waters
and night time in shallower waters (Righton etal.,2021). Eels choose different types of substrates in different
seasons, preferring soft substrates with abundant organics or silts in the spring, muddy substrates in the summer,
and rubble substrates in the autumn (Tomie etal.,2016). Their substrate preference varies with body size, with
smaller eels preferring coarse substrates (gravels, cobbles, and boulders) and larger eels preferring fine substrates.
These habitat variations are likely due to the combined effects of changes in physical space requirements and prey
preference with the increasing body size of eels (Lloyst etal.,2015).
Alterations of water temperature regime, flow velocity and riverbed substrate due to river damming can have
dramatic effects on the living conditions of eels (Righton etal.,2021). The release of hypolimnetic water from
reservoirs has a cooling effect on water temperature in summer and thus degrades the growth of eels downstream
of dams (Maheu etal.,2016a). The migration speed and route selection of eels are affected by flow veloc-
ity, and thus decreases in flow velocity due to dams may cause failure of their migration (Jansen etal., 2007).
During downstream migration, eels usually swim actively with a speed of 0.3–1.2m/s (Behrmann-Godel &
Eckmann,2003), and those that manage to migrate downstream pass over the dam crest only when the flow
velocity is high (Besson etal., 2016). Artificial fluctuations in water depth due to dam operations inevitably
affect the living environment of eels (Righton etal.,2021). Dams are known to cause substrate-sorting effects
and thus result in local habitat homogeneity, which limits the diverse substrate requirements of eels (Naganna
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
25 of 64
& Deka,2018). Dams intercept sediment and reduce downstream substrate availability for the critical life stages
such as nesting and refuge of eels (Černý etal.,2003). The decreased sediment flux can also cause delta recession
(Baisre & Arboleya,2006), which increases the susceptibility of eels to natural disturbances (Day etal.,2007;
O'Connor etal., 2015). The introduction of non-native species due to river damming squeezes the habitat and
even leads to extinction of native eels and lampreys (Marohn etal.,2014). Nonetheless, there are also some
marginal benefits of river damming for eels. Reservoir eutrophication has been reported to increase chirono-
mid populations, which comprises the major food of eels (Aprahamian etal.,2021). The increase in plant and
zooplankton production caused by reservoir eutrophication is considered to provide an additional benefit to the
populations of eels (Bunting etal.,2007). In some cases, the isolation created by dams can be beneficial to eels
by preventing the introduction of contaminants, parasites and diseases into their habitats (Liermann etal.,2012;
Righton etal.,2021).
Limited by monitoring technology, comprehensive and reliable data on the migration routes and spawning
grounds of eels are sparse, which hinders accurate assessments of the impact of river damming on eels. To date,
relevant studies outside of Europe, the Americas and Japan are scarce, and the lack of information across Africa
is a severe barrier to the appropriate assessment of the conservation status of eels on a global scale (Righton
etal.,2021). In addition, when compared to intensive studies concerning the impacts of river damming on eels,
similar research on lampreys needs to be enhanced. Although an individual small dam has much less impact on
eels and lampreys than a large dam, the cumulative impact of many small dams requires adequate attention, given
their widespread and intensive distribution (Lehner etal.,2011). Climate change could alter water temperature
regime and thereby induce impacts on the habitat of eels and lampreys, and is regarded as one of the least under-
stood risks to the species worldwide (Jacoby etal.,2015).
3.5. Impact on Other Fish Species
River damming also impacts on other fish species, which have less socioeconomic and cultural value, such as
giant catfish, tilapia and sardines (Figure11).
Giant catfishes are widely known for their giant bodies and mysterious trails in deep water areas, and they
can grow to more than two m in length (Boulêtreau & Santoul,2016). Giant catfish species mainly include
the Mekong catfish (Pangasianodon gigas, Pangasius krempfi, Pangasius sanitwongsei, and Pangasius mekon-
gensis), Amazonian catfish (Brachyplatystoma rousseauxii and Brachyplatystoma filamentous), and European
catfish (Silurus glanis). Megafaunal species have disproportionate per capita effects on community structure
and ecological processes, and any shift in their abundance is likely to affect food webs and ecosystem functions
(Malhi etal.,2016). Giant catfishes generally spawn in freshwater rivers and live in estuaries with long-distance
migration (Figure11). The Mekong catfish migrates long distances to spawn, spending much of their lives in
the brackish waters of the Mekong Delta and in the South China Sea near Vietnam before returning to spawn in
Figure 11. Distribution of giant catfish, tilapia, and sardines around the world, as well as migration route for spawning of giant catfish.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
26 of 64
the Mekong River in Laos and Thailand (Hogan etal.,2007). Amazonian catfish adopt a basin-wide migratory
life cycle between the Andean piedmont and the Amazon estuary, which makes them possess the longest river
migration that may total up to 12,000km in the case of Brachyplatystoma rousseauxii (Duponchelle etal.,2021).
Amazonian catfish are of high commercial importance to fisheries of the Amazon River basin, forming a major
ecosystem service provided by this river system (Fraser,2018). Due to the impact of dam construction, giant
catfish, and especially long-distance migratory species such as the Mekong and Amazonian catfish, have become
endangered in the wild, threatening the stability of ecosystem structure due to their top-tier status in the food web
(Duponchelle etal.,2021; Fraser,2018; Hermann etal.,2016; Hogan,2011). Dams on the Mekong River have
disrupted the migration and spawning of giant catfish, and the situation may become worse as Laos plans to build
more dams on the mainstream of the Mekong River (Soukhaphon etal.,2021). The migration route and access
to a substantial portion of the spawning grounds of the Amazonian catfish in the Madeira River are blocked by
two dams built a decade ago, which profoundly affects the populations and fisheries of the Amazonian catfish
and further alters the food web of the river ecosystem controlled by this apex predator (Duponchelle etal.,2016;
Fearnside,2013; Fraser,2018). It has also been reported that the body size and trophic level of Amazonian catfish
have declined in the reservoir and downstream reach of the Belo Monte Dam on the Xingu River after large
hydropower development (Keppeler etal.,2022). Dams intercept sediment and nutrients that nourish the Amazon
River and its floodplain, and damage fish physical habitats with a cascade of ecological effects (Fraser,2018).
The disruption of nutrient connectivity is more pronounced by dams in the Andean-Amazonian transition, whose
rivers provide most of the sediment and nutrients to the Amazon (Lees etal.,2016).
Tilapia is the third most-cultured finfish in inland aquaculture, with its production reaching 4.41 million tons in
2020, and accounting for 9% of world inland aquaculture (FAO,2022). Tilapia mainly includes the Nile tilapia
(Oreochromis niloticus) and Mozambique tilapia (Oreochromis mossambicus). The Nile tilapia is indigenous in
the Nile River; however, it is highly invasive elsewhere and has invaded many ecosystems worldwide (Faunce &
Paperno,1999). The Mozambique tilapia is indigenous in South Africa, Zimbabwe, Mozambique, and Eswatini
(Moyo & Rapatsa,2021). Tilapia possess a strong survivability and are physiologically tolerant to a wide range of
salinities, dissolved oxygen content and water temperature, and are characterized by multiple spawning, parental
care, and extreme feeding plasticity (Avella etal.,1993; G. Farmer & Beamish,1969; Vitule etal.,2009). Tilapia
inhabits the middle and bottom layers of water bodies, and can survive in waters of 16–40°C, with the optimal
temperature being 28°C. Tilapia is a warm-water fish, which will hibernate or even die when water temperature
is lower than 15°C. Tilapia feeds on the base of the food web at the bottom layer of a water body, most often on
sediment resources such as nematodes, rotifers, bryozoans and hydrozoans, and is well adapted to surviving and
growing in non-native environments (Peterson etal.,2006). Tilapia is widely cultured in reservoirs using cages in
tropical and subtropical regions (Hishamunda,2007), which can alleviate pressures from water scarcity and also
help tilapia to overwinter safely (Moyo & Rapatsa,2021). However, due to the strong environmental adaptation
ability of tilapia, non-native tilapia that escapes accidently from aquaculture in reservoirs could rapidly become
the dominant population. This increases the risk of invasive species to the original ecosystem of reservoirs, result-
ing in the loss of genetic integrity of native species, damage to the native biodiversity, and further affecting the
stability of the native river ecosystem (Bernery etal.,2022; Canonico etal.,2005; Cucherousset & Olden,2011).
Therefore, it is critical to quantify the impact of cultivated tilapia on the native fish species in reservoirs, and
conduct actions to control their habitat range and prevent their escape into the wild, in order to protect the ecosys-
tem stability of dammed rivers.
River damming also affects non-migratory offshore marine fishes such as sardines (Sardina pilchardus), which
are mainly distributed in the Atlantic and Mediterranean (Figure11). Sardines are a small pelagic fish that typi-
cally feeds on plankton, and plays an important role in global fisheries (FAO,2019). The availability of food such
as phytoplankton and zooplankton depend on the nutrient inflows from rivers, and thus river damming indirectly
affects sardine populations (Biswas & Tortajada, 2012). For example, the autumn flooding of the Nile River
irrigates and fertilizes the floodplain annually, and supplies sufficient nutrients to the Mediterranean Sea. Before
1965, the flood of Nile River delivered about 7×10
3 tons year
−1 of nitrogen and 7–11×10
3 tons year
−1 of phos-
phorous to the Mediterranean coast (Nixon,2003). The nutrients in the Nile floodwater support a massive diatom
bloom and a productive fishery, particularly for sardines (Halim,1960; Halim etal.,1964). However, completion
of the Aswan High Dam in 1965 decreased the fall flood by about 90% (Dorozynski,1975) and reduced the fertil-
ity of the southeastern Mediterranean waters, leading to a sharp decline in the marine fisheries (Milliman,1997;
Mohamed,2019; Rzóska,1976). It has been reported that sardine catches along the Egyptian coast declined
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
27 of 64
from 18,000 tons in 1962 to a mere 460 tons in 1968 (Abul-Atta,1978). Since the late 1980s, the total fish
catch, including sardines, of the Nile and its estuaries have recovered back to the level before construction of the
Aswan High Dam (General Authority for Fish Resources Development,2012). The cause for this recovery may
be attributed to the increasing discharge of sewage from urban expansion stimulated by damming in the Nile
(Nixon,2003; A. J. Oczkowski etal.,2009; Richards,1982). However, the long-term fertilizing effects of sewage
require further investigation, as the persistent discharge of highly poor-quality water could damage the coastal
ecosystem and lead to an unsustainable increase, or even decline, of the fishery (A. Oczkowski & Nixon,2008).
In many cases, the time of dam construction coincides with adjustment of fishery policies that aim at increasing
fish catch and consequently results in overfishing. The contribution of overfishing and river damming to the
decline of fish catches demands quantification, in order to investigate and reveal the extent of fishery decline
attributed to dam construction (Biswas & Tortajada,2012). Meanwhile, due to rapid urbanization and indus-
trialization, nutrient supply to downstream river reaches and estuaries has largely increased (A. J. Oczkowski
etal.,2009), leading to rebounds of estuarine fisheries after river damming. It is essential to understand whether
this effect is sustainable or brings new challenges. Most current studies have focused on endangered or major
commercial fish species in order to prevent extinction or sustain fishery economies. However, fish diversity plays
an irreplaceable role in the aquatic food web and its integrity, as well as the stability of river ecosystems. There-
fore, we highlight the necessity for future investigations on the impacts of river damming on more fish species
and fish communities.
4. Fish Conservation Measures in Dammed Rivers
To offset or mitigate the impacts of river damming on fish, a variety of conservation measures, including fish
passage facilities, artificial breeding and release, ecological reservoir operation, and habitat compensation in trib-
utaries, have been proposed and implemented. Each measure has its own advantages and limitations, application
conditions, efficiency, and cost-effectiveness, and thus should be selected according to the specific situations of
a given dammed river. Of course, policies concerning fishery conservation play an essential role, but are outside
the scope of the present review. These policies include, but are not limited to, setting up nature reserves and
germplasm resources reserves, implementing policies and laws for no-catch measures, and navigation restrictions
during fish-sensitive seasons such as the spawning season (S. Huang & He,2019; Maxwell etal.,2020; H. Zhang,
Kang, etal.,2020).
4.1. River Connectivity Restoration and Fish Passage Facilities
Restoring longitudinal and lateral connectivity, such as river-floodplain reconnection and fish passage facilities,
could be the most direct approach to rehabilitate physical habitat and migration routes of fish in dammed rivers.
In recent decades, there have been increasing efforts to restore fish floodplain habitats in riverine rehabilitation
practices, as floodplain channels that are lost or disconnected from the main river have demonstrated visible
impacts on river ecosystems. Establishing the connectivity of floodplain channels to the main channel and restor-
ing lateral connectivity of river can provide essential nursery areas for fish and mitigate the loss of fish diversity
(Stoffers etal.,2022). The restored floodplain channels in the Rhine River, Netherlands, have served as suitable
nursery areas for rheophilic fish species (Stoffers etal.,2021). The restoration of hydrologic linkages between
the main channel and floodplain in the Kissimmee River, USA, has demonstrated positive effects on food web
structure and ecosystem functioning (Jordan & Arrington,2014). In the upper Danube River, a secondary flood-
plain channel has been artificially created following a nature-based construction scheme, which has provided
additional habitats and restored migration routes, thereby making an important contribution to restoring the
population of endangered fishes (Pander etal.,2015). The restoration on a tributary of the upper South Esk River,
Scotland, has reconstructed a natural meandering channel guided by historical maps, which reinstates floodplain
connectivity and habitat for Atlantic salmon and trout (Addy etal.,2016).
Fish passage is potentially an effective engineering approach to reconnect fragmented ecological corridor due
to river damming and restore river longitudinal connectivity. It is the earliest, and also the most widely used,
measure to conserve migratory fish in dammed rivers (Schilt,2007). Fish passage mainly includes fishway, fish
lift, fish collection, and transportation, turbine passage, juvenile bypass, and other engineering transport meas-
ures (Figure12). The earliest fishway can be dated to the mid-18th century in Europe (Clay,1995). In the early
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
28 of 64
20th century, field and laboratory experiments on different fishway designs were performed. Denil(1909, 1938)
created the unique Denil fishway to reduce the flow velocity inside the path (Figures12a and 13a). In 1946,
a vertical slot fishway (Figures12b and13b) was constructed on both sides of Hells Gate in the Fraser River,
Canada, to allow salmonids to successfully cross the channel barrier caused by a landslide (Jackson,1950). Monk
etal.(1989) proposed a fishway configuration of pool and weir, in which nearly 100% of shad and almost all
Figure 12. Different types of fish passage facilities. (a) Denil fishway. (b) Vertical-slot fishway. (c) Nature-based bypass system. (d) Fish lift. (e) Fish collection and
transport system. (f) Fish-friendly turbine. Panel (a)–(c) are redrawn from Thorncraft & Harris(2000).
Figure 13. Different types of fishways. (a) Denil fishway for Arctic grayling in the Big Hole River watershed in Montana,
Canada (source: Montana State University, photo by Matt Blank, 2015; https://www.montana.edu/ecohydraulics/research/).
(b) Vertical-slot fishway in the Mosel River in Koblenz, Germany (source: The Federal Waterways Engineering and Research
Institute (BAW); https://www.baw.de/en/die_baw/wasserbau/umwelt/umwelt.html). (c) Pool-weir fishway for migrating pink
and coho salmon to spawning grounds in Anderson Creek, Canada (source: McElhanney Company; https://www.mcelhanney.
com/project/anderson-creek-fishway/). (d) Nature-based bypass fishway beside the Se San River of Cambodia (source: the
Xinhua News Agency; https://new.qq.com/rain/a/20191026A099LC00).
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
29 of 64
other species could successfully pass through (Figure13c). These early efforts were mostly oriented at salmon
species, with only a small number of studies aiming at shad. Recent legislation for endangered species in the
United States, Canada, and Europe has re-emphasized the importance of fishways for migratory species other
than salmonids and shad; meanwhile, some successful initiatives have been taken for migratory species in other
parts of the world (Katopodis & Williams,2012).
With the rapid growth of large dams, the applicability of fishway has been facing great challenges due to high
cost, low efficiency and engineering complexity, which promotes the exploration of other types of fish passage.
Fish lift has a design principle similar to an elevator, which can actively move and release fish from the down-
stream reach of the dam to the upstream reservoir (Figure12d). These lifts induce fishes into a hopper that is
raised from the bottom to the upstream side of the dam (Santos etal.,2021). Barry and Kynard(1986) found
that the tailrace lift was more efficient than the early fish lift for American shad. It has been reported that the
combination of an Archimedes screw and fish lift can significantly improve the efficiency of fish lift (McNabb
etal.,2003; Zielinski etal.,2022). In some cases of high dams, where the construction of fishway and fish lift
is unfeasible, a fish collection and transportation system could be an appropriate alternative (Figure12e). Fish
collection and transportation is a special form of fish passage facility, which is mainly adopted in high dams or
in projects where fish need to climb several steps continuously. It attracts fishes into cabins or other boxes, and
then transports the fish over the dam by ships or vehicles (Figure14). In 1981, massive collection and transpor-
tation was implemented as an operational program by the United States Army Corps of Engineers (USACE) to
reduce losses of juvenile salmonids during their seaward migration. Fish collection and transportation systems
can improve the survival rate of fish passing through dams and has proved to be more effective than other fish
passage facilities in some contexts (D. L. Ward etal.,1997).
Traditional turbines are extremely harmful for fish to pass through, thus fish-friendly turbines such as Archi-
medes screw turbines and blunt blade turbines have been gradually used (Figure 12f). Fish-friendly turbines
have slowed rotational speeds and large openings, which can allow safe passage of small objects (Bracken &
Lucas,2013; YoosefDoost & Lubitz,2020). Bypass systems (Figures12c and13d), such as the curved-bar rack
bypass and the horizontal rod rack bypass, can guide downstream-moving fish toward a reasonably safe corridor
around water intakes, and thus effectively reduce the mortality of fish passing through dams (Beck etal.,2020;
Meister etal.,2022). In the eight dams of the lower Snake River and lower Columbia River, USA, most of the fish
entering the powerhouse are diverted to a juvenile bypass system, providing safe and efficient passage for juvenile
salmon to migrate downstream (Faulkner etal.,2019). River-like side channels are constructed in some dammed
rivers, which serve as a bypass for fish migration and even as a supplementary habitat for fish reproduction (L.
Zhang etal.,2023). In the Don River, Russia, a natural-type channel has been constructed to allow sturgeon to
bypass the Konstantinovskiy Dam. Stellate sturgeon (Acipenser stellatus) eggs are found in the bypass channel,
indicating that the bypass channel has been used as a spawning habitat (Pavlov & Skorobogatov,2014).
Overall, fish passage facilities can assist fish to pass through the barrier of dams and mitigate the impact of habitat
fragmentation. The efficiency of fish passages differs significantly between different types of fish passages. On
average, the efficiency of downward passage is slightly higher than that of upward passage; pool-weir, vertical
slot and naturalized fish passages have higher efficiency than Denil fishways and fish lifts (Noonan etal.,2012).
Figure 14. Fish collection and transportation system in Fengman hydropower station in Jilin, China (source: the Xinhua
News Agency; https://www.gov.cn/xinwen/2021-08/31/content_5634534.htm#10). (a) Fish collection from the box of fish lift.
(b) Transportation of collected fish to upstream.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
30 of 64
Fishways are used mostly in low dams, which are usually built in low gradient rivers, to improve river longitudinal
connectivity. The Denil and nature-based fishways are the most efficient fish passages (Baumgartner etal.,2018;
Bunt etal.,2001). These fish passages can remain effective even when significant fluctuations in upstream or
downstream water level occur (Quaranta etal.,2019). The Vianney-Legendre vertical slot fishway in the Riche-
lieu River, Canada, has shown to be successful for passing a variety of fish species, including lake sturgeon
(Marriner etal.,2016). The Denil fishways of the 63 diversion dams in the Big Hole River basin in southwestern
Montana provide opportunities for Arctic grayling (Thymallus arcticus) and other fish for year-round access to
critical habitats (Triano etal.,2022). The nature-based fishway installed at a low dam in Indian Creek, Canada,
has proven effective for multiple fish species to traverse the dam (Steffensen etal.,2013). However, fishways
may have a poor effectiveness for some fish. The effectiveness of a fish weir at Foster Dam in the Santiam River,
USA, is low to moderate for the downward passage of juvenile Chinook salmon, while it is consistently high for
steelhead trout, indicating that fish weirs may not be a suitable solution for all species (J. S. Hughes etal.,2021).
Overall, the passage efficiency of fishway is dependent on the slope, baffle and other characteristics affecting
the flow field inside the pools, the water head drop between the pools, as well as the turbulence levels and flow
velocities (Quaranta etal.,2019).
Fish lifts, fish collection and transportation systems are regarded as the most cost-effective fish passage facil-
ities for high dams. The fish lift of large hydropower dams on the Lima River, Portugal, effectively prevented
fragmentation of potamodromous populations between different reaches (Mameri etal.,2019). Fish collection
and transportation systems have the advantages of high flexibility, no interference to the structural layout of the
dam, suitability to large variations in reservoir water level, and small occupation of space for fish to cross the
dam. In the 1940s, a temporary collection and transportation system successfully transferred thousands of adult
salmonids at the Rock Island Dam on the lower Columbia River. Fish collection and transportation systems have
also been used to transport shad successfully at the Mactaquac Dam on the St. John River, Canada, and Essex
Dam on the Merrimack River, USA (Clay,1995). D. L. Ward etal.(1997) reviewed studies conducted by the
US National Marine Fisheries Service from 1968 to 1989 concerning the efficiency of using trucks and barges
to transport migrating juvenile Chinook salmon from the Snake River around dams to reservoirs in the lower
Snake and Columbia rivers, and suggested that the use of barges to transport juvenile Chinook salmon could
improve their survival rate. However, only 47% of Atlantic salmon succeeded in passing through the fish lift of
the Golfech-Malause hydroelectric complex on the Garonne River, France. In the Gezhouba dam, it has proven
successful to lure the bottom fishes by releasing jet flow during the collecting procedure (Y. Liang etal.,2014).
The main difficulty specific to fish lifts involves fish trapping, as the V-shaped entrance of fish lift may inhibit
salmonids entering the holding pool and cannot guarantee that the entered fishes will not return back to the river
(Croze etal.,2008).
The passage efficiency of different fish species under different hydraulic conditions in passage facilities varies
greatly from case to case (Bunt etal., 2016; Nieminen etal., 2016; Williams & Katopodis,2016). Salmonids
and clupeids have been found to efficiently pass through the vertical slot, pool-weir fishway, and Denil fishway,
with an efficiency of 63%, 45%, and 51%, respectively (Castro etal., 2016; Mallen-Cooper & Stuart,2007;
Noonan etal.,2012). Brown bullhead (Ameiurus nebulosus) and striped bass (Morone saxatilis), which have a
smaller body size than adult carp, prefer nature-based fishways, and have a passage efficiency of up to 70% (Bunt
etal.,2012). Fish lifts are the most effective up-migration measure for lamprey and brown trout, but the difficulty
is to capture small-sized individuals (Castro etal.,2016; Pompeu & Martinez,2007). Tummers etal.(2016) high-
lighted that the physical characteristics of baffles and high turbulence may inhibit lamprey ascending the passage,
and Moser etal.(2019) proposed a novel modification of fishway entrance for Pacific lamprey. The passage
efficiency of fishways is also related to the behavior of fishes (Shahabi etal.,2021), as their swimming direction
in fishways is dependent on their experience with the flow field (Goodwin etal.,2014). Inadequate attractiveness
for fish is recognized to be a major factor limiting the efficiency of fish passages (David etal.,2022). According
to Laine(1995), fish often need to become acquainted with the passage entrances before they start to climb the
passage facilities. Mensinger etal.(2021) suggested that fish may segregate at barriers based on their personality
and sizes, and this could be alleviated by increasing fishway attraction and maximizing passage opportunity,
leading to more exploratory eels passing through successfully. Generally, the functionality of fish passages in
alleviating dam barrier effects on fish is limited, and requires both good design of the facilities and good swim-
ming ability of fishes (Noonan etal.,2012).
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
31 of 64
Fish passage has demonstrated to be an effective way in many cases to reconnect fragmented fish habitats in
dammed rivers. However, it is argued that moving fish preferring a lotic environment from downstream reaches
to reservoirs could cause further damage to these fish species, as the lentic environment of reservoirs may form
an ecological trap that makes the transferred fish unable to find their migration path or suitable habitats. Fish-
ways are mostly favorable to the fish species with strong swimming ability, which could potentially change the
structure of fish communities both upstream and downstream of a dam, and hence further impair fish biodiversity
in the dammed river. There are also plenty of fish passages that have not achieved their expected functionality.
Their low effectiveness could be attributed to insufficient considerations of fish swimming ability and hydraulic
characteristics in the design of the facilities. This demands renewed efforts to develop innovative solutions, at
the core of which is the need for engineers and biologists to work together and design passages based on the
preferred hydraulic conditions of multiple fish species. In addition, nature-based solutions and application of
natural materials, instead of concrete and metal, in fish passage constructions are an important aspect. The lack
of long-term monitoring and prompt assessment of the effectiveness of the operational fish passages also restricts
the modification and improvement of such facilities.
4.2. Artificial Breeding and Release
Artificial breeding and release of target fish species in dammed rivers is a viable conservation measure to replen-
ish the populations of endangered fish species in the wild and restore fishery resources (Molony etal.,2003;
Naish etal.,2007; J. Yang etal.,2013), although the evidence base for effectiveness is incomplete (Rytwinski
etal.,2021). Artificial breeding and release can be categorized into two main types: “ecological restoration”
and “resource restoration.” “Ecological restoration” aims to conserve the endangered native fish populations
and prevent extinction by releasing hatchery-reared fish into the original habitats of dammed rivers. “Resource
restoration” aims to recover fishery resources and improve the economic fishery in dammed rivers by artificial
breeding and release (L. Wang,2016).
This measure has been adopted as an essential conservation strategy for more than 20 vulnerable fish species in
China (J. Yang etal.,2013). For instance, Chinese sturgeon are preferentially distributed in the lower reaches of
the Yangtze River; however, their spawning migration route is blocked and spawning grounds are damaged by
hydropower dams on the upper Yangtze River, which reduces the length of natural spawning ground from 600 to
7km and causes their gonads to degenerate (L. Wang & Huang,2020; Xie,2003; Y. Zheng etal.,2022). Since
1984, artificial breeding and release of larvae and juveniles (Figure15) into the natural environment has become
an important approach to conservation efforts of the Chinese sturgeon (Chang etal.,2021; Gao etal.,2009; Qin
etal.,2020; Stone,2008; Wei etal.,2004). From 1983 to 1998, approximately 6 million Chinese sturgeon fry
and juveniles were released into the Yangtze River (H. Wang etal.,2019). About 500 to 1,500 adult sturgeons
were released each year into the spawning ground of the Yichang section of the Yangtze River from 1997 to 2003
(J. Li etal.,2021;P.Zhuang etal.,1997). In addition, a breakthrough in artificial breeding technology for the
two sturgeon species has been made, so that the artificially bred Yangtze sturgeons can be developed to a third
generation in the laboratory (D. Li, Prinyawiwatkul, etal.,2021).
The first stock enhancement program by breeding and release in Brazil was carried out for non-native fishes in the
northwest region, which improved fishery yields significantly (Paiva etal.,1994). Nowadays, stock enhancement
Figure 15. Artificial breeding and releasing of Chinese sturgeon. (a) Artificially bred fry and juveniles of Chinese sturgeon.
(b) Release of artificially bred fry and juveniles of Chinese sturgeon into the Yangtze River. Photos are provided by China
Three Gorges Corporation.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
32 of 64
has become a mandatory management action, which is considered a primary measure to mitigate the impacts of
reservoirs on fish populations and protect ichthyofauna in Brazil (F.P.Arantes etal.,2011; Casimiro etal.,2022).
In China, dam construction in the Yangtze River basin has significantly affected natural reproduction processes
and decreased the fishery resources of FMCCs (Q. Peng et al., 2012). After impoundment of the TGR, the
resources of FMCCs have decreased by more than 90%, compared to that in the 1960s (Y. Yi & Wang,2009).
The development of artificial breeding techniques and large-scale implementations of stock enhancement have
significantly recovered the FMCCs resources in the Yangtze River basin. The annual release of over 10,000kg of
bloodstock of FMCCs has been carried out in the Shishou and Jianli sections of the middle Yangtze River since
2010, which has brought substantial economic, social, and ecological benefits (H. J. Chen,2019). The improved
FMCCs fishery has alleviated limitations depending only on the catches of natural FMCCs resources (H. J.
Chen,2019; J. Yang etal.,2013). In the Lancang River, China, fish breeding and release has been introduced for
almost all the dams, which has proven effective in recovering fish resources (H. Xu & Pittock,2018).
Overall, artificial breeding and release of target fish species is an important measure to recover endangered fish
species, maintain genetic diversity, support population expansion, and further preserve the ecosystem integrity of
dammed rivers (Le Luyer etal.,2017; Leinonen etal.,2020). However, monitoring data shows that the number of
juvenile Chinese sturgeon in the estuary of the Yangtze River has not shown a visible increase (Wei etal.,2004),
as there are only small-scale natural breeding activities in the spawning grounds below the Gezhouba Dam
(P.Zhuang etal.,2016). Thus, artificial breeding and release has not recovered the natural reproductive processes
of wild Chinese sturgeon, but only maintains their populations to some extent and prevents them from extinction.
In the artificial breeding process, the selection and renewal of breeding fishes are sometimes overlooked during
artificial reproduction, leading to parent fishes with inferior genetic characteristics arising from inbreeding. The
genetic introgression of the released populations has the potential to reduce the genetic diversity and impair the
performance of wild populations due to genetic drift (Lin etal.,2022). For instance, artificially bred FMCCs
with inferior genetic characteristics escaping from tributaries and connected lakes, have affected the natural
high-quality germplasm resource of FMCCs in the mainstream Yangtze River, further leading to a decrease in
the quality of the germplasm resource and adaptation ability to the wild environment of the wild FMCCs (H. J.
Chen,2019). Fish stocking activities can introduce exotic diseases and parasites into the water body, which is
potentially harmful to the endangered populations (J. Yang etal.,2013). Therefore, the genetic admixture between
artificially bred fishes and wild fishes could cause genetic contamination and affect the genetic structure as well
as the stability of wild populations, leading to genetic and ecological risks (Abdolhay etal.,2011). In addition,
there is competition between large-scale released hatchery fish and wild recipient fish, squeezing the population
of wild species (J. D. Bell etal.,2008). It is necessary to understand the carrying capacity of the receptor envi-
ronment and the size of the wild population before conducting artificial stocking and release, in order to reduce
the negative impacts and maximize the benefits (Agostinho etal., 2016). In particular, artificial breeding and
release of non-native species can lead to biological invasions, which will result in changes and declines in native
fish diversity (Bernery etal.,2022). For instance, tilapia, which possess a wide environmental adaptive ability,
have become the dominant species in reservoirs and lakes due to artificial enhancement of stocking, threatening
the survival of native fish species (Cucherousset & Olden,2011).
Long-term field monitoring has shown that most artificially bred fishes cannot reproduce naturally in the wild,
but merely maintain the size of the target fish population. Failure to reproduce a second generation naturally
leads to the challenge that the wild population of the fish species cannot increase, and thus the stability of the
wild population is vulnerable in the long-term. Therefore, it is important to investigate the mechanisms of natural
reproduction of the target fish species in dammed rivers, including both artificially bred and wild individuals, for
restoring populations. Moreover, due to the lack of long-term and continuous monitoring data, the quantitative
impacts of artificial breeding and release on river ecosystems remain unclear. It is essential to establish a risk
assessment system on genetic admixture and species invasion to evaluate quantitatively the negative effects of
artificial breeding and release. This could help improve conservation measures to increase the target fish popula-
tion and protect river biodiversity.
4.3. Nature-Based Solutions and Reservoir Ecological Operation
Nature-based Solutions (NbS) have been recognized as an umbrella concept to capture eco-friendly strategies
that mimic nature with broad public acceptance (Cohen-Shacham etal.,2016; Y. A. Song etal., 2019). NbS
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
33 of 64
follows the laws of nature to protect and restore degraded ecosystems as a whole, which can be applied in various
domains (Maes & Jacobs,2017; Nesshöver etal.,2017; Raymond etal.,2017). After the first appearance in the
official report from the World Bank(2008), the idea has gained increasing attention in scientific communities,
governmental agencies and non-governmental organizations worldwide (Cohen-Shacham etal.,2016; Griscom
etal.,2017; Krull etal.,2015). NbS is particularly suitable for watershed ecosystem protection and river ecosys-
tem remediation so as to preserve and restore the functions and valuable services that nature provides.
Conventional reservoir operations aim predominantly to maximize social-economic benefits, which may cause
severe damage to the ecological structure and functioning of river ecosystems. By contrast, ecological reser-
voir operations follow the idea of NbS, which aims to balance social-economic benefits and river ecosystem
requirements (X. Xia etal.,2009). Studies on reservoir ecological operation have spanned several decades, since
Schlueter(1971) proposed that reservoir operation schemes should take the habitat diversity of river ecosystems
into account when meeting the social-economic water demand. Many studies have been dedicated to developing
optimization models for reservoir ecological operation (D. A. Hughes & Ziervogel,1998; C. Liu etal., 2019;
Suen & Eheart,2006). The United States and Australia are two countries that have undertaken relatively early
implementations of reservoir ecological operation (Higgins & Brock,1999; A. J. King etal.,2010). In recent
years, China has made great efforts in reservoir ecological operations, which have made visible contributions to
restoring fish resources in dammed rivers (Q. Chen etal.,2021; C. Ma etal.,2020).
Flow regulation can create suitable hydrological conditions for fish to acquire their demand environment at
critical life stages, especially during spawning periods (Yin etal.,2011). The NFR, which shapes river species
community and ecosystem structure and function (Poff,2018), is crucial to riverine biodiversity and healthy
(Bower etal.,2022; Poff etal.,1997). Releasing ecological flow (Maavara etal.,2020; Poff,2018) is an impor-
tant measure of reservoir ecological operation (Y. Miao etal.,2020; X. Xia etal.,2009; Y. Xia etal., 2019). Q.
Chen etal.(2012), D. Chen, Chen, etal.(2016) proposed an operation scheme for the Qingshitan Reservoir in the
Lijiang River, China, to maintain a quasi-NFR in the downstream reach for fish conservation when meeting the
demands of irrigation, cruise navigation and water supply. Creating artificial floods through reservoir operation
has been used to stimulate spawning of drifting eggs in dammed rivers (C. Ma etal.,2020; Zhou etal.,2019). In
the Colorado River, artificial flood experiments have been conducted at the Glen Canyon Dam for many years,
which have successfully restored the endangered humpback chub (Gila cypha) and maintained populations of
other native fish species (Jacobson & Galat,2008; Melis etal.,2015; Yao etal.,2015). The Hume Reservoir in
the Murray River, Australia, rebuilds some small-to-medium floods, which modifies the timing of peaks to trig-
ger the spawning of native golden and silver perch, and increase the duration of floods to extend the recruitment
of species from mid-October to mid-December 2005 (A. J. King etal.,2010). In China, ecological operation of
the TGR has substantially increased the spawning of the FMCCs (Figure16). During the 4-day experiment of
ecological operation of TGR conducted in mid-June 2018, an initial outflow of 11,000m
3/s and a flow increment
in the range of 1,000 to 1,500m
3/(sd) was implemented, which formed a concentrated egg spawning activ-
ity of FMCCs in the Yichang section of the Yangtze River (C. Ma etal., 2020). In Brazil, reservoir operators
have recently proposed a flow regime that could restore flooding for 32 fish breeding sites in the Xingu River
(Moutinho,2023).
Discharge regulation alone is sometimes insufficient to meet the demand of fish reproduction, since fish spawn-
ing also requires suitable water temperature. Controlling the selective withdrawal device to adjust the water
temperature of the outflow from temperature-stratified reservoirs can improve the water temperature rhythm to
some extent for fish reproduction downstream of dammed rivers (Saadatpour etal.,2021). A selective-withdrawal
device was installed at the Shasta Dam in 1997 to manage the water temperature regime in the downstream reach
of the dam to meet the year-round thermal requirements of salmonid while fulfilling the obligation on water deliv-
ery, power generation and flood control. Its use resulted in a significant increase in the salmonid fish population
(Bartholow etal.,2001; Hanna etal.,1999). A water temperature regulation experiment was conducted in the
Xiluodu-Xiangjiaba cascade reservoirs in the Yangtze River in May 2017, which increased outflow temperature
and thereby facilitated fish spawning in the reaches downstream of the Xiangjiaba dam. About 10 million and 100
million fish eggs were monitored in the Yibin and Jiangjin section of the Yangtze River, respectively. In particu-
lar, the annual peak in fish spawning occurred during the period of the experiment (Ren etal.,2020).
Sediment regulation of reservoirs can change the morphology of the downstream reaches, and this potentially
affects the location and quality of fish habitats (W. Wang etal.,2012; H. Zhang, Jarić, etal.,2021). Sediment
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
34 of 64
regulation has been implemented at the Xiaolangdi Reservoir in the Yellow River since 2002, resulting in the main
channel of the lower reach of the dam being fully scoured (C. Miao etal.,2016). The regulation also brings abun-
dant nutrients and hence plankton blooms downstream, providing suitable feeding sites for pelagic and epipelagic
fish that congregate in the Yellow River estuary (L. Zheng etal.,2014). Some studies show that sediment flushing
could erode the areas of spawning and feeding grounds of fish, and even lead to fish mortality (Kong etal.,2017).
However, proper design of sediment flushing could control suspended sediment concentration and reduce adverse
effects on fish in the downstream reaches (Cattaneo etal.,2021). In addition, fish can possess a certain resistance
to the negative impacts caused by sediment flushing operations (Grimardias etal.,2017). The three-year field
investigation conducted from 2009 to 2011 concerning the controlled sediment flushing of a small reservoir on
the Adda River in Italy showed that fish resources in the downstream reach were not affected significantly (Espa
etal.,2015). Similarly, fish density showed no significant impairment during controlled sediment flushing oper-
ations conducted in 2016 at the Verbois Reservoir in Switzerland (Cattaneo etal.,2021).
Control of intermittent reservoir discharges is possible to reduce the damage of TDG supersaturation on fish.
Modeling results illustrate that intermittent discharges from the Bala Reservoir could diminish TDG supersatura-
tion and reduce the negative effects on fish in the Zumuzu River, China (J. Feng etal.,2014). Based on numerical
modeling of TDG, Wan etal. (2021) found that a proper discharge scheme for the Xiluodu Reservoir on the
Yangtze River could reduce the level and maximum residence time of TDG in downstream waters, thereby alle-
viating negative impacts on fish. The mixing of tail water and spill discharge can create areas of low TDG level,
providing shelter zones for fish to avoid damage from high TDG levels (Wan etal.,2020).
Reservoir ecological operation has developed from single factors to the coupling of multiple factors of fish phys-
ical habitat. W. He etal.(2020) proposed an operation model for the Sanbanxi Reservoir in the Yuanjiang River,
China, to meet the demand of outflow water temperature and downstream ecological flow. Z. Xu etal.(2017)
proposed an eco-friendly operation scheme considering flow velocity and water temperature demands of target
fish, which significantly facilitated the spawning of this fish species. In South Africa, an experiment has been
conducted at the Clanwilliam Dam on the Olifants River by creating small pulses of high flow and making the
water temperature at the spawning site reach above 19°C, which has resulted in a visible increase of successful
spawning activities of yellowfish (Barbus capensis) downstream of the dam (J. King etal.,1998). In the USA, a
Figure 16. Effect of reservoir ecological operation on fish spawning in dammed river. Adapted from Q. Chen etal.(2021). (a) The proposed ecological operation of
the Three Gorges Reservoir (TGR) for improving the spawning of four major Chinese carps (FMCCs) in the downstream river of the TGR. (b) The changes of spawned
eggs of FMCCs in the downstream river of the TGR before and after the proposed ecological operation. Data are available from China Three Gorges Corporation and
China National Environmental Monitoring Center. There was no measurement of spawned eggs during the TGR ecological operation in 2020 due to COVID-19.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
35 of 64
flow pulse has been released from the Gavins Point Dam, together with the improvement of water temperature, to
promote spawning of pallid sturgeon in the lower Missouri River (Jacobson & Galat,2008).
The effect of ecological operation of a single reservoir is sometimes limited. Joint operations of multiple reser-
voirs, which can be in a cascade along one river or are distributed on multiple rivers within the same watershed,
can better coordinate the social-economic interests and ecological requirements. Dalcin etal.(2022) reported a
methodological framework to guide operations of cascade reservoirs for rebuilding expected flow regimes. The
application to the reach between Porto Primavera Reservoir and the Itaipu Reservoir in the upper Paraná River
basin in Brazil provided suitable conditions for the successful recruitment of migratory fish species. Q. Chen
etal.(2013), D. Chen etal.(2015) proposed an adaptive operation scheme for two cascade reservoirs, Jinping-I
and Jinping-II, in the Yalong River, China, to meet the requirements of daily ecological flow and water tempera-
ture for the conservation of the indigenous fish species Schizothorax chongi in the dewatered river reach between
the two dams. D. Chen, Leon, etal.(2016), D. Chen, Leon, Engle, etal.(2017), and D. Chen, Leon, Hosseini,
etal.(2017) optimized the operation of 10 reservoirs in the Columbia River to maximize total power revenue
and fulfill the ecological flow requirements of fish, providing an important reference to ecological operation
of multiple reservoirs. Z. Jiang etal. (2019) developed a multi-objective optimization model to balance flood
control, power production, and ecological flow requirement of a large-scaled cluster of reservoirs with mixed
types in the Pearl River basin, China. The model results played a significant role in the operations to improve both
social-economic benefits and fish conservation.
Although there are studies and engineering practices of reservoir ecological operation concerning multiple
factors, such as flow and water temperature, the spatiotemporal matching of different factors with regards to
habitat requirements of different fish species and their life cycles remain challenging. In the joint operation of
cascade reservoirs, available studies mostly use water balance methods to link the reservoirs. In future, it is better
to adopt hydrodynamic models, which can simulate the distribution of flow fields, water temperature, and TDG
to achieve refined operation strategies for more effective conservation of fish. In addition, climate change could
bring large uncertainties to the inflow of reservoirs (Y. Wang etal.,2019), which implies that current reservoir
ecological operations based on historical data and deterministic models must be updated to incorporate inflow
uncertainties under future climate changes.
4.4. Habitat Compensation and Dam Removal in Tributaries
The shift from a lotic to a lentic environment of the reservoir after river damming leads to a permanent loss of
habitats for the maturation and spawning of fish species (Antonio etal.,2007; Liermann etal.,2012), and such
loss cannot mostly be remediated through reservoir ecological operations. The situation is more severe in rivers
with cascade dams, for instance the Lancang River, where an upstream dam is located in the backwater zone of
the nearest downstream reservoir, and the lentic section stretches for over one thousand kilometers. Under such
circumstances, relatively unaltered tributaries can provide possible alterative places to conserve indigenous species
of the dammed mainstream by serving as high-value natural surrogates or supplements that can restore some
function of the mainstem ecosystems (Neely etal.,2009; Nunes etal.,2015; Pracheil etal.,2009,2013). The
elements of natural flow fluctuations and the availability of food resources and shelter areas could be main-
tained in these tributaries. Moreover, diverse hydraulic conditions exist in the upper, middle, and lower tributary
reaches, giving the migratory or rheophilic fish species chances to colonize new habitats. In addition, for some
native fish species, their survival and life history requirements are directly related to intact longitudinal pathways,
including the possibility of migration into tributaries for reproduction and rearing (Da Silva etal.,2015). There-
fore, the use of unregulated tributaries to alleviate the adverse impact of impoundments in the mainstream has
recently been forwarded as a key alternative for fish conservation in dammed rivers (Figure17).
Successful spawning of humpback chub is found to be related to the migration of adult individuals from the
highly altered Colorado River to a relatively unregulated tributary, which offers necessary spawning habitat
and hydrological variability (Gorman & Stone,1999). Adults of American paddlefish prefer to migrate up the
unaltered Yellowstone River rather than the regulated Missouri River when they move above the confluence of
the two rivers, probably because the more natural flow pattern in the Yellowstone River tributary provides better
spawning conditions than the Missouri River (Firehammer & Scarnecchia,2006). In South America, recent
findings indicate that at least eight long-distance migratory species utilize alternative spawning and nursery
habitat in four tributaries of the upper Paraná River, after the construction of Porto Primavera Reservoir (Da Silva
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
36 of 64
etal.,2015,2019). The Congonhas River in Brazil provides possible reproductive routes and feeding sites for six
important rheophilic and migratory species, Piaractus mesopotamicus, Megaleporinus obtusidens, Prochilodus
lineatus, Salminus brasiliensis, Pinirampus pirinampu, and Pseudoplatystoma corruscans, after the construction
of Capivara Dam. These species are long-distance migrators and represent about 29% of all migratory species
inhabiting the upper Paraná River basin (Garcia etal., 2019). In China, indigenous fish species in the upper
Yangtze River and Lancang River are facing great threats incurred by the cascade dams. Conservations of the
indigenous fish in these severely impounded rivers often focus on their large tributaries, which remain natural or
are lightly dammed and have been recognized as alternative habitats for spawning and larvae development (L.
Tang etal.,2021). W. Cao(2000) and Park etal.(2003) suggested that three tributaries, the Jialing, Chishui and
Tuo rivers, in the upper Yangtze River can be a potential refuge for 22 endemic species, including Dabry's stur-
geon that is listed as a first-class protected animal in China. In the Lancang River, its tributary, the Luosuo River,
serves as an important habitat for the migration and spawning of red mahseer (Tor sinensis), one of the most
famous economic fishes of Yunnan province in southwest China (Hong etal.,2022; Y. Peng, Hong, etal.,2022).
These findings demonstrate that unregulated tributaries can provide habitats required by fish at different devel-
opment stages.
Preservation of free flow is essential for tributaries to serve as alternative habitats of fish in dammed mainstreams.
However, numerous tributaries of large rivers have been developed for individual or cascades of SHPs. In the
United States, there are more than 75,000 reservoirs, and many of them are impounded by small dams (<10m in
height) that are now aged and in disrepair (Ahearn & Dahlgren,2005). In China, more than 45,000 small dams
have been built by the end of 2012 to meet rural electricity demand (Ding etal.,2019; Hennig & Harlan,2018),
and approximately 6,590 small dams are reported to be out of service due to their age and loss of function
Figure 17. Habitat compensation and rehabilitation in dammed river. (a) Laomuhe Dam on the Heishui River, a tributary of upper Yangtze River. (b) The removal
of the Laomuhe Dam. (c) The Heishui River after the removal of the Laomuhe Dam. Photo by Lei Tang. (d) Habitat restoration measures in dammed river, including
ecological spur dike construction, creating diverse habitats with large woody debris, big rocks, pool and riffle. (e) Changes of fish diversity and mass per unit effort in
the Heishui River before and after the removal of Laomuhe Dam. Adapted from S. He etal.(2021).
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
37 of 64
(Gao etal.,2018). Removal instead of maintenance of these aged facilities has been growing around the world.
Since the early 1900s, nearly 1,800 small dams have been removed from rivers in the USA (Fox etal.,2022). In
Europe, at least 4,000dams and weirs have been removed since the mid-1990s (Kim & Choi,2019). Removal of
small dams in tributaries can open up spawning and rearing habitats in previously inaccessible regions upstream,
which has shown positive effects on fish diversity of the dammed mainstreams (Barbarossa etal.,2020; Lasne
etal.,2015; Magilligan, Graber, etal.,2016; O'Connor etal.,2015; Y. Peng, Hong, etal.,2022). Foley, Bellmore,
etal.(2017) reported that anadromous fishes, such as salmonids, passed the former dam site within days to
weeks after the removal of the Marmot dam in the Sandy River, USA, a tributary of the Columbia River. Simi-
larly, following the removal of small dams in the Jidu River, a tributary of the Lancang River, seven migratory
species recolonized newly accessible habitats (Hong etal.,2022). In addition to the direct effect of reconnection
and upstream access, dam removal has strong influences on fish physical habitat by changing sediment regime
and channel morphology, which increase the heterogeneity of hydraulic features and create new habitats for fish
species (Hatten etal.,2016; Im etal.,2011; Magilligan, Nislow, etal.,2016). The removal of a small dam in the
Heishui River (Figures17a–17c), a tributary of upper Yangtze River, has increased the percentage of suitable
spawning habitat for Jinshaia sinensis by a factor of four (Figure17e), due to the change in river morphology and
hydrological regimes following the removal (L. Tang etal.,2021).
Although removing small dams to restore the natural conditions of rivers is shown effective to conserve fishes
in dammed rivers, removals of large dams still face great challenges due to social-economic constraints and
intensive as well as long lasting ecological consequences (O'Connor etal.,2015). To restore the access of salmo-
nids to their spawning grounds in the mainstream of Elwha River, two large dams, the Elwha Dam and Glines
Canyon Dam, have been removed, resulting in increased species richness and functional stability of the riverine
ecosystem (Foley, Warrick, etal.,2017; Shaffer etal.,2018; Warrick etal.,2015). Similar achievements have
been made in the removal of the San Clemente Dam on the Carmel River, USA (Smith etal.,2020), which has
gradually improved the spawning grounds of steelhead salmon near the original dam site and in its downstream
reach (Harrison etal.,2018). Despite the visible ecological benefits, removal of large dams often comes at a huge
cost, and long-term consequence of fish community demands further investigation.
After dam removal, naturalized artificial habitat, following the NbS concept, can be created to expand living spaces
of fish (Figure17d). The addition of gravel to rivers, known as gravel augmentation, is an early attempt at artificial
habitat creation, which has proven to increase the available spawning grounds for Atlantic salmon and brown trout
in regulated rivers (Barlaup etal.,2008; Pulg et al.,2008). Three types of artificial habitats (straw bales, straw
tubes and moss tubes) have been implemented to enhance egg production of Galaxias maculatus, an important
fishery species in New Zealand (Hickford & Schiel,2013). In the Youjiang River, a tributary of the Pearl River,
China, artificial habitats made of bamboo and palm slices have been deployed to serve as spawning grounds for
fish that produce sticky eggs and as refuges that improve the survival rates of juvenile fishes (D. Guo etal.,2020).
The installation of large woody debris can also restore degraded river ecosystems due to dam construction through
“rewilding,” which has been proven to significantly improve the abundance of food resources and thereby increase
the population of fishes in the restored reaches (Thompson etal.,2018). In the lower Mulde River, Germany, fish
abundance increased nearly 10-fold eight months after the installation of large wood (Anlanger etal.,2022).
Tributaries cannot offer an identical replacement for the degraded habitats of previously undammed mainstreams,
as the former have lower discharge, longitudinal distance, morphological variability, and habitat complexity than
the latter. Discharge and its variability are essential to provide flow-related cues that initiate fish maturation and
spawning, or create conditions for recruitment of larvae and juveniles. The longitudinal distance is critical for
fish species that require long egg-drifting distances for survival. To date, quantitative methods for determining
suitable tributaries for fish spawning, foraging and refuging have not been established. Future studies should
attempt to coordinate the relations between conservation efforts in the mainstream channel and its tributaries to
help achieve maximum effectiveness. In addition, despite the removal of a large number of dams worldwide over
the past decades, our knowledge concerning these effects is still limited due to the lack of long-term monitoring
data, which is critical to quantify the rate, magnitude and sequence of tributary habitat recovery to dam removal.
4.5. Efficiency Assessment on Conservation Measures
A variety of measures to conserve fish species impacted by river damming are available and each measure can
be effective under specific conditions. However, a challenge is to select appropriate measures based on real
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
38 of 64
situations and cost-effectiveness. Table2 summarizes the major conservation measures and their advantages,
disadvantages as well as applicable conditions. We acknowledge that environmental decisions are complex and
require a nuanced understanding of local context and will almost always involve trade-offs.
Fishways are of benefit to fish species with strong swimming ability, but are mainly suitable for dams where
the water head is <60m. Fish lifts are space saving and mainly suitable for concrete gravity dams where the
Conservation measures Advantages Disadvantages Suitability of applications
Fish passage facilities Fishway 1. Promote timely fish passage Only effective for fish species
with strong swimming
abilities
All types of dams with a water
head between upstream and
downstream within 60m
2. Reduce damage to fish bodies
Fish lifts 1. Space-saving 1. Complex mechanical facilities Concrete gravity dams with
a water head between
upstream and downstream
over 60m
2. Higher possibility of failure
2. Easy to arrange in the dams 3. Limited number of passed
fishes
Fish collection and
transportation system
1. Flexible 1. High power consumption All types of dams with a water
head between upstream and
downstream over 60m
2. Adjust the replenishment
flow according to the fish
preference
2. Complex operation and
management
3. Fish mortality during
transportation
Fish-friendly turbine passage Ecological friendly turbines can
allow for the safe passage of
small objects
Physically damage or traumatize
fishes
Low and medium water head
hydropower plants
Juvenile bypass system 1. Long length, with good
ecological landscape function
1. Need more space All types of dams with a water
head between upstream
and downstream within
30m, and rely on tributary
projects
2. Effectively reduce the mortality
rate of fish passing through
the dam
2. High requirements of local
terrain conditions
3. Easy to adjust and expand the
trial run after completion
Artificial breeding and release 1. More convenient operation and
management
1. Longer exploration time All dammed rivers, especially
for conservation of
endangered or economic
fish species
2. Relatively mature technology 2. Reducing survival chance of
wild individuals
3. Protect fish species and increase
fish populations
3. Dilution of genetic diversity in
wild populations
Reservoir ecological operation 1. Restore the structure and
function of river ecosystems;
1. Loss of social and economic
benefits;
All types of dams with large or
medium-sized reservoirs
2. Difficulties in coordinating
departments;
2. Promote the fish spawning 3. Difficult to meet the needs
of different fish species
simultaneously
Habitat compensation in tributaries 1. Provide compensatory
habitat for fish affected by
hydropower in the mainstream;
1. Limited compensation ability
of the tributaries without
sufficient free-flowing length;
The presence of tributaries
with high habitat similarity
to the dammed mainstream
2. Restore the natural connectivity
of the river;
3. Promote changes in
the diversity of river
characteristics;
2. Demolition of small dams
in tributaries causes fish
casualties and adenosis
4. Promote fish reproduction
Table 2
Summary of Advantages, Disadvantages and Applicable Conditions of Fish Conservation Measures
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
39 of 64
water head is >60m, but they are mechanically complex and have limited passage capacity. Fish collection and
transportation systems are flexible in space and time, and suitable for dams with water head >60m, but they are
complicated to operate and usually cause a high mortality rate during transportation. Fish-friendly turbines are
designed for low and medium water head hydropower plants, which can decrease the mortality and mechanical
damage to fish (Hogan etal.,2014; Pracheil etal.,2016; Watson etal.,2022). Juvenile bypass systems are mostly
used where the water head is <30m and there is a tributary present. Artificial breeding and release are effective to
protect endangered fish species and restore the resources of economic fish species, but it could affect the survival
and genetic diversity of the wild population. Reservoir ecological operation is an effective non-engineering
method for fish conservation in dammed rivers and is particularly applicable to large and medium-sized reser-
voirs, but it may cause certain loss of social-economic benefits. Habitat compensation in tributaries can be a
potential approach to conserve fish species that permanently lose their original habitats in the impounded main-
stream upstream of high dams, but their effectiveness depends on the ecological status of the tributaries and the
future development of the mainstream.
Quantitative evaluation on the effects of conservation measures is essential to select proper approaches and
improve their efficiencies. Table3 summarizes the major indicators and methods to assess these different conser-
vation measures.
Passage effectiveness and efficiency are the main indicators to evaluate the capability of fish passage facilities
(Bravo-Córdoba etal.,2021). Passage effectiveness is used to describe qualitatively the potential effect of fish
passage facilities on fish proliferation by checking that the passage facility is capable of letting all target species
pass through within the range of environmental conditions observed in nature during the migration period.
Passage efficiency is a quantitative evaluation indicator of fish passage effectiveness, which is defined as the
ratio of the number of species and quantities of fish individuals that migrate upstream through the fish passage
facility to the number of species and quantities of fish individuals that demand to pass the dam in a specific period
(Larinier,2008). The average efficiency of fish passage facilities is 50–60% (Hershey,2021). Evaluation of the
effectiveness of artificial breeding and release depends on the specific objectives of the measure, and generally
focuses on the growth of fry and the contribution to target fish resources as well as the related economic, ecolog-
ical and social benefits (Rytwinski etal.,2021). Ecological restoration aims at conserving endangered species,
and the evaluation mainly focuses on the survival rate and natural reproduction of artificially hatched fry after
Conservation measures Evaluation indicators Description and calculation
Fish passage facilities Fish passage effectiveness The potential effect of fish passage facilities on fish proliferation by checking that
the passage facility is capable of letting all target species pass through within the
range of environmental conditions observed in nature during the migration period.
Calculations are based on the monitoring data of fish species, numbers, sizes, life
stages, and behavior under the specific conditions of the operating fish passage
Fish passage efficiency The ratio of the number of fish species and quantities of fish individuals that migrate
upstream through the fish passage facility to the number of fish species and
quantities of fish individuals that demand to pass the dam in a specific period
Artificial breeding and releasing Survival rate after releasing Mark the fry to be released using marking techniques, and calculate the percentage of
fish containing the number of marks in the fish catch by recapture
Catch rate The percentage of the catch in a certain water body and a certain period of time to the
total resource of the fishing object in the water body during the same period
Reservoir ecological operation Ecological flow replenishment Replenishing the flow through reservoir operations to meet the minimum flow
demand of downstream ecosystems
Spawning volume The spawning volume that is monitored for the spawning grounds during the reservoir
ecological operation
Habitat compensation in tributaries Habitat diversity The overall richness of various types of habitats that accommodate various organisms
Fish diversity Biodiversity of indigenous fish species, including species richness, species
abundance, and phylo genetic diversity
Fish populations The total number of fishes that inhabits a certain area
Table 3
Summary of Evaluation Indicators for Fish Conservation Measures
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
40 of 64
release (Lyu etal.,2021). Resource restoration aims at restoring fishery resources and improving the economic
benefits of fisheries, and thus the evaluation has mainly focused on the catch rate. The evaluation indicators for
reservoir ecological operation include replenishment of ecological flow and spawning volume of target fish in the
downstream reach of the dam (J. Li etal.,2019), particularly, the amount of spawning volume during reservoir
operation can well evaluate the effectiveness of ecological dispatch. The primary objective of habitat compen-
sation in tributaries is to protect the biodiversity of indigenous fish species, and the indicators for effectiveness
evaluation include habitat diversity, fish diversity and fish populations.
There are significant differences in the cost-benefits of different fish conservation measures for dammed rivers
(Table4). Fish passage has significant variations in cost-benefits, depending on the types of facilities. The Denil
fishway and fish lift have relatively low cost-benefits, while the juvenile bypass system and fish collection system
have relatively high cost-benefits. The Denil fishway costs approximately USD 124,000 per vertical meter and
has low maintenance as well as operation costs, but their average passage efficiency is only 16% (Noonan
etal.,2012). Fish lifts are expensive to build and operate, and costs roughly USD 2.4 million to install and an
annual maintenance charge of 5% (Noonan etal.,2012), although it has relatively high passage efficiency. Fish
collection systems need to temporarily preserve the transferred fish during long-distance transport, leading to
high mortality loss and operational costs. Juvenile bypass systems have relatively low construction and opera-
tional costs, and the average passage efficiency can reach 70%. However, it demands more space, which limits
its applicability.
The disposable investment of artificial breeding and release ranges from millions to hundreds of millions of USD,
and the cost-benefits depends on the bred fish species and scales of release (H. J. Chen,2019). For instance,
Chinese sturgeon spawns in the autumn, so they need to be incubated and reared indoor using heating devices to
promote their growth, which is expensive to maintain (Wei etal.,2004). Ecological restoration for the endangered
Conservation measures Costs Benefits
Fish passage facilities 1. Denil fishway: approximately US $124,000 per vertical
meter with low maintenance and operation costs
1. Denil fishway cost: average passage efficiency is only 16%
2. Fish lifts: roughly US $2.4 million to install and annual
maintenance charge of 5%
2. Fish lifts: relatively high operation efficiency
3. Fish collection systems: high mortality loss and operational
cost
3. Fish collection systems: high passage efficiency
4. Juvenile bypass system: relatively low construction and
operation cost
4. Juvenile bypass system: average passage efficiency is 70%
Artificial breeding and releasing 1. Range from millions to hundreds of millions of USD for
different scales of releasing and different bred fish species
(H. J. Chen,2019)
1. “Ecological restoration”: conserved the endangered native
fish populations and prevent their extinction. “Resource
restoration”: improved fishery with high economic values
2. A total of 31.55 million endemic fishes, including 24.16
million economic fishes which cost about US $0.48 billion
2. The number of fish species increased by 18 species in the
Yangtze River basin, effectively restoring the fishery
resources of the Yangtze River (Sun & Wang,2020)
Reservoir ecological operation 1. The power generation of cascade reservoirs decreased by
1.76%
1. The ecological flow coordination degree increased by
17.45%, promoting the spawning of the four major carps
(Dai etal.,2022)
2. The Gezhouba Hydropower Station lost 0.15% of its power
generation
2. The suitability of the spawning ground for Chinese
sturgeon increased by 39% (Y. Y. Wang etal.,2013)
3. The loss of power generation benefits is about 2.5% 3. Protect at least 50% of the target fish habitat in the river
(D. Chen etal.,2014)
Habitat compensation in tributaries 1. The Waterworks Dam was removed at a cost of US $0.214
million
1. Two years after the removal of the dam, the number of fish
species at the original site of Waterworks Dam increased
from the previous 11 to 26 species (Catalano etal.,2007)
2. The Marmot Dam was removed at a cost of US $4.86 million 2. The removal of the dam restored nearly seven miles of
river to migratory habitat for steelhead, Chinook salmon
and coho salmon (Xiao,2021)
Table 4
Summary of the Costs and Benefits of Conservation Measures
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
41 of 64
native fish species can prevent their extinction in dammed rivers, and the value cannot be simply capitalized.
Resource restoration for fishery improvement in dammed rivers can bring high economic values. From 2005 to
2018, a total of 31.55 million endemic fishes, including 24.16 million economic fishes that cost about US $0.48
billion, have been released into the upper Yangtze River. The release of economic fishes has effectively restored
fishery resources, and the release of rare fishes has increased the number of fish species from 22 in 2006 to 40
in 2018 (Sun & Wang,2020).
It is widely perceived that reservoir ecological operations may sacrifice social-economic interests due to the
dedicated discharge of ecological flow or water head loss for withdrawing temperature stratified water. The
temperature regulation device on the Shasta Dam resulted in an estimated US $63 million loss of hydropower
revenue between 1987 and 1996 (Hallnan etal.,2020). However, many studies have shown that by optimizing
the operation of reservoirs, the ecological benefits can be ensured with a marginal loss, or even an increase of
socio-economic benefits. Under the optimal operation scheme, the Jinping cascade reservoirs in the Yalong River,
China, only sacrifice annually 2.5% hydropower production to conserve more than 50% habitat for Schizothorax
chongi in the dewatered reach (D. Chen etal.,2014). In the upper Yangtze River, under an optimal operation
scheme, the hydropower production loss of the cascade dams is only 1.76%, while the fulfillment of ecological
flow increases 17.45%, which greatly promoted the spawning of FMCCs (Dai etal.,2022). X. Wang etal.(2020)
proposed an optimized operation for the Three Gorges and Gezhouba cascade reservoirs that the total hydro-
power production increased by 250,089.2MW·h and the area of suitable spawning grounds of Chinese sturgeon
increased by 2.16%. Cioffi & Gallerano(2012) optimized hydropower production and fish habitat protection for
the Pieve di Cadore Reservoir in the Piave River, Italy, and the results showed that the area of fish habitat could
be increased with little loss of power generation. W. Chen & Olden(2017) designed a reservoir release scheme,
which could create proper conditions to favor native over non-native fish, while human water demands were
barely sacrificed. In general, reservoir ecological operation is an effective non-engineering conservation measure
with high cost-benefits.
Studies have shown that the cost of removing small weirs (≤3.0m in height) is US $69,000 on average or US
$23,000 per meter height, which is less than 20% of the cost of building a fish passage or less than 12% of the
cost of building a fish ladder (Garcia de Leaniz,2008), and dam removal can significantly increase fish species
richness. The removal of the Waterworks Dam in the Baraboo River, USA, a tributary of the Wisconsin River,
cost approximately US $0.214 million in 1998. Two years after the removal, the number of fish species increased
from 11 to 26 at the original dam site (Catalano etal.,2007). The Marmot Dam in the Sandy River, a tributary of
the Columbia River, was removed at a cost of approximately US $4.86 million. The removal has restored nearly
seven miles of river habitat for migratory fishes of steelhead, Chinook and coho salmons (Xiao,2021). In general,
habitat compensation and removing small dams in tributaries is likely a cost-effective approach to conserve fish
diversity in dammed mainstreams.
5. Future Perspectives
5.1. Strategic Plans for Development and Conservation
Mitigation actions should be taken to minimize the potential impacts on fish during the complete process of
river damming, including planning and operation. Planning of dams should be conducted at a system level to
ensure that decisions are made in a more holistic manner. Prior to dam construction, the intensity of hydro-
power development should be determined at a basin scale, in order to balance river ecosystem conservation and
economic benefits. It is also imperative to develop strategic dam planning, especially at the basin or regional
scale, by performing multi-criteria optimization schemes (Flecker etal.,2022). Adequate investigations should
be conducted to identify the siting of dams so as to minimize impacts on fish spawning, feeding and wintering
grounds within the local hydro-geophysical constraints. Research on the impacts of dam construction on river
ecosystem services, such as water supply, sediment transport, and biodiversity maintenance, should be strength-
ened. The planning of dams on mainstream and tributary rivers must be collaborative, and consider the optimal
combination of high dams, low dams and run-of-river dams, which is important to reduce impacts on the river
ecosystem at a basin scale (Couto and Olden,2018; Couto etal.,2021; Schmitt etal.,2019). For fish that have
lost their habitats in a mainstream that has been dammed, it is valuable to investigate the possibility of restoring
tributaries to provide fish with alternative habitats.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
42 of 64
Although there are plenty of studies concerning native fish conservation in dammed rivers, most focus on a
specific issue (e.g., such as fish passage or habitat restoration), and comprehensive conservation strategies that
consider the habitat requirements of fish species throughout their life history are lacking. Existing conservation
measures that aim at restoring the physical habitats for fish have had limited effectiveness, and there is demand
for more studies to refine these measures and ensure they better achieve their intended goals. Specifically, the
effectiveness and efficiency of the measures need to be assessed quantitatively by using long-term monitoring
data of dammed rivers, so that emerging problems can be identified in a timely manner, and thereby aid improved
designs for more efficient measures. It is of great value to develop a framework to assess systematically the
effectiveness and efficiency of conservation measures for target fish species from the genetic to population,
metapopulation, community, and ecosystem levels. NbS have shown a high potential to conserve fish species in
dammed rivers and deserve more attention in future studies. Moreover, it is important to incorporate more target
fish species when investigating and planning a conservation program. Finally, consideration of hydrokinetic
energy and submerged turbines is valuable, as these can generate electricity without many of the drawbacks of
dams, although the power yield is less.
5.2. Long-Term Systematic Observations
One important aspect of the knowledge gap concerning the impacts of river damming on fish and associated
conservation measures is the lack of long-term monitoring data. Compared to the long-term records of natu-
ral rivers from gauge-stations, similar data for dammed rivers are relatively short-term. Dam construction also
causes inconsistencies between data before and after impoundment. Changes in river morphology are usually
slow, and thus the evolution of channel morphology after river damming, and its impacts on fish physical habitats,
demand longer monitoring periods. Hysteresis exists in the response of river ecosystems to hydro-geophysical
changes, which also requires sufficiently long observations. TDG supersaturation only occurs during the occa-
sional flood discharge of dams, providing few monitoring opportunities for data collection. It is also difficult, if
not impossible, to obtain the near-field data of TDG. Moreover, the changes to the physical habitat and adaptation
of fish species in dammed rivers may interact with each other, and these changes increase the complexity of eval-
uating the impact of dam construction on fish communities and developing conservation measures. Therefore,
there is an urgent need to establish dedicated monitoring networks in dammed rivers to strengthen the collection
of long-term and systematic data to improve our understanding of the impacts of river damming on fish. Future
hydro-geophysical monitoring of large rivers can be revolutionalized by observing from satellites, such as the
SWOT mission. Emerging technology, such as eDNA, otolith microchemistry, and biotelemetry, can be used
to characterize the dynamics of fish communities in dammed and undammed rivers, thereby providing useful
comparisons. Such long-term monitoring data concerning hydro-geophysical conditions and fish communities
will deepen knowledge on the impacts of river damming on fish, which will then provide a fuller scientific basis
for developing conservation strategies.
Available information on the impact of dams on fish is based mainly on statistical analyses of data from field
observations and laboratory experiments concerning the relationships between key hydro-geophysical factors and
various fish-related endpoints. Some studies examine the behavioral responses of target fish species to variations
of key hydro-geophysical factors in the laboratory to establish suitability curves, which are then used to develop
fish habitat models for impact assessment or prediction. These studies have made significant contributions to
evaluating the impacts of river damming and the design of fish conservation measures. In future, the adoption of
genomic, transcriptomic, proteomic, metabolomic and bioinformatic methods can be a viable approach to inves-
tigate the physiological mechanisms of how altered hydro-geophysical conditions affect gonad development, sex
differentiation, gene regulation, and gene expression of target fish species (Natri etal.,2019; Ortega-Recalde
etal.,2020). In addition, neurotoxic effects of river damming on fish behavior, personality and cognition could, in
turn, potentially generate feedback loops that may amplify the effects on fish. Therefore, integrative approaches
that combine field observations with novel technologies (e.g., molecular omics techniques, biotelemetry) are
recommended to bridge the knowledge gap in assessing river damming impacts on fish. This will enable the
development of more reliable models to predict long-term consequences, and thus support the implementation of
effective conservation measures.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
43 of 64
5.3. Complex Effects of Climate Change and Land Cover Changes
Climate change has dramatic impacts on river hydrology and thermal regimes, and thus affects the migration,
reproduction, growth and distribution of fish species in various ways. Regional patterns of warming-induced
changes in surface hydroclimate are complex, with evidence of increases and decreases in the magnitude of
precipitation and runoff, as well as frequency, depending on the local context (Milly etal.,2005). The mean and
high water temperature of global rivers are projected to increase under climate change, with the impact varying
between regions and across seasons. Climate change may thus exacerbate the effect of river damming on fish
habitats. Reservoir operation has disturbed natural hydrological regimes and riverine hydrodynamic conditions,
and climate change could bring additional alterations by affecting the patterns of global precipitation and snow-
melt. Climate warming may increase river water temperatures, prolong stratification and decrease vertical mixing
in reservoirs. Such prolonged stratification intensifies anoxia, which promotes the release of nutrients from sedi-
ment and stimulates eutrophication in reservoir waters, thus affecting fish community structure and food web
dynamics. Meanwhile, increases in water temperature reduce the dissolved oxygen content of water bodies, and
the stratification of reservoirs intensifies the deficiency of dissolved oxygen in the bottom layer, which affects
the survival of benthic fish species. However, reservoir operation can also mitigate impacts of future climatic
variability and climate change on fish habitats, for example, by releasing cold water to offset impacts of warming
temperatures on cold-water fish species (Benjankar etal.,2018). Climate change and reservoir retentions jointly
influence river sediment regime, which affects river morphology and substrate composition, thereby impacting
on fish habitat (J. Li etal.,2021). Future studies concerning the effects of climate change on fish should take into
account changes in watershed soil erosion, sediment flux and nutrient loss, which play an increasingly important
role in river ecosystems, when assessing the impacts of dams on fish (Best etal.,2022; J. Li etal.,2021). Due
to the lack of integrated data concerning climate change, dam regulation and the evolution of fish communities,
the interactive multi-stressor impacts of climate change and river damming on fish demand further investigation.
Climate change and dam construction also affect land use, which can directly or indirectly impact river ecosys-
tems. For example, climate change affects the amount and distribution of vegetation, agriculture and forestry, and
dam construction can promote rapid urbanization along the river. This significantly increases impervious surface
areas that lead to higher flood flows and earlier flood timing, and thus intensifies the impact on fish in dammed
rivers. The development of industrial and residential areas results in increased discharge of industrial wastewater
and domestic sewage, which could change local water temperatures and nutrient levels, and thereby affect fish
migration, spawning and feeding. Therefore, the impacts of river damming on fish could become extraordinarily
complex under the effects of climate and land use change. This highlights the need to consider the full range of
stressors affecting rivers, identify the major factors, and assess both their interactions and timescales of their
effects in future studies (Best & Darby,2020). Advances in numerical modeling and data collection could be
used to develop virtual watersheds, where different climate change forcings can be used to simulate processes in
river-reservoir systems (Benjankar etal.,2018; Tranmer etal.,2020). Preliminary implementations of this meth-
odology have shown its benefits in understanding the impact of dam operations on fish habitat.
Data Availability Statement
Most of data supporting the figures are available via the cited references. Data of Figures1a and1b are avail-
able through Lehner etal.(2011). Data of Figures1c and1d are available through Zarfl etal.(2015). Data of
Figure3a are calculated from Hydrological Data of Changjiang River Basin in Annual Hydrological ReportP.R.
China, including average monthly discharge data of Pingshan station (2007–2010) and Xiangjiaba hydrological
station (2016–2020). Data of Figure 3b are available from Hydrological Data of Changjiang River Basin in
Annual Hydrological ReportP.R. China, including average monthly discharge data of Ningnan hydrological
station in 2015 and 2019. Data of Figure3c are from the authors (Q. Li,2023). Data of Figure3d are available
from Q. Chen etal.(2021). Data of Figure5a are calculated from Hydrological Data of Changjiang River Basin
in Annual Hydrological Report P.R. China, including average monthly water temperature data of Pingshan
station (2007–2010) and Xiangjiaba hydrological station (2016–2020). Data of Figure5c are available through
T. Li etal.(2021). Data of Figure6b are available through Y. Wang etal.(2015). Data of Figure16 are available
through Q. Chen etal.(2021). Data of Figure17e are available through S. He etal.(2021).
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
44 of 64
References
Abdolhay, H. A., Daud, S. K., Ghilkolahi, S. R., Pourkazemi, M., Siraj, S. S., & Satar, M. K. A. (2011). Fingerling production and stock enhance-
ment of Mahisefid (Rutilus frisii kutum) lessons for others in the south of Caspian Sea. Reviews in Fish Biology and Fisheries, 21(2), 247–257.
https://doi.org/10.1007/s11160-010-9163-9
Abul-Atta, A. (1978). Egypt and the Nile after the construction of high Aswan dam (p.222). Study for Ministry of Irrigation.
Addy, S., Cooksley, S., Dodd, N., Waylen, K., Stockan, J., Byg, A., etal. (2016). River restoration and biodiversity: Nature-based solutions for
restoring the rivers of the UK and Republic of Ireland. In The International Union for the Conservation of Nature (IUCN) National Committee
UK (NCUK). Retrieved from https://portals.iucn.org/library/node/46347
Agostinho, A. A., Gomes, L. C., Santos, N. C. L., Ortega, J. C. G., & Pelicice, F. M. (2016). Fish assemblages in Neotropical reservoirs: Coloni-
zation patterns, impacts and management. Fisheries Research, 173, 26–36. https://doi.org/10.1016/j.fishres.2015.04.006
Ahearn, D. S., & Dahlgren, R. A. (2005). Sediment and nutrient dynamics following a low-head dam removal at Murphy Creek, California.
Limnology & Oceanography, 50(6), 1752–1762. https://doi.org/10.4319/lo.2005.50.6.1752
Ahmad, S. K., Hossain, F., Holtgrieve, G. W., Pavelsky, T., & Galelli, S. (2021). Predicting the likely thermal impact of current and future dams
around the world. Earth's Future, 9(10), e2020EF001916. https://doi.org/10.1029/2020EF001916
Alabaster, J. S. (1990). The temperature requirements of adult Atlantic salmon, Salmo salar L., during their upstream migration in the River Dee.
Journal of Fish Biology, 37(4), 659–661. https://doi.org/10.1111/j.1095-8649.1990.tb05901.x
Algera, D. A., Kamal, R., Ward, T. D., Pleizier, N. K., Brauner, C. J., Crossman, J. A., etal. (2022). Exposure risk of fish downstream of a hydropower
facility to supersaturated total dissolved gas. Water Resources Research, 58(6), e2021WR031887. https://doi.org/10.1029/2021WR031887
Almeida, P.R., Arakawa, H., Aronsuu, K., Baker, C., Blair, S., Beaulaton, L., etal. (2021). Lamprey fisheries: History, trends and management.
Journal of Great Lakes Research, 47, 159–185. https://doi.org/10.1016/j.jglr.2021.06.006
Almeida, R. M., Hamilton, S. K., Rosi, E. J., Barros, N., Doria, C. R., Flecker, A. S., etal. (2020). Hydropeaking operations of two run-of-river
mega-dams alter downstream hydrology of the largest Amazon tributary. Frontiers in Environmental Science, 8, 120. https://doi.org/10.3389/
fenvs.2020.00120
Alsip, P.J., Zhang, H., Rowe, M. D., Rutherford, E., Mason, D. M., Riseng, C., & Su, Z. (2020). Modeling the interactive effects of nutrient loads,
meteorology, and invasive mussels on suitable habitat for Bighead and Silver Carp in Lake Michigan. Biological Invasions, 22(9), 2763–2785.
https://doi.org/10.1007/s10530-020-02296-4
Anderson, D., Moggridge, H., Warren, P., & Shucksmith, J. (2015). The impacts of ‘run-of-river’ hydropower on the physical and ecological
condition of rivers. Water and Environment Journal, 29(2), 268–276. https://doi.org/10.1111/wej.12101
Anderson, E., Jenkins, C., Heilpern, S., Maldonado-Ocampo, J., Carvajal-Vallejos, F., Encalada, A., et al. (2018). Fragmentation of
Andes-to-Amazon connectivity by hydropower dams. Science Advances, 4(1), eaao1642. https://doi.org/10.1126/sciadv.aao1642
Anlanger, C., Attermeyer, K., Hille, S., Kamjunke, N., Koll, K., König, M., etal. (2022). Large wood in river restoration: A case study on the
effects on hydromorphology, biodiversity, and ecosystem functioning. International Review of Hydrobiology, 107(1–2), 34–45. https://doi.
org/10.1002/iroh.202102089
Antonio, R. R., Agostinho, A. A., Pelicice, F. M., Bailly, D., Okada, E. K., & Dias, J. H. P. (2007). Blockage of migration routes by dam construc-
tion: Can migratory fish find alternative routes? Neotropical Ichthyology, 5(2), 177–184. https://doi.org/10.1590/S1679-62252007000200012
Aprahamian, M. W., Evans, D. W., Briand, C., Walker, A. M., McElarney, Y., & Allen, M. (2021). The changing times of Europe's largest
remaining commercially harvested population of eel Anguilla anguilla L. Journal of Fish Biology, 99(4), 1201–1221. https://doi.org/10.1111/
jfb.14820
Arantes, C. C., Laufer, J., David, M., Moran, E. F., Claudia, M., Jynessa, L., etal. (2021). Functional responses of fisheries to hydropower dams
in the Amazonian floodplain of the Madeira River. Journal of Applied Ecology, 59, 680–692. https://doi.org/10.1111/1365-2664.14082
Arantes, F. P., Santos, H. B., Rizzo, E., Sato, Y., & Bazzoli, N. (2011). Collapse of the reproductive process of two migratory fish (Prochilodus
argenteus and Prochilodus costatus) in the Três Marias reservoir, São Francisco river, Brazil. Journal of Applied Ichthyology, 27(3), 847–853.
https://doi.org/10.1111/j.1439-0426.2010.01583.x
Arismendi, I., Penaluna, B., Gomez-Uchida, D., Di Prinzio, C., Rodríguez-Olarte, D., Carvajal-Vallejos, F. M., etal. (2019). Trout and char of
south America. In Trout and char of the world (pp.279–311).
Arthington, A., Dulvy, N., Gladstone, W., & Winfield, I. (2016). Fish conservation in freshwater and marine realms: Status, threats and manage-
ment. Aquatic Conservation: Marine and Freshwater Ecosystems, 26(5), 838–857. https://doi.org/10.1002/aqc.2712
Atkinson, S., Bruen, M., O'Sullivan, J. J., Turner, J. N., Ball, B., Carlsson, J., etal. (2020). An inspection-based assessment of obstacles to
salmon, trout, eel and lamprey migration and river channel connectivity in Ireland. Science of the Total Environment, 719, 137215. https://doi.
org/10.1016/j.scitotenv.2020.137215
Auer, S., Hayes, D. S., Führer, S., Zeiringer, B., & Schmutz, S. (2022). Effects of cold and warm thermopeaking on drift and stranding of juvenile
European grayling (Thymallus thymallus). River Research and Applications, 39(3), 1–11. https://doi.org/10.1002/rra.4077
Avella, M., Berhaut, J., & Bornancin, M. (1993). Salinity tolerance of two tropical fishes, Oreochromis aureus and O. Niloticus. I. Biochem-
ical and morphological changes in the gill epithelium. Journal of Fish Biology, 42(2), 243–254. https://doi.org/10.1111/j.1095-8649.1993.
tb00325.x
Baisre, J. A., & Arboleya, Z. (2006). Going against the flow: Effects of river damming in Cuban fisheries. Fisheries Research, 81(2–3), 283–292.
https://doi.org/10.1016/j.fishres.2006.04.019
Ban, X., Diplas, P., Shih, W., Pan, B., Xiao, F., & Yun, D. (2019). Impact of Three Gorges Dam operation on the spawning success of four major
Chinese carps. Ecological Engineering, 127, 268–275. https://doi.org/10.1016/j.ecoleng.2018.10.011
Bao, H., Zhang, S., Tang, C., & Yang, X. (2022). Effect of dam construction on spawning activity of Yellow River carp (Cyprinus carpio) in the
lower Yellow River. Frontiers in Earth Science, 10, 975433. https://doi.org/10.3389/feart.2022.975433
Barbarossa, V., Schmitt, R. J. P., Huijbregts, M. A. J., Zarfl, C., King, H., & Schipper, A. M. (2020). Impacts of current and future large dams
on the geographic range connectivity of freshwater fish worldwide. Proceedings of the National Academy of Sciences of the United States of
America, 117(7), 3648–3655. https://doi.org/10.1073/pnas.1912776117
Barlaup, B. T., Gabrielsen, S. E., Skoglund, H., & Wiers, T. (2008). Addition of spawning gravel—A means to restore spawning habitat of atlantic
salmon (Salmo salar L.), and anadromous and resident brown trout (Salmo trutta L.) in regulated rivers. River Research and Applications,
24(5), 543–550. https://doi.org/10.1002/rra.1127
Barrios-O’Neill, D., Dick, J. T., Emmerson, M. C., Ricciardi, A., & MacIsaac, H. J. (2015). Predator-free space, functional responses and biolog-
ical invasions. Functional Ecology, 29(3), 377–384. https://doi.org/10.1111/1365-2435.12347
Barrios-O’Neill, D., Kelly, R., Dick, J. T., Ricciardi, A., MacIsaac, H. J., & Emmerson, M. C. (2016). On the context-dependent scaling of
consumer feeding rates. Ecology Letters, 19(6), 668–678. https://doi.org/10.1111/ele.12605
Acknowledgments
This work was supported by National
Key Program of Science and Technol-
ogy (2022YFC3203900), and National
Natural Science Foundation of China
(No. 52121006, 92047303). Dr. Chen Q.
was also supported by the Xplorer Prize.
The authors are grateful to Dr. Honghai
Ma, Dr. Ting Li, Mr. Zhiyuan Zhang, Mr.
Yumeng Tang, Ms. Qi Zhang, Ms. Mo
Chen, Ms. Mengru Wei, Mr. Zaoli Yang,
and Mr. Peisi Yang for their contribution
in material collections. We also greatly
appreciate the valuable comments and
constructive suggestions from the anony-
mous reviewers.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
45 of 64
Barry, T., & Kynard, B. (1986). Attraction of adult American shad to fish lifts at Holyoke Dam, Connecticut River. North American Journal of
Fisheries Management, 6(2), 233–241. https://doi.org/10.1577/1548-8659(1986)6<233:AOAAST>2.0.CO;2
Barth, J. M., Gubili, C., Matschiner, M., Tørresen, O. K., Watanabe, S., Egger, B., etal. (2020). Stable species boundaries despite ten million years
of hybridization in tropical eels. Nature Communications, 11(1), 1–13. https://doi.org/10.1038/s41467-020-15099-x
Bartholow, J., Hanna, R. B., Saito, L., Lieberman, D., & Horn, M. (2001). Simulated limnological effects of the Shasta Lake temperature control
device. Environmental Management, 27(4), 609–626. https://doi.org/10.1007/s0026702324
Baumgartner, L. J., Boys, C. A., Marsden, T., McPherson, J., Ning, N., Phonekhampheng, O., etal. (2018). Comparing fishway designs for appli-
cation in a large tropical river system. Ecological Engineering, 120, 36–43. https://doi.org/10.1016/j.ecoleng.2018.05.027
Beck, C., Albayrak, I., Meister, J., Peter, A., Selz, O. M., Leuch, C., etal. (2020). Swimming behavior of downstream moving fish at innovative
curved-bar rack bypass systems for fish protection at water intakes. Water, 12(11), 3244. https://doi.org/10.3390/w12113244
Beeman, J. W., & Maule, A. G. (2006). Migration depths of juvenile Chinook salmon and steelhead relative to total dissolved gas supersaturation
in a Columbia River reservoir. Transactions of the American Fisheries Society, 135(3), 584–594. https://doi.org/10.1577/T05-193.1
Béguer-Pon, M., Castonguay, M., Shan, S., Benchetrit, J., & Dodson, J. J. (2015). Direct observations of American eels migrating across the
continental shelf to the Sargasso Sea. Nature Communications, 6(1), 1–9. https://doi.org/10.1038/ncomms9705
Behrmann-Godel, J., & Eckmann, R. (2003). A preliminary telemetry study of the migration of silver European eel (Anguilla anguilla L.) in the
River Mosel, Germany. Ecology of Freshwater Fish, 12(3), 196–202. https://doi.org/10.1034/j.1600-0633.2003.00015.x
Bell, E., Kramer, S., Zajanc, D., & Aspittle, J. (2008). Salmonid fry stranding mortality associated with daily water level fluctuations in Trail
Bridge Reservoir, Oregon. North American Journal of Fisheries Management, 28(5), 1515–1528. https://doi.org/10.1577/M07-026.1
Bell, J. D., Leber, K. M., Blankenship, H. L., Loneragan, N. R., & Masuda, R. (2008). A new era for restocking, stock enhancement and sea
ranching of coastal fisheries resources. Reviews in Fisheries Science, 16(1–3), 1–9. https://doi.org/10.1080/10641260701776951
Belletti, B., de Leaniz, C. G., Jones, J., Bizzi, S., Börger, L., Segura, G., etal. (2020). More than one million barriers fragment Europe’s rivers.
Nature, 588(7838), 436–441. https://doi.org/10.1038/s41586-020-3005-2
Benjankar, R., Tonina, D., McKean, J. A., Sohrabi, M. M., Chen, Q., & Vidergar, D. (2018). Dam operations may improve aquatic habitat and offset
negative effects of climate change. Journal of Environmental Management, 213, 126–134. https://doi.org/10.1016/j.jenvman.2018.02.066
Benjankar, R., Vidergar, D., Tonina, D., & Chen, Q. (2023). The role of water management and river morphology on stranding pool formation.
Ecological Engineering, 196, 107101. https://doi.org/10.1016/j.ecoleng.2023.107101
Bernery, C., Bellard, C., Courchamp, F., Brosse, S., Gozlan, R. E., Jarić, I., etal. (2022). Freshwater fish invasions: A comprehensive review.
Annual Review of Ecology, Evolution, and Systematics, 53(19), 1–30. https://doi.org/10.1146/annurev-ecolsys-032522-015551
Bertola, N., Wang, H., & Chanson, H. (2018). A physical study of air–water flow in planar plunging water jet with large inflow distance. Inter-
national Journal of Multiphase Flow, 100(3), 155–171. https://doi.org/10.1016/j.ijmultiphaseflow.2017.12.015
Besson, M. L., Trancart, T., Acou, A., Charrier, F., Mazel, V., Legault, A., & Feunteun, E. (2016). Disrupted downstream migration behaviour of
European silver eels (Anguilla anguilla L.) in an obstructed river. Environmental Biology of Fishes, 99(10), 779–791. https://doi.org/10.1007/
s10641-016-0522-9
Best, J. (2019). Anthropogenic stresses on the world’s big rivers. Nature Geoscience, 12(1), 7–21. https://doi.org/10.1038/s41561-018-0262-x
Best, J., Ashmore, P., & Darby, S. E. (2022). Beyond just floodwater. Nature Sustainability, 5(10), 811–813. https://doi.org/10.1038/
s41893-022-00929-1
Best, J., & Darby, S. E. (2020). The pace of human-induced change in large rivers: Stresses, resilience and vulnerability to extreme events. One
Earth, 2(6), 510–514. https://doi.org/10.1016/j.oneear.2020.05.021
Bevacqua, D., Melià, P., Gatto, M., & De Leo, G. A. (2015). A global viability assessment of the European eel. Global Change Biology, 21(9),
3323–3335. https://doi.org/10.1111/gcb.12972
Bhattarai, B., Hilliard, B., Reeder, W. J., Budwig, R., Martin, B. T., Xing, T., & Tonina, D. (2023). Effect of surface hydraulics and salmon redd
size on redd-induced hyporheic exchange. Water Resources Research, 59(6), 1–24. https://doi.org/10.1029/2022WR033977
Billard, R., & Lecointre, G. (2000). Biology and conservation of sturgeon and paddlefish. Reviews in Fish Biology and Fisheries, 10(4), 355–392.
https://doi.org/10.1023/A:1012231526151
Biswas, A. K., & Tortajada, C. (2012). Impacts of the high Aswan Dam. In Impacts of large Dams: A global assessment (pp.379–395). Springer.
Bittencourt, A. C. D. S. P., Dominguez, J. M. L., Fontes, L. C. S., Sousa, D. L., Silva, I. R., & Da Silva, F. R. (2007). Wave refraction, river
damming, and episodes of severe shoreline erosion: The São Francisco river mouth, Northeastern Brazil. Journal of Coastal Research, 23(4),
930–938. https://doi.org/10.2112/05-0600.1
Bohlin, T., Dellefors, C., & Faremo, U. (1993). Timing of sea-run brown trout (Salmo trutta) smolt migration: Effects of climatic variation.
Canadian Journal of Fisheries and Aquatic Sciences, 50(6), 1132–1136. https://doi.org/10.1139/F93-128
Boldyrev, V. (2018). Anadromous sturgeons (Acipenseridae, Actinopterygii) of the Don River upstream from the Tsimlyansk hydroengineering
complex. Biology Bulletin, 45(10), 1129–1138. https://doi.org/10.1134/S1062359018100084
Boscari, E., Wu, J., Jiang, T., Zhang, S., Cattelan, S., Wang, C., etal. (2022). The last giants of the Yangtze River: A multidisciplinary picture of
what remains of the endemic Chinese sturgeon. Science of the Total Environment, 843, 157011. https://doi.org/10.1016/j.scitotenv.2022.157011
Bouck, G. R. (1980). Etiology of gas bubble disease. Transactions of the American Fisheries Society, 109(6), 703–707. https://doi.
org/10.1577/1548-8659(1980)109<703:EOGBD>2.0.CO;2
Bouck, G. R. (1984). Annual variation of gas supersaturation in four spring-fed Oregon streams. The Progressive Fish-Culturist, 46(2), 139–140.
https://doi.org/10.1577/1548-8640(1984)46<139:AVOGSI>2.0.CO;2
Boulêtreau, S., & Santoul, F. (2016). The end of the mythical giant catfish. Ecosphere, 7(11), e01606. https://doi.org/10.1002/ecs2.1606
Bower, L. M., Peoples, B. K., Eddy, M. C., & Scott, M. C. (2022). Quantifying flow-ecology relationships across flow regime class and ecore-
gions in South Carolina. Science of the Total Environment, 802, 149721. https://doi.org/10.1016/j.scitotenv.2021.149721
Bracken, F. S. A., & Lucas, M. C. (2013). Potential impacts of small-scale hydroelectric power generation on downstream moving lampreys. River
Research and Applications, 29(9), 1073–1081. https://doi.org/10.1002/rra.2596
Bravo-Córdoba, F. J., Valbuena-Castro, J., García-Vega, A., Fuentes-Pérez, J. F., Ruiz-Legazpi, J., & Sanz-Ronda, F. J. (2021). Fish passage
assessment in stepped fishways: Passage success and transit time as standardized metrics. Ecological Engineering, 162, 106172. https://doi.
org/10.1016/j.ecoleng.2021.106172
Brevé, N. W., Leuven, R. S., Buijse, A. D., Murk, A. J., Venema, J., & Nagelkerke, L. A. (2022). The conservation paradox of critically endan-
gered fish species: Trading alien sturgeons versus native sturgeon reintroduction in the Rhine-Meuse River delta. Science of the Total Environ-
ment, 848, 157641. https://doi.org/10.1016/j.scitotenv.2022.157641
Brinker, A., Chucholl, C., Behrmann-Godel, J., Matzinger, M., Basen, T., & Baer, J. (2018). River damming drives population fragmentation and
habitat loss of the threatened Danube streber (Zingel streber): Implications for conservation. Aquatic Conservation: Marine and Freshwater
Ecosystems, 28(3), 587–599. https://doi.org/10.1002/aqc.2878
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
46 of 64
Brosnan, I. G., Welch, D. W., & Scott, M. J. (2016). Survival rates of out-migrating yearling chinook salmon in the lower Columbia River and
plume after exposure to gas-supersaturated water. Journal of Aquatic Animal Health, 28(4), 240–251. https://doi.org/10.1080/08997659.201
6.1227398
Bruijs, M., & Durif, C. M. (2009). Silver eel migration and behaviour. In Spawning migration of the European eel (pp.65–95). Springer.
Budy, P., Thiede, G. P., Bouwes, N., Petrosky, C. E., & Schaller, H. (2002). Evidence linking delayed mortality of Snake River salmon to their
earlier hydrosystem experience. North American Journal of Fisheries Management, 22(1), 35–51. https://doi.org/10.1577/1548-8675(2002)0
22<0035:ELDMOS>2.0.CO;2
Bunn, S. E., & Arthington, A. H. (2002). Basic principles and ecological consequences of altered flow regimes for aquatic biodiversity. Environ-
mental Management, 30(4), 492–507. https://doi.org/10.1007/s00267-002-2737-0
Bunt, C. M., Castro-Santos, T., & Haro, A. (2012). Performance of fish passage structures at upstream barriers to migration. River Research and
Applications, 28(4), 457–478. https://doi.org/10.1002/rra.1565
Bunt, C. M., Castro-Santos, T., & Haro, A. (2016). Reinforcement and validation of the analyses and conclusions related to fishway evaluation
data from Bunt etal.: ‘Performance of fish passage structures at upstream barriers to migration’. River Research and Applications, 32(10),
2125–2137. https://doi.org/10.1002/rra.3095
Bunt, C. M., van Poorten, B. T., & Wong, L. (2001). Denil fishway utilization patterns and passage of several warmwater species relative to
seasonal, thermal and hydraulic dynamics. Ecology of Freshwater Fish, 10(4), 212–219. https://doi.org/10.1034/j.1600-0633.2001.100403.x
Bunting, L., Leavitt, P. R., Gibson, C. E., McGee, E. J., & Hall, V. A. (2007). Degradation of water quality in Lough Neagh, Northern
Ireland, by diffuse nitrogen flux from a phosphorus-rich catchment. Limnology & Oceanography, 52(1), 354–369. https://doi.org/10.4319/
lo.2007.52.1.0354
Cai, H., Piccolroaz, S., Huang, J., Liu, Z., Liu, F., & Toffolon, M. (2018). Quantifying the impact of the three Gorges dam on the thermal dynam-
ics of the Yangtze River. Environmental Research Letters, 13(5), 054016. https://doi.org/10.1088/1748-9326/aab9e0
Caissie, D. (2006). The thermal regime of rivers: A review. Freshwater Biology, 51(8), 1389–1406. https://doi.org/10.1111/j.1365-2427.2006.01597.x
Calderon, M. S., & An, K. G. (2016). An influence of mesohabitat structures (pool, riff le, and run) and land-use pattern on the index of biological
integrity in the Geum River watershed. Journal of Ecology and Environment, 40(1), 1–13. https://doi.org/10.1186/s41610-016-0018-8
Cambray, J. A., King, J. M., & Bruwer, C. (1997). Spawning behaviour and early development of the Clanwilliam yellowfish (Barbus
capensis; Cyprinidae), linked to experimental dam releases in the Olifants River, South Africa. Regulated Rivers: Research &
Management: An International Journal Devoted to River Research and Management, 13(6), 579–602. https://doi.org/10.1002/
(SICI)1099-1646(199711/12)13:6<579::AID-RRR486>3.0.CO;2-F
Campton, D. E. (2004). Sperm competition in salmon hatcheries—The need to institutionalize genetically benign spawning protocols: Response
to comment. Transactions of the American Fisheries Society, 133(5), 1277–1289. https://doi.org/10.1577/T03-200.1
Canonico, G. C., Arthington, A., Mccrary, J. K., & Thieme, M. L. (2005). The effects of introduced tilapias on native biodiversity. Aquatic
Conservation: Marine and Freshwater Ecosystems, 15(5), 463–483. https://doi.org/10.1002/aqc.699
Cao, C., Deng, Y., Yin, Q., Li, N., Liu, X., Shi, H., etal. (2020). Effects of continuous acute and intermittent exposure on the tolerance of juvenile
yellow catfish (Pelteobagrus fulvidraco) in total dissolved gas supersaturated water. Ecotoxicology and Environmental Safety, 201, 110855.
https://doi.org/10.1016/j.ecoenv.2020.110855
Cao, L., Li, K., Liang, R., Chen, S., Jiang, W., & Li, R. (2016). The tolerance threshold of Chinese sucker to total dissolved gas supersaturation.
Aquaculture Research, 47(9), 2804–2813. https://doi.org/10.1111/are.12730
Cao, W. (2000). Some considerations on construction of national nature reserve for rare and endemic fishes in the upper Yangtze River Basin.
Resources and Environment in the Yangtze Basin, 9(2), 131–132.
Casimiro, A. C. R., Vizintim Marques, A. C., Claro-Garcia, A., Garcia, D. A. Z., de Almeida, F. S., & Orsi, M. L. (2022). Hatchery fish stock-
ing: Case study, current Brazilian state, and suggestions for improvement. Aquaculture International, 30(5), 1–18. https://doi.org/10.1007/
s10499-022-00898-4
Castello, L., & Macedo, M. N. (2016). Large-scale degradation of Amazonian freshwater ecosystems. Global Change Biology, 22(3), 990–1007.
https://doi.org/10.1111/gcb.13173
Castro, S. T., Shi, X., & Haro, A. (2016). Migratory behavior of adult sea lamprey and cumulative passage performance through four fishways.
Canadian Journal of Fisheries and Aquatic Sciences, 74(5), 790–800. https://doi.org/10.1139/cjfas-2016-0089
Catalano, M. J., Bozek, M. A., & Pellett, T. D. (2007). Effects of dam removal on fish assemblage structure and spatial distributions in the Bara-
boo River, Wisconsin. North American Journal of Fisheries Management, 27(2), 519–530. https://doi.org/10.1577/M06-001.1
Cattaneo, F., Guillard, J., Diouf, S., O'Rourke, J., & Grimardias, D. (2021). Mitigation of ecological impacts on fish of large reservoir sediment
management through controlled flushing - the case of the Verbois dam (Rhone River, Switzerland). Science of the Total Environment, 756,
144053. https://doi.org/10.1016/j.scitotenv.2020.144053
Caudill, C. C., Keefer, M. L., Clabough, T. S., Naughton, G. P., Burke, B. J., & Peery, C. A. (2013). Indirect effects of impoundment on migrating
fish: Temperature gradients in fish ladders slow dam passage by adult chinook salmon and steelhead. PLoS One, 8(12), e85586. https://doi.
org/10.1371/journal.pone.0085586
Černý, J., Copp, G. H., Kováč, V., Gozlan, R., & Vilizzi, L. (2003). Initial impact of the Gabčíkovo hydroelectric scheme on the species richness
and composition of 0+ fish assemblages in the Slovak flood plain, River Danube. River Research and Applications, 19(7), 749–766. https://
doi.org/10.1002/rra.716
Chai, Y., Huang, J., Zhu, T., Yang, D., & Wu, X. (2019). Preliminary study on substrate preference of Schizothorax wangchiachii larvae fish.
Freshwater Fisheries, 49(1), 44–47.
Chang, T., Gao, X., & Liu, H. (2021). Potential hydrological regime requirements for spawning success of the Chinese sturgeon Acipenser sinen-
sis in its present spawning ground of the Yangtze River. Ecohydrology, 14(8), e2339. https://doi.org/10.1002/eco.2339
Chang, T., Lin, P., Gao, X., Liu, F., Duan, Z., & Liu, H. (2017). Using adaptive resolution imaging sonar to investigate Chinese sturgeon (Acip-
enser sinensis Gray, 1835) behaviour on its only spawning ground in the Yangtze River. Journal of Applied Ichthyology, 33(4), 681–688.
https://doi.org/10.1111/jai.13406
Chapman, P.M., & Wang, F. (2001). Assessing sediment contamination in estuaries. Environmental Toxicology and Chemistry, 20(1), 3–22.
https://doi.org/10.1002/etc.5620200102
Chapuis, M., Dufour, S., Provansal, M., Couvert, B., & de Linares, M. (2015). Coupling channel evolution monitoring and RFID tracking in a
large, wandering, gravel-bed river: Insights into sediment routing on geomorphic continuity through a riffle–pool sequence. Geomorphology,
231, 258–269. https://doi.org/10.1016/j.geomorph.2014.12.013
Chen, D., Chen, Q., Leon, A. S., & Li, R. (2016). A genetic algorithm parallel strategy for optimizing the operation of reservoir with multiple
eco-environmental objectives. Water Resources Management, 30(7), 2127–2142. https://doi.org/10.1007/s11269-016-1274-1
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
47 of 64
Chen, D., Chen, Q., Li, R., Blanckaert, K., & Cai, D. (2014). Ecologically-friendly operation scheme for the Jinping cascaded reservoirs in the
Yalongjiang River, China. Frontiers of Earth Science, 8(2), 282–290. https://doi.org/10.1007/s11707-013-0396-5
Chen, D., Leon, A. S., Engle, S. P., Fuentes, C., & Chen, Q. (2017). Offline training for improving online performance of a genetic algorithm
based optimization model for hourly multi-reservoir operation. Environmental Modelling & Software, 96, 46–57. https://doi.org/10.1016/j.
envsoft.2017.06.038
Chen, D., Leon, A. S., Gibson, N. L., & Hosseini, P. (2016). Dimension reduction of decision variables for multireservoir operation: A spectral
optimization model. Water Resources Research, 52(1), 36–51. https://doi.org/10.1002/2015WR017756
Chen, D., Leon, A. S., Hosseini, P., Gibson, N. L., & Fuentes, C. (2017). Application of cluster analysis for finding operational patterns of
multireservoir system during transition period. Journal of Water Resources Planning and Management, 143(8), 04017028. https://doi.
org/10.1061/(ASCE)WR.1943-5452.0000772
Chen, D., Li, R., Chen, Q., & Cai, D. (2015). Deriving optimal daily reservoir operation scheme with consideration of downstream ecological
hydrograph through a time-nested approach. Water Resources Management, 29(9), 3371–3386. https://doi.org/10.1007/s11269-015-1005-z
Chen, D., Xiong, F., Wang, K., & Chang, Y. (2009). Status of research on Yangtze fish biology and fisheries. Environmental Biology of Fishes,
85(4), 337–357. https://doi.org/10.1007/s10641-009-9517-0
Chen, H. J. (2019). Effects of broodstock enhancement on fish larval resources and genetic diversity of the four major Chinese carps in the middle
reaches of the Yangtze River. Southwest University.
Chen, J. Z., Huang, S. L., & Han, Y. S. (2014). Impact of long-term habitat loss on the Japanese eel Anguilla japonica. Estuarine, Coastal and
Shelf Science, 151, 361–369. https://doi.org/10.1016/j.ecss.2014.06.004
Chen, Q., Chen, D., Han, R., Li, R., Ma, J., & Blanckaert, K. (2012). Optimizing the operation of the Qingshitan Reservoir in the Lijiang River for
multiple human interests and quasi-natural flow maintenance. Journal of Environmental Sciences, 24(11), 1923–1928. https://doi.org/10.1016/
S1001-0742(11)61029-2
Chen, Q., Chen, D., Li, R., Ma, J., & Blanckaert, K. (2013). Adapting the operation of two cascaded reservoirs for ecological flow requirement
of a de-watered river channel due to diversion-type hydropower stations. Ecological Modelling, 252(SI), 266–272. https://doi.org/10.1016/j.
ecolmodel.2012.03.008
Chen, Q., Zhang, J., Chen, Y., Mo, K., Wang, J., Tang, L., etal. (2021). Inducing flow velocities to manage fish reproduction in regulated rivers.
Engineering, 7(2), 178–186. https://doi.org/10.1016/j.eng.2020.06.013
Chen, W., & Olden, J. D. (2017). Designing flows to resolve human and environmental water needs in a dam-regulated river. Nature Communi-
cations, 8(1), 2158. https://doi.org/10.1038/s41467-017-02226-4
Chen, Y., Li, W., Deng, H., Fang, G., & Li, Z. (2016). Changes in central Asia’s water tower: Past, present and future. Scientific Reports, 6(1),
1–12. https://doi.org/10.1038/srep35458
Cheng, F., Li, W., Castello, L., Murphy, B. R., & Xie, S. (2015). Potential effects of dam cascade on fish: Lessons from the Yangtze River. Reviews
in Fish Biology and Fisheries, 25(3), 569–585. https://doi.org/10.1007/s11160-015-9395-9
Chezik, K. A., Lester, N. P., & Venturelli, P.A. (2014). Fish growth and degree-days I: Selecting a base temperature for a within-population study.
Canadian Journal of Fisheries and Aquatic Sciences, 71(1), 47–55. https://doi.org/10.1139/cjfas-2013-0295
Chong, X. Y., Vericat, D., Batalla, R. J., Teo, F. Y., Lee, K. S. P., & Gibbins, C. N. (2021). A review of the impacts of dams on the hydromorphol-
ogy of tropical rivers. Science of the Total Environment, 794, 148686. https://doi.org/10.1016/j.scitotenv.2021.148686
Chu, C. R., & Jirka, G. H. (2003). Wind and stream flow induced reaeration. Journal of Environmental Engineering, 129(12), 1129–1136. https://
doi.org/10.1061/(ASCE)0733-9372(2003)129:12(1129)
Cioffi, F., & Gallerano, F. (2012). Multi-objective analysis of dam release flows in rivers downstream from hydropower reservoirs. Applied
Mathematical Modelling, 36(7), 2868–2887. https://doi.org/10.1016/j.apm.2011.09.077
Clay, C. H. (1995). Design of fishways and other fish facilities (2nd ed.,p.248). Lewis Publishers.
Cohen-Shacham, E., Walters, G., Janzen, C., & Maginnis, S. (2016). Nature-based Solutions to address global societal challenges (Vol. 13,p.97).
IUCN. https://doi.org/10.2305/IUCN.CH.2016.13.en
Collingsworth, P.D., Bunnell, D. B., Murray, M. W., Kao, Y. C., Feiner, Z. S., Claramunt, R. M., etal. (2017). Climate change as a long-term
stressor for the fisheries of the Laurentian Great lakes of North America. Reviews in Fish Biology and Fisheries, 27(2), 363–391. https://doi.
org/10.1007/s11160-017-9480-3
Collins, D. N. (2009). Seasonal variations of water temperature and discharge in rivers draining ice-free and partially-glacierised Alpine basins.
In Northern research basins symposium (Vol. 17,pp.67–74). Retrieved from https://usir.salford.ac.uk/id/eprint/18324
Connor, W. P., Burge, H. L., Waitt, R., & Bjornn, T. C. (2002). Juvenile life history of wild fall Chinook salmon in the Snake and Clearwater rivers.
North American Journal of Fisheries Management, 22(3), 703–712. https://doi.org/10.1577/1548-8675(2002)022<0703:JLHOWF>2.0.CO;2
Connor, W. P., Burge, H. L., Yearsley, J. R., & Bjornn, T. C. (2003). Influence of flow and temperature on survival of wild subyear-
ling fall Chinook salmon in the Snake River. North American Journal of Fisheries Management, 23(2), 362–375. https://doi.
org/10.1577/1548-8675(2003)023<0362:IOFATO>2.0.CO;2
Consuegra, S., O'Rorke, R., Rodríguez Barreto, D., Fernández, S., Jones, J., & Garcia de Leaniz, C. (2021). Impacts of large and small barriers
on fish assemblage composition assessed using environmental DNA metabarcoding. Science of the Total Environment, 790, 148054. https://
doi.org/10.1016/j.scitotenv.2021.148054
Cooper, A., Infante, D., Daniel, W., Wehrly, K., Wang, L., & Brenden, T. (2017). Assessment of dam effects on streams and fish assemblages of
the conterminous USA. Science of the Total Environment, 586, 879–889. https://doi.org/10.1016/j.scitotenv.2017.02.067
Cott, P.A., Sibley, P.K., Somers, W. M., Lilly, M. R., & Gordon, A. M. (2008). A review of water level fluctuations on aquatic biota
with an emphasis on fishes in ice-covered lakes. Journal of the American Water Resources Association, 44(2), 343–359. https://doi.
org/10.1111/j.1752-1688.2007.00166.x
Coulter, D. P., MacNamara, R., Glover, D. C., & Garvey, J. E. (2018). Possible unintended effects of management at an invasion front: Reduced
prevalence corresponds with high condition of invasive bigheaded carps. Biological Conservation, 221, 118–126. https://doi.org/10.1016/j.
biocon.2018.02.020
Counihan, T. D., & Chapman, C. G. (2018). Relating river discharge and water temperature to the recruitment of age-0 White Sturgeon (Acipenser
transmontanus Richardson, 1836) in the Columbia River using over-dispersed catch data. Journal of Applied Ichthyology, 34(2), 279–289.
https://doi.org/10.1111/jai.13570
Couto, T., Messager, M. L., & Olden, J. D. (2021). Safeguarding migratory fish via strategic planning of future small hydropower in Brazil.
Nature Sustainability, 4(5), 409–416. https://doi.org/10.1038/s41893-020-00665-4
Couto, T., & Olden, J. (2018). Global proliferation of small hydropower plants - Science and policy. Frontiers in Ecology and the Environment,
16(2), 91–100. https://doi.org/10.1002/fee.1746
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
48 of 64
Croze, O., Bau, F., & Delmouly, L. (2008). Efficiency of a fish lift for returning Atlantic salmon at a large-scale hydroelectric complex in France.
Fisheries Management and Ecology, 15(5–6), 467–476. https://doi.org/10.1111/j.1365-2400.2008.00628.x
Crozier, L. G., Burke, B. J., Chasco, B. E., Widener, D. L., & Zabel, R. W. (2021). Climate change threatens Chinook salmon throughout their life
cycle. Communications Biology, 4(1), 1–14. https://doi.org/10.1038/s42003-021-01734-w
Csiki, S., & Rhoads, B. L. (2010). Hydraulic and geomorphological effects of run-of-river dams. Progress in Physical Geography, 34(6),
755–780. https://doi.org/10.1177/0309133310369435
Cucherousset, J., & Olden, J. D. (2011). Ecological impacts of nonnative freshwater fishes. Fisheries, 36(5), 215–230. https://doi.org/10.1080/0
3632415.2011.574578
Cupp, A. R., Brey, M. K., Calfee, R. D., Chapman, D. C., Erickson, R., Fischer, J., etal. (2021). Emerging control strategies for integrated pest
management of invasive carps. Journal of Vertebrate Biology, 70(4), 21057. https://doi.org/10.25225/jvb.21057
Cyrus, D. P., & Blaber, S. J. M. (1987). The influence of turbidity on juvenile marine fishes in estuaries. Part 2. Laboratory stud-
ies, comparisons with field data and conclusions. Journal of Experimental Marine Biology and Ecology, 109(1), 71–91. https://doi.
org/10.1016/0022-0981(87)90186-9
Cyrus, D. P., & Blaber, S. J. M. (1992). Turbidity and salinity in a tropical northern Australian estuary and their influence on fish distribution.
Estuarine, Coastal and Shelf Science, 35(6), 545–563. https://doi.org/10.1016/S0272-7714(05)80038-1
Dai, L., Dai, H., Li, W., Tang, Z., Ren, Y., & Chang, M. (2022). Optimal operation of cascade hydropower plants considering spawning of four
major Chinese carps. Journal of Hydroelectric Engineering, 41(5), 21–30. https://doi.org/10.11660/slfdxb.20220503
Dalcin, A. P., Marques, G. F., de Oliveira, A. G., & Tilmant, A. (2022). Identifying functional flow regimes and fish response for multiple reservoir
operating solutions. Journal of Water Resources Planning and Management, 148(6). https://doi.org/10.1061/(ASCE)WR.1943-5452.0001567
Dang, T. H., Coynel, A., Orange, D., Blanc, G., Etcheber, H., & Le, L. A. (2010). Long-term monitoring (1960–2008) of the river-sediment
transport in the Red River Watershed (Vietnam): Temporal variability and dam-reservoir impact. Science of the Total Environment, 408(20),
4654–4664. https://doi.org/10.1016/j.scitotenv.2010.07.007
Da Silva, P.S., Makrakis, M. C., Miranda, L. E., Makrakis, S., Assumpção, L., Paula, S., et al. (2015). Importance of reservoir tributaries to
spawning of migratory fish in the upper Paraná River. River Research and Applications, 31(3), 313–322. https://doi.org/10.1002/rra.2755
Da Silva, P.S., Miranda, L. E., Makrakis, S., de Assumpção, L., Dias, J. H. P., & Makrakis, M. C. (2019). Tributaries as biodiversity preserves:
An ichthyoplankton perspective from the severely impounded Upper Paraná River. Aquatic Conservation: Marine and Freshwater Ecosystems,
29(2), 258–269. https://doi.org/10.1002/aqc.3037
David, G., Céline, C., Morgane, B., & Franck, C. (2022). Ecological connectivity of the upper Rhône River: Upstream fish passage at two successive
large hydroelectric dams for partially migratory species. Ecological Engineering, 178, 106545. https://doi.org/10.1016/j.ecoleng.2022.106545
Day, J. W., Boesch, D. F., Clairain, E. J., Kemp, G. P., Laska, S. B., Mitsch, W. J., etal. (2007). Restoration of the Mississippi delta: Lessons from
hurricanes Katrina and Rita. Science, 315(5819), 1679–1684. https://doi.org/10.1126/science.1137030
Deinet, S., Scott-Gatty, K., Rotton, H., Twardek, W. M., Marconi, V., McRae, L., etal. (2020). The living planet index (LPI) for migratory fresh-
water fish. Technical report (p.30). World Fish Migration Foundation.
Deng, Y., Cao, C., Liu, X., Yuan, Q., Feng, C., Shi, H., etal. (2020). Effect of total dissolved gas supersaturation on the survival of bighead carp
(Hypophthalmichthys bilis). Animals, 10(1), 1–14. https://doi.org/10.3390/ani10010166
Denil, G. (1909). Les échelles à poissons et leur application aux barrage de Meuse et d'Ourthe (p.152). Annales des travaux publics de Belgique.
Denil, G. (1938). La mécanique du poisson de rivière (p.395). Annales des travaux publics de Belgique.
Dettinger, M., & Diaz, H. (2000). Global characteristics of stream flow seasonality and variability. Journal of Hydrometeorology, 1(4), 289–310.
https://doi.org/10.1175/1525-7541(2000)001<0289:GCOSFS>2.0.CO;2
Díaz, G., Górski, K., Heino, J., Arriagada, P., Link, O., & Habit, E. (2020). The longest fragment drives fish beta diversity in fragmented river
networks: Implications for river management and conservation. Science of the Total Environment, 766, 144323. https://doi.org/10.1016/j.
scitotenv.2020.144323
Dietrich, W. E., Kirchner, J. W., Ikeda, H., & Iseya, F. (1989). Sediment supply and the development of the coarse surface layer in gravel-bedded
rivers. Nature, 340(6230), 215–217. https://doi.org/10.1038/340215a0
Ding, C., Jiang, X., Wang, L., Fan, H., Chen, L., Hu, J., etal. (2019). Fish assemblage responses to a low-head dam removal in the Lancang River.
Chinese Geographical Science, 29(1), 26–36. https://doi.org/10.1007/s11769-018-0995-x
Dorozynski, A. (1975). After the dam, the depression? Nature, 255(5510), 570. https://doi.org/10.1038/255570a0
Dorts, J., Grenouillet, G., Douxfils, J., Mandiki, S. N., Milla, S., Silvestre, F., & Kestemont, P. (2012). Evidence that elevated water temperature
affects the reproductive physiology of the European bullhead Cottus gobio. Fish Physiology and Biochemistry, 38(2), 389–399. https://doi.
org/10.1007/s10695-011-9515-y
Du, K., Stöck, M., Kneitz, S., Klopp, C., Woltering, J. M., Adolfi, M. C., etal. (2020). The sterlet sturgeon genome sequence and the mechanisms
of segmental rediploidization. Nature ecology & evolution, 4(6), 841–852. https://doi.org/10.1038/s41559-020-1166-x
Duan, X., Liu, S., Huang, M., Qiu, S., Li, Z., Wang, K., & Chen, D. (2009). Changes in abundance of larvae of the four domestic Chinese carps
in the middle reach of the Yangtze River, China, before and after closing of the Three Gorges Dam. Environmental Biology of Fishes, 86(1),
13–22. https://doi.org/10.1007/s10641-009-9498-z
Duarte, G., Segurado, P., Haidvogl, G., Pont, D., & Branco, P. (2021). Damn those damn dams: Fluvial longitudinal connectivity impair-
ment for European diadromous fish throughout the 20th century. Science of the Total Environment, 761, 143293. https://doi.org/10.1016/j.
scitotenv.2020.143293
Dugdale, S. J., Hannah, D. M., & Malcolm, I. A. (2017). River temperature modelling: A review of process-based approaches and future direc-
tions. Earth-Science Reviews, 175, 97–133. https://doi.org/10.1016/j.earscirev.2017.10.009
Duponchelle, F., Isaac, V. J., Doria, C., Van Damme, P.A., Herrera-R, G. A., Anderson, E. P., etal. (2021). Conservation of migratory fishes in
the Amazon basin. In Aquatic Conservation: Marine and Freshwater Ecosystems, aqc (p.3550). https://doi.org/10.1002/aqc.3550
Duponchelle, F., Pouilly, M., Pécheyran, C., Hauser, M., Renno, J. F., Panfili, J., etal. (2016). Trans-Amazonian natal homing in giant catfish.
Journal of Applied Ecology, 53(5), 1511–1520. https://doi.org/10.1111/1365-2664.12665
Dursun, B., & Gokcol, C. (2011). The role of hydroelectric power and contribution of small hydropower plants for sustainable development in
Turkey. Renewable Energy, 36(4), 1227–1235. https://doi.org/10.1016/j.renene.2010.10.001
Dusek Jennings, E., & Hendrix, A. N. (2020). Spawn timing of winter-run chinook salmon in the upper Sacramento River. San Francisco Estuary
and Watershed Science, 18(2). https://doi.org/10.15447///sfews.2020v18iss2art5
Elhetawy, A. I., Vasilyeva, L. M., Lotfy, A. M., Emelianova, N., Abdel-Rahim, M. M., Helal, A. M., etal. (2020). Effects of the rearing system of
the Russian sturgeon (Acipenser gueldenstaedtii) on growth, maturity, and the quality of produced caviar. Aquaculture, Aquarium, Conserva-
tion & Legislation, 13(6), 3798–3809. Retrieved from http://www.bioflux.com.ro/aacl
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
49 of 64
Elliott, J. M., & Elliott, J. A. (2010). Temperature requirements of atlantic salmon Salmo salar, brown trout Salmo trutta and arctic charr Salvelinus
alpinus: Predicting the effects of climate change. Journal of Fish Biology, 77(8), 1793–1817. https://doi.org/10.1111/j.1095-8649.2010.02762.x
Embke, H. S., Kocovsky, P.M., Richter, C. A., Pritt, J. J., Mayer, C. M., & Qian, S. S. (2016). First direct confirmation of grass carp spawning in
a Great Lakes tributary. Journal of Great Lakes Research, 42(4), 899–903. https://doi.org/10.1016/j.jglr.2016.05.002
EPA. (1987). Quality criteria for water 1986.
Espa, P., Crosa, G., Gentili, G., Quadroni, S., & Petts, G. (2015). Downstream ecological impacts of controlled sediment flushing in an alpine
valley river: A case study. River Research and Applications, 31(8), 931–942. https://doi.org/10.1002/rra.2788
Fan, J., & Morris, G. L. (1998). Reservoir sedimentation handbook: Design and management of dams, reservoirs, and watersheds for sustainable
use. McGraw-Hill.
FAO. (2015). Fishery and aquaculture Statistics (Global aquaculture production 1950–2013) (FishStatJ). In FAO fisheries and aquaculture
department (online), Rome. Retrieved from http://www.fao.org./fishery/en/statistics/software/FishStatJ/en
FAO. (2019). The state of world fisheries and aquaculture 2018. In Meeting the sustainable development goals, Rome.
FAO. (2022). The state of world fisheries and aquaculture 2022. In Towards blue transformation. Rome.
Farmer, G., & Beamish, F. (1969). Oxygen consumption of Tilapia nilotica in relation to swimming speed and salinity. Journal of the Fisheries
Research Board of Canada, 26(11), 2807–2821. https://doi.org/10.1139/f69-277
Farmer, T. M., Marschall, E. A., Dabrowski, K., & Ludsin, S. A. (2015). Short winters threaten temperate fish populations. Nature Communica-
tions, 6(1), 1–10. https://doi.org/10.1038/ncomms8724
Faulkner, J. R., Bellerud, B. L., Widener, D. L., & Zabel, R. W. (2019). Associations among fish length, dam passage history, and survival to
adulthood in two at-risk species of Pacific salmon. Transactions of the American Fisheries Society, 148(6), 1069–1087. https://doi.org/10.1002/
tafs.10200
Faunce, C., & Paperno, R. (1999). Tilapia-Dominated fish assemblages within an impounded mangrove ecosystem in east-central Florida.
Wetlands, 19(1), 126–138. https://doi.org/10.1007/BF03161741
Fearnside, P. (2013). Impacts of Brazil's Madeira River dams: Unlearned lessons for hydroelectric development in Amazonia. Environmental
Science & Policy, 38, 164–172. https://doi.org/10.1016/j.envsci.2013.11.004
Fencl, J. S., Mather, M. E., Costigan, K. H., & Daniels, M. D. (2015). How big of an effect do small dams have? Using geomorphological foot-
prints to quantify spatial impact of low-head dams and identify patterns of across-dam variation. PLoS One, 10(11), e0141210. https://doi.
org/10.1371/journal.pone.0141210
Feng, C., Li, N., Wang, Y., Liu, X., Shi, X., Fu, C., etal. (2019). Effects of total dissolved gas supersaturated water at varying suspended
sediment concentrations on the survival of rock carp Procypris rabaudi. Fisheries Science, 85(6), 1067–1075. https://doi.org/10.1007/
s12562-019-01344-w
Feng, J., Li, R., & Li, K. (2010). Study on the supersaturated TDG release process downstream of high dam. Journal of Hydraulic Power, 29(1),
7–12.
Feng, J., Li, R., Liang, R., & Shen, X. (2014). Eco-environmentally friendly operational regulation: An effective strategy to diminish the TDG
supersaturation of reservoirs. Hydrology and Earth System Sciences, 18(3), 1213–1223. https://doi.org/10.5194/hess-18-1213-2014
Feng, J., Li, R., Tang, C., & Yong, X. (2012). Experimental study on the sediment effect on releasing process of supersaturated total dissolved
gas. Advances in Water Science, 23(5), 702–708.
Feng, J., Wang, L., Li, R., Li, K., Pu, X., & Li, Y. (2018). Operational regulation of a hydropower cascade based on the mitigation of the total
dissolved gas supersaturation. Ecological Indicators, 92(3), 124–132. https://doi.org/10.1016/j.ecolind.2017.04.015
Ficke, A. D., Myrick, C. A., & Hansen, L. J. (2007). Potential impacts of global climate change on freshwater fisheries. Reviews in Fish Biology
and Fisheries, 17(4), 581–613. https://doi.org/10.1007/s11160-007-9059-5
Figueiredo, J. S. M. C. D., Fantin-Cruz, I., Silva, G. M. S., Beregula, R. L., Girard, P., Zeilhofer, P., etal. (2021). Hydropeaking by small hydro-
power facilities affects flow regimes on tributaries to the Pantanal Wetland of Brazil. Frontiers in Environmental Science, 9, 1–13. https://doi.
org/10.3389/fenvs.2021.577286
Firehammer, J. A., & Scarnecchia, D. L. (2006). Spring migratory movements by paddlefish in natural and regulated river segments of the
Missouri and Yellowstone rivers, North Dakota and Montana. Transactions of the American Fisheries Society, 135(1), 200–217. https://doi.
org/10.1577/T05-058.1
Fiseha, B. M., Setegn, S., Melesse, A., Volpi, E., & Fiori, A. (2014). Impact of climate change on the hydrology of upper Tiber River Basin using
bias corrected regional climate model. Water Resources Management, 28(5), 1327–1343. https://doi.org/10.1007/s11269-014-0546-x
Flecker, A. S., Shi, Q., Almeida, R. M., Angarita, H., Gomes-Selman, J. M., García-Villacorta, R., etal. (2022). Reducing adverse impacts of
Amazon hydropower expansion. Science, 375(6582), 753–760. https://doi.org/10.1126/science.abj4017
Foley, M. M., Bellmore, J., O'Connor, J. E., Duda, J. J., East, A. E., Grant, G., etal. (2017). Dam removal: Listening in. Water Resources Research,
53(7), 5229–5246. https://doi.org/10.1002/2017WR020457
Foley, M. M., Warrick, J. A., Ritchie, A., Stevens, A. W., Shafroth, P.B., Duda, J. J., etal. (2017). Coastal habitat and biological community
response to dam removal on the Elwha River. Ecological Monographs, 87(4), 552–577. https://doi.org/10.1002/ecm.1268
Förstner, U., Heise, S., Ahlf, W., & Westrich, B. (2008). Data quality assurance of sediment monitoring. In The water framework directive—
Ecological and chemical status monitoring (pp.371–386). https://doi.org/10.1002/9780470716090.ch22
Förstner, U., Heise, S., Schwartz, R., Westrich, B., & Ahlf, W. (2004). Historical contaminated sediments and soils at the river basin scale. Journal
of Soils and Sediments, 4(4), 247–260. https://doi.org/10.1007/BF02991121
Fox, C. A., Reo, N. J., Fessell, B., & Dituri, F. (2022). Native American tribes and dam removal: Restoring the Ottaway, Penobscot and Elwha rivers.
Water Alternatives, 15(1), 31–55. Retrieved from https://www.water-alternatives.org/index.php/alldoc/articles/vol15/v15issue1/652-a15-1-3/file
Fraser, B. (2018). Dams nudge Amazon's ecosystems off-kilter. Science, 359(6375), 508–509. https://doi.org/10.1126/science.359.6375.50
Freedman, J., Lorson, B., Taylor, R., Carline, R., & Stauffer, J. (2014). River of the Dammed: Longitudinal changes in fish assemblages in
response to dams. Hydrobiologia, 727(1), 19–33. https://doi.org/10.1007/s10750-013-1780-6
Freeman, M. C., Pringle, C. M., & Jackson, C. R. (2007). Hydrologic connectivity and the contribution of stream headwaters to ecological integ-
rity at regional scales. Journal of the American Water Resources Association, 43(1), 5–14. https://doi.org/10.1111/j.1752-1688.2007.00002.x
Fritts, A. K., Knights, B. C., Stanton, J. C., Milde, A. S., Vallazza, J. M., Brey, M. K., et al. (2021). Lock operations influence upstream
passages of invasive and native fishes at a Mississippi River high-head dam. Biological Invasions, 23(3), 771–794. https://doi.org/10.1007/
s10530-020-02401-7
Fuchs, N. T., Caudill, C. C., Murdoch, A. R., & Truscott, B. L. (2021). Overwintering distribution and post spawn survival of steelhead in the
upper Columbia River basin. North American Journal of Fisheries Management, 41(3), 757–774. https://doi.org/10.1002/nafm.10585
Fuller, M. R., Doyle, M. W., & Strayer, D. L. (2016). Causes and consequences of habitat fragmentation in river networks. Annals of the New York
Academy of Sciences, 1355(1), 31–51. https://doi.org/10.1111/nyas.12853
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
50 of 64
Fullerton, A., Kelly, B., Steel, E., Flitcroft, R., Pess, G., Feist, B., etal. (2010). Hydrological connectivity for riverine fish: Measurement chal-
lenges and research opportunities. Freshwater Biology, 55(11), 2215–2237. https://doi.org/10.1111/j.1365-2427.2010.02448.x
Gao, X., Brosse, S., Chen, Y., Lek, S., & Chang, J. (2009). Effects of damming on population sustainability of Chinese sturgeon, Acipenser sinensis:
Evaluation of optimal conservation measures. Environmental Biology of Fishes, 86(2), 325–336. https://doi.org/10.1007/s10641-009-9521-4
Gao, Y., Liu, Y., Wang, H., & Zhang, Z. (2018). Status of decommissioned dam removal and review of research progress on its effects. Journal
of Water Resources and Water Engineering, 29, 133–139.
Garcia, D. A. Z., Vidotto-Magnoni, A. P., Costa, A. D. A., Casimiro, A. C. R., Jarduli, L. R., Ferraz, J. D., etal. (2019). Importance of the
Congonhas River for the conservation of the fish fauna of the upper Paraná basin, Brazil. Biodiversitas Journal of Biological Diversity, 20(2),
474–481. https://doi.org/10.13057/biodiv/d200225
Garcia, T., Jackson, P.R., Murphy, E. A., Valocchi, A. J., & Garcia, M. H. (2013). Development of a Fluvial Egg Drift Simulator to evaluate the
transport and dispersion of Asian carp eggs in rivers. Ecological Modelling, 263, 211–222. https://doi.org/10.1016/j.ecolmodel.2013.05.005
Garcia de Leaniz, C. (2008). Weir removal in salmonid streams: Implications, challenges and practicalities. Hydrobiologia, 609(1), 83–96. https://
doi.org/10.1007/s10750-008-9397-x
Garde, R. J., & Raju, K. R. (1979). Mechanics of sediment transportation and alluvial stream problems. Halsted Press. (Wiley Eastern Limited,
New Delhi), 1977. https://doi.org/10.1002/esp.3290040119
Geist, D. R., Linley, T. J., Cullinan, V., & Deng, Z. (2013). The effects of total dissolved gas on Chum Salmon fry survival, growth, gas bubble
disease, and seawater tolerance. North American Journal of Fisheries Management, 33(1), 200–215. https://doi.org/10.1080/02755947.2012
.750634
General Authority for Fish Resources Development. (2012). Fisheries statistics yearbook (1996–2008).
George, A. E., Garcia, T., & Chapman, D. C. (2017). Comparison of size, terminal fall velocity, and density of bighead carp, silver carp, and
grass carp eggs for use in drift modeling. Transactions of the American Fisheries Society, 146(5), 834–843. https://doi.org/10.1080/0002848
7.2017.1310136
Glowa, S. E., Watkinson, D. A., Jardine, T. D., & Enders, E. C. (2022). Evaluating the risk of fish stranding due to hydropeaking in a large conti-
nental river. River Research and Applications, 39(3), 1–16. https://doi.org/10.1002/rra.4083
Goodwell, A. E., & Bassiouni, M. (2022). Source relationships and model structures determine information flow paths in ecohydrologic models.
Water Resources Research, 58(9), e2021WR031164. https://doi.org/10.1029/2021WR031164
Goodwin, R. A., Politano, M., Garvin, J. W., Nestler, J. M., Hay, D., Anderson, J. J., etal. (2014). Fish navigation of large dams emerges from their
modulation of flow field experience. Proceedings of the National Academy of Sciences of the United States of America, 111(14), 5277–5282.
https://doi.org/10.1073/pnas.131187411
Gorman, O. T., & Stone, D. M. (1999). Ecology of spawning humpback chub, Gila cypha, in the little Colorado river near Grand Canyon,
Arizona. Environmental Biology of Fishes, 55(1), 115–133. https://doi.org/10.1023/A:1007450826743
Graf, W. L. (2006). Downstream hydrologic and geomorphic effects of large dams on American rivers. Geomorphology, 79(3–4), 336–360.
https://doi.org/10.1016/j.geomorph.2006.06.022
Grant, E., Lowe, W., & Fagan, W. (2007). Living in the branches: Population dynamics and ecological processes in dendritic networks. Ecology
Letters, 10(2), 165–175. https://doi.org/10.1111/j.1461-0248.2006.01007.x
Grill, G., Lehner, B., Lumsdon, A., MacDonald, G., Zarfl, C., & Liermann, C. (2015). An index-based framework for assessing patterns and
trends in river fragmentation and flow regulation by global dams at multiple scales. Environmental Research Letters, 10(1), 015001. https://
doi.org/10.1088/1748-9326/10/1/015001
Grill, G., Lehner, B., Thieme, M., Geenen, B., Tickner, D., Antonelli, F., etal. (2019). Mapping the world’s free-flowing rivers. Nature, 569(7755),
215–221. https://doi.org/10.1038/s41586-019-1111-9
Grimardias, D., Guillard, J., & Cattanéo, F. (2017). Drawdown flushing of a hydroelectric reservoir on the Rhône River: Impacts on the fish
community and implications for the sediment management. Journal of Environmental Management, 197, 239–249. https://doi.org/10.1016/j.
jenvman.2017.03.096
Griscom, B. W., Adamsa, J., Ellisa, P.W., Houghtonc, R. A., Lomaxa, G., Mitevad, D. A., etal. (2017). Natural climate solutions. Proceedings of
the National Academy of Sciences of the United States of America, 114(44), 11645–11650. https://doi.org/10.1073/pnas.1710465114
Groot, G. (1991). Pacific salmon life histories. UBC Press.
Guo, C., Jin, Z., Guo, L., Lu, J., Ren, S., & Zhou, Y. (2020). On the cumulative dam impact in the upper Changjiang River: Streamflow and
sediment load changes. Catena, 184, 104250. https://doi.org/10.1016/j.catena.2019.104250
Guo, D., Zhou, L., Wang, G., Lai, H., Bi, S., Chen, X., etal. (2020). Use of artificial structures to enhance fish diversity in the Youjiang River, a
dammed river of the Pearl River in China. Ecology and Evolution, 10(23), 13439–13450. https://doi.org/10.1002/ece3.6949
Guy, C. S., Treanor, H. B., Kappenman, K. M., Scholl, E. A., Ilgen, J. E., & Webb, M. A. (2015). Broadening the regulated-river management
paradigm: A case study of the forgotten dead zone hindering pallid sturgeon recovery. Fisheries, 40(1), 6–14. https://doi.org/10.1080/03632
415.2014.987236
Habel, M., Mechkin, K., Podgorska, K., Saunes, M., Babiński, Z., Chalov, S., et al. (2020). Dam and reservoir removal projects: A mix of
social-ecological trends and cost-cutting attitudes. Scientific Reports, 10, 1–16. https://doi.org/10.1038/s41598-020-76158-3
Hager, C., Kahn, J., Watterson, C., Russo, J., & Hartman, K. (2014). Evidence of atlantic sturgeon spawning in the York river system. Transac-
tions of the American Fisheries Society, 143(5), 1217–1219. https://doi.org/10.1080/00028487.2014.925971
Haines, A., Finlayson, B., & McMahon, T. (1988). A global classification of river regimes. Applied Geography, 8(4), 255–272. https://doi.
org/10.1016/0143-6228(88)90035-5
Halim, Y. (1960). Observations on the Nile bloom of phytoplankton in the Mediterranean. ICES Journal of Marine Science, 26(1), 57–67. https://
doi.org/10.1093/icesjms/26.1.57
Halim, Y., Guergues, S. K., & Salah, H. H. (1964). Hydrodynamic conditions and plankton in the South East Mediterranean during the last nor mal
Nile flood. Internationale Revue der gesamten Hydrobiologie und Hydrographie, 52(3), 401–425. https://doi.org/10.1002/IROH.19670520305
Hallnan, R., Saito, L., Busby, D., & Tyler, S. (2020). Modeling Shasta Reservoir water-temperature response to the 2015 drought and response
under future climate change. Journal of Water Resources Planning and Management, 146(5), 04020018. https://doi.org/10.1061/(ASCE)
WR.1943-5452.0001186
Hanna, R. B., Saito, L., Bartholow, J. M., & Sandelin, J. (1999). Results of simulated temperature control device operations on
in-reservoir and discharge water temperatures using CE-QUAL-W2. Lake and Reservoir Management, 15(2), 87–102. https://doi.
org/10.1080/07438149909353954
Hannah, D. M., Malcolm, I. A., Soulsby, C., & Youngson, A. F. (2004). Heat exchanges and temperatures within a salmon spawning stream in the
Cairngorms, Scotland: Seasonal and sub-seasonal dynamics. River Research and Applications, 20(6), 635–652. https://doi.org/10.1002/rra.771
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
51 of 64
Hansford, M. R., Plink-Björklund, P., & Jones, E. R. (2020). Global quantitative analyses of river discharge variability and hydrograph shape with
respect to climate types. Earth-Science Reviews, 200, 102977. https://doi.org/10.1016/j.earscirev.2019.102977
Harlan, T., Xu, R., & He, J. (2021). Is small hydropower beautiful? Social impacts of river fragmentation in China's red River Basin. Ambio, 50(2),
436–447. https://doi.org/10.1007/s13280-020-01367-z
Harnish, R. A., Sharma, R., McMichael, G. A., Langshaw, R. B., & Pearsons, T. N. (2014). Effect of hydroelectric dam operations on the fresh-
water productivity of a Columbia River fall Chinook salmon population. Canadian Journal of Fisheries and Aquatic Sciences, 71(4), 602–615.
https://doi.org/10.1139/cjfas-2013-0276
Haro, A. (2014). Anguillidae: Freshwater eels. Freshwater Fishes of North America, 1, 313–331.
Harrison, L. R., East, A. E., Smith, D. P., Logan, J. B., Bond, R. M., Nicol, C. L., etal. (2018). River response to large-dam removal in a Medi-
terranean hydroclimatic setting: Carmel River, California, USA. Earth Surface Processes and Landforms, 45(15), 3009–3021. https://doi.
org/10.1002/esp.4464
Harvey, A. C., Glover, K. A., Wennevik, V., & Skaala, Ø. (2020). Atlantic salmon and sea trout display synchronised smolt migration relative to
linked environmental cues. Scientific Reports, 10(1), 3529. https://doi.org/10.1038/s41598-020-60588-0
Hatten, J. R., Batt, T. R., Skalicky, J. J., Engle, R., Barton, G. J., Fosness, R. L., & Warren, J. (2016). Effects of dam removal on Tule fall
Chinook salmon spawning habitat in the White Salmon River, Washington. River Research and Applications, 32(7), 1481–1492. https://doi.
org/10.1002/rra.2982
He, F., Thieme, M., Zarfl, C., Grill, G., Lehner, B., Hogan, Z., etal. (2021). Impacts of loss of free-flowing rivers on global freshwater megafauna.
Biological Conservation, 263, 109335. https://doi.org/10.1016/j.biocon.2021.109335
He, S., Tang, L., Wang, J., Zhu, C., Li, T., Mo, K., & Chen, Q. (2021). Effects of dam removal on fish community structure and spatial distribution
in Heishui River. Acta Ecologica Sinica, 41(9), 3525–3534. https://doi.org/10.5846/stxb202002050203
He, W., Ma, C., Zhang, J., Lian, J., Wang, S., & Zhao, W. (2020). Multi-objective optimal operation of a large deep reservoir during stor-
age period considering the outflow-temperature demand based on NSGA-II. Journal of Hydrology, 586, 124919. https://doi.org/10.1016/j.
jhydrol.2020.124919
Heer, T., Wells, M. G., & Mandrak, N. E. (2019). Assessment of Asian carp spawning potential in tributaries to the Canadian Lake Ontario basin.
Journal of Great Lakes Research, 45(6), 1332–1339. https://doi.org/10.1016/j.jglr.2019.09.019
Heggenes, J., Alfredsen, K., Bustos, A. A., Huusko, A., & Stickler, M. (2018). Be cool: A review of hydro-physical changes and fish responses in
winter in hydropower-regulated northern streams. Environmental Biology of Fishes, 101(1), 1–21. https://doi.org/10.1007/s10641-017-0677-z
Hennig, T., & Harlan, T. (2018). Shades of green energy: Geographies of small hydropower in Yunnan, China and the challenges of
over-development. Global Environmental Change, 49, 116–128. https://doi.org/10.1016/j.gloenvcha.2017.10.010
Hermann, T., Stewart, D., Limburg, K., & Castello, L. (2016). Unravelling the life history of Amazonian fishes through otolith microchemistry.
Royal Society Open Science, 3(6), 160206. https://doi.org/10.1098/rsos.160206
Hershey, H. (2021). Updating the consensus on fishway efficiency: A meta-analysis. Fish and Fisheries, 22(4), 735–748. https://doi.org/10.1111/
faf.12547
Hickford, M. J. H., & Schiel, D. R. (2013). Artificial spawning habitats improve egg production of a declining diadromous fish, Galaxias macu-
latus (Jenyns, 1842). Restoration Ecology, 21(6), 686–694. https://doi.org/10.1111/rec.12008
Higgins, J. M., & Brock, W. G. (1999). Overview of reservoir release improvements at 20 TVA dams. Journal of Energy Engineering, 125(1),
1–17. https://doi.org/10.1061/(ASCE)0733-9402(1999)125:1(1)
Hilborn, R. (2013). Ocean and dam influences on salmon survival. Proceedings of the National Academy of Sciences, 110(17), 6618–6619.
https://doi.org/10.1073/pnas.130365311
Hishamunda, N. (2007). Chapter 8: Aquaculture in Africa: Reasons for failures and ingredients for success. In P.S. Leung, C. S. Lee, & P.J.
O’Bryen (Eds.), Species and system selection for sustainable aquaculture (pp.103–116). Blackwell Publishing.
Hogan, T. W., Cada, G. F., & Amaral, S. V. (2014). The status of environmentally enhanced hydropower turbines. Fisheries, 39(4), 164–172.
https://doi.org/10.1080/03632415.2014.897195
Hogan, Z. (2011). Imperiled giant fish and mainstream Mekong dams in the lower Mekong Basin: Assessment of current status, threats, and
mitigation.
Hogan, Z., Baird, I., Radtke, R., & Vander Zanden, J. (2007). Long distance migration and marine habitation in the tropical Asian catfish,
Pangasius krempfi. Journal of Fish Biology, 71(3), 818–832. https://doi.org/10.1111/j.1095-8649.2007.01549.x
Holsman, K. K., Scheuerell, M., Buhle, E., & Emmett, R. (2012). Interacting effects of translocation, artificial propagation, and environmental
conditions on the marine survival of Chinook salmon from the Columbia River, Washington, USA. Conservation Biology, 26(5), 912–922.
https://doi.org/10.1111/j.1523-1739.2012.01895.x
Holt, C. R., Pfitzer, D., Scalley, C., Caldwell, B. A., & Batzer, D. P. (2015). Macroinvertebrate community responses to annual flow variation
from river regulation: An 11-year study. River Research and Applications, 31(7), 798–807. https://doi.org/10.1002/rra.2782
Hong, Y., Liu, D., Ma, H., Chen, Y., Chen, Q., Shi, W., etal. (2022). Effects of fish habitat substitution in tributaries under the cascade hydro-
power development of Lancang River. Acta Ecologica Sinica, 42(8), 3191–3205. https://doi.org/10.5846/stxb202009042303
Honsey, A. E., Venturelli, P.A., & Lester, N. P. (2018). Bioenergetic and limnological foundations for using degree-days derived from air temper-
atures to describe fish growth. Canadian Journal of Fisheries and Aquatic Sciences, 76(4), 657–669. https://doi.org/10.1139/cjfas-2018-0051
Huang, J., Li, R., Feng, J., Xu, W., & Wang, L. (2016). Relationship investigation between the dissipation process of supersaturated total dissolved
gas and wind effect. Ecological Engineering, 95(10), 430–437. https://doi.org/10.1016/J.ECOLENG.2016.06.042
Huang, J., Lu, J., Li, R., Feng, J., & Yang, H. (2021). Numerical study on the cumulative effect of supersaturated TDG through the spillway.
Ecohydrology and Hydrobiology, 21(2), 292–298. https://doi.org/10.1016/j.ecohyd.2021.01.003
Huang, S., & He, Y. (2019). Management of China's capture fisheries: Review and prospect. Aquaculture and Fisheries, 4(5), 173–182. https://
doi.org/10.1016/j.aaf.2019.05.004
Huang, X., Li, K. F., Du, J., & Li, R. (2010). Effects of gas supersaturation on lethality and avoidance responses in juvenile rock carp (Procypris
rabaudi Tchang). Journal of Zhejiang University - Science B, 11(10), 806–811. https://doi.org/10.1631/jzus.B1000006
Huang, Z. (2019). Drifting with flow versus self-migrating—How do young anadromous fish move to the sea? iScience, 19, 772–785. https://
doi.org/10.1016/j.isci.2019.08.029
Huang, Z., & Wang, L. (2018). Yangtze dams increasingly threaten the survival of the Chinese sturgeon. Current Biology, 28(22), 3640–3647.
https://doi.org/10.1016/j.cub.2018.09.032
Huchzermeyer, K. D. A. (2003). Clinical and pathological observations on Streptococcus sp.infection on South African trout farms with gas
supersaturated water supplies. Onderstepoort Journal of Veterinary Research, 70(2), 95–105.
Hughes, D. A., & Ziervogel, G. (1998). The inclusion of operating rules in a daily reservoir simulation model to determine ecological reserve
releases for river maintenance. WaterSA, 24(4), 293–302.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
52 of 64
Hughes, J. S., Khan, F., Liss, S. A., Harnish, R. A., Johnson, G. E., Bellgraph, B. B., & Znotinas, K. R. (2021). Evaluation of Juvenile salmon
passage and survival through a fish weir and other routes at Foster Dam, Oregon, USA. Fisheries Management and Ecology, 28(3), 241–252.
https://doi.org/10.1111/fme.12481
Im, D., Kang, H., Kim, K. H., & Choi, S. U. (2011). Changes of river morphology and physical fish habitat following weir removal. Ecological
Engineering, 37(6), 883–892. https://doi.org/10.1016/j.ecoleng.2011.01.005
Irvine, R. L., Thorley, J. L., Westcott, R., Schmidt, D., & DeRosa, D. (2015). Why do fish strand? An analysis of ten years of flow reduction moni-
toring data from the Columbia and Kootenay rivers, Canada. River Research and Applications, 31(10), 1242–1250. https://doi.org/10.1002/
rra.2823
Isaak, D., Wollrab, S., Horan, D., & Chandler, G. (2012). Climate change effects on stream and river temperatures across the northwest US from
1980–2009 and implications for salmonid fishes. Climatic Change, 113(2), 499–524. https://doi.org/10.1007/s10584-011-0326-z
IUCN. (2022). The IUCN red list of threatened species. Retrieved from https://www.iucnredlist.org/en
Jackson, R. I. (1950). Variations in flow patterns at Hell’s Gate and their relationships to the migration of sockeye salmon. New Westminster, B.
C. International pacific salmon fisheries commission. Bulletin, III, 81–129.
Jacobson, R. B., & Galat, D. L. (2008). Design of a naturalized flow regime-an example from the Lower Missouri River, USA. Ecohydrology,
1(2), 81–104. https://doi.org/10.1002/eco.9
Jacoby, D. M., Casselman, J. M., Crook, V., DeLucia, M., Ahn, H., Kaifu, K., etal. (2015). Synergistic patterns of threat and the challenges facing
global anguillid eel conservation. Global Ecology and Conservation, 4, 321–333. https://doi.org/10.1016/j.gecco.2015.07.009
Jansen, H. M., Winter, H. V., Bruijs, M. C., & Polman, H. J. (2007). Just go with the flow? Route selection and mortality during downstream
migration of silver eels in relation to river discharge. ICES Journal of Marine Science, 64(7), 1437–1443. https://doi.org/10.1093/icesjms/
fsm132
Jardim, P.F., Melo, M. M. M., Ribeiro, L. C., Collischonn, W., & Paz, A. R. (2020). A modeling assessment of large-scale hydrologic alter-
ation in South American Pantanal due to upstream dam operation. Frontiers in Environmental Science, 8, 1–15. https://doi.org/10.3389/
fenvs.2020.567450
Jarzyna, M. A., & Jetz, W. (2016). Detecting the multiple facets of biodiversity. Trends in Ecology & Evolution, 31(7), 527–538. https://doi.
org/10.1016/j.tree.2016.04.002
Jellyman, D. J. (2022). An enigma: How can freshwater eels (Anguilla spp.) be such a successful genus yet be universally threatened? Reviews in
Fish Biology and Fisheries, 32(2), 701–718. https://doi.org/10.1007/s11160-021-09658-8
Jellyman, D. J., & Arai, T. (2016). Juvenile eels: Upstream migration and habitat use. Biology and ecology of anguillid eels, 143–170. https://
doi.org/10.1201/b19925-9
Ji, Q., Xue, S., Yuan, Q., Yuan, Y., Wang, Y., Liang, R., etal. (2019). The tolerance characteristic of resident fish in the upper Yangtze River
under varying gas supersaturation. International Journal of Environmental Research and Public Health, 16(11), 2021. https://doi.org/10.3390/
ijerph16112021
Jiang, W., Liu, H., Duan, Z., & Cao, W. (2010). Seasonal variation in drifting eggs and larvae in the upper Yangtze, China. Zoological Science,
27(5), 402–409. https://doi.org/10.2108/zsj.27.402
Jiang, Z., Liu, P., Ji, C., Zhang, H., & Chen, Y. (2019). Ecological flow considered multi-objective storage energy operation chart optimization of
large-scale mixed reservoirs. Journal of Hydrology, 577, 123949. https://doi.org/10.1016/j.jhydrol.2019.123949
Johnsen, B. O., & Jensen, A. J. (1994). The spread of furunculosis in salmonids in Norwegian rivers. Journal of Fish Biology, 45(1), 47–55.
https://doi.org/10.1111/j.1095-8649.1994.tb01285.x
Johnson, E. L., Clabough, T. S., Bennett, D. H., Bjornn, T. C., Peery, C. A., Caudill, C. C., & Stuehrenberg, L. C. (2005). Migration depths of
adult spring and summer Chinook salmon in the lower Columbia and Snake Rivers in relation to dissolved gas supersaturation. Transactions
of the American Fisheries Society, 134(5), 1213–1227. https://doi.org/10.1577/T04-116.1
Johnson, E. L., Clabough, T. S., Peery, C. A., Bennett, D. H., Bjornn, T. C., Caudill, C. C., & Richmond, M. C. (2007). Estimating adult
Chinook salmon exposure to dissolved gas supersaturation downstream of hydroelectric dams using telemetry and hydrodynamic models.
River Research and Applications, 23(9), 963–978. https://doi.org/10.1002/rra.1019
Johnson, N. S., Buchinger, T. J., & Li, W. (2015). Reproductive ecology of lampreys. In Lampreys: Biology, conservation and control
(pp.265–303). Springer.
Johnston, M., Frantzich, J., Espe, M. B., Goertler, P., Singer, G., Sommer, T., & Klimley, A. P. (2020). Contrasting the migratory behavior and
stranding risk of White Sturgeon and Chinook Salmon in a modified floodplain of California. Environmental Biology of Fishes, 103(5),
481–493. https://doi.org/10.1007/s10641-020-00974-9
Jonsson, B., & Jonsson, N. (2009). A review of the likely effects of climate change on anadromous Atlantic salmon Salmo salar and brown
trout Salmo trutta, with particular reference to water temperature and flow. Journal of Fish Biology, 75(10), 2381–2447. https://doi.
org/10.1111/j.1095-8649.2009.02380.x
Jordan, F., & Arrington, D. A. (2014). Piscivore responses to enhancement of the channelized Kissimmee River, Florida, USA. Restoration Ecol-
ogy, 22(3), 418–425. https://doi.org/10.1111/rec.12060
Jung, E., Joo, G. J., Kim, H. G., Kim, D. K., & Kim, H. W. (2022). Effects of seasonal and diel variations in thermal stratification on phytoplank-
ton in a regulated river. In Biogeosciences discussions (pp.1–18). https://doi.org/10.5194/bg-2022-42
Junk, W. J., An, S., Finlayson, C., Gopal, B., Květ, J., Mitchell, S. A., etal. (2013). Current state of knowledge regarding the world’s wetlands and
their future under global climate change: A synthesis. Aquatic Sciences, 75(1), 151–167. https://doi.org/10.1007/s00027-012-0278-z
Kamal, R., David, Z., Asce, M., Leake, A., & Crossman, J. (2018). Dissipation of supersaturated total dissolved gases in the intermediate mixing
zone of a regulated river. Journal of Environmental Engineering, 145(2), 04018135. https://doi.org/10.1061/(ASCE)EE.1943-7870.0001477
Karim, F., Dutta, D., Marvanek, S., Petheram, C., Ticehurst, C., Lerat, J., etal. (2015). Assessing the impacts of climate change and dams on
floodplain inundation and wetland connectivity in the wet–dry tropics of northern Australia. Journal of Hydrology, 522, 80–94. https://doi.
org/10.1016/j.jhydrol.2014.12.005
Karr, J. R. (1991). Biological integrity: A long-neglected aspect of water resource management. Ecological Applications, 1(1), 66–84. https://
doi.org/10.2307/1941848
Katopodis, C., & Williams, J. G. (2012). The development of fish passage research in a historical context. Ecological Engineering, 48, 8–18.
https://doi.org/10.1016/j.ecoleng.2011.07.004
Keefer, M. L., Boggs, C. T., Peery, C. A., & Caudill, C. C. (2008). Overwintering distribution, behavior, and survival of adult summer
steelhead: Variability among Columbia River populations. North American Journal of Fisheries Management, 28(1), 81–96.
https://doi.org/10.1577/M07-011.1
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
53 of 64
Kemp, P. S., Russon, I. J., Vowles, A. S., & Lucas, M. C. (2011). The influence of discharge and temperature on the ability of upstream
migrant adult river lamprey (Lampetra fluviatilis) to pass experimental overshot and undershot weirs. River Research and Applications, 27(4),
488–498. https://doi.org/10.1002/rra.1364
Kennedy, J., Nash, R. D. M., Slotte, A., & Kjesbu, O. S. (2011). The role of fecundity regulation and abortive maturation in the reproductive strategy
of Norwegian spring-spawning herring (Clupea harengus). Marine Biology, 158(6), 1287–1299. https://doi.org/10.1007/s00227-011-1648-0
Keppeler, F. W., Andrade, M., Trindade, P., Sousa, L., Arantes, C., Winemiller, K., etal. (2022). Early impacts of the largest Amazonian hydro-
power project on fish communities. Science of the Total Environment, 838, 155951. https://doi.org/10.1016/j.scitotenv.2022.155951
Kerr, J. R., Karageorgopoulos, P., & Kemp, P.S. (2015). Efficacy of a side-mounted vertically oriented bristle pass for improving upstream
passage of European eel (Anguilla anguilla) and river lamprey (Lampetra fluviatilis) at an experimental Crump weir. Ecological Engineering,
85, 121–131. https://doi.org/10.1016/j.ecoleng.2015.09.013
Kim, S. K., & Choi, S. U. (2019). Ecological evaluation of weir removal based on physical habitat simulations for macroinvertebrate community.
Ecological Engineering, 138, 362–373. https://doi.org/10.1016/j.ecoleng.2019.08.003
King, A. J., Ward, K. A., O'Connor, P., Green, D., Tonkin, Z., & Mahoney, J. (2010). Adaptive management of an environmental watering event
to enhance native fish spawning and recruitment. Freshwater Biology, 55(1), 17–31. https://doi.org/10.1111/j.1365-2427.2009.02178.x
King, J., Cambray, J. A., & Impson, N. D. (1998). Linked effects of dam-released floods and water temperature on spawning of the Clanwilliam
yellowfish Barbus capensis. Hydrobiologia, 384(1), 245–265. https://doi.org/10.1023/A:1003481524320
Klemetsen, A., Amundsen, P.A., Dempson, J. B., Jonsson, B., Jonsson, N., O'connell, M. F., & Mortensen, E. (2003). Atlantic salmon Salmo
salar L., brown trout Salmo trutta L. and arctic charr Salvelinus alpinus (L.): A review of aspects of their life histories. Ecology of Freshwater
Fish, 12(1), 1–59. https://doi.org/10.1034/j.1600-0633.2003.00010.x
Kobayashi, S., Nakanishi, S., Akamatsu, F., Yajima, Y., & Amano, K. (2012). Differences in amounts of pools and riffles between upper and
lower reaches of a fully sedimented dam in a mountain gravel-bed river. Landscape and Ecological Engineering, 8(2), 145–155. https://doi.
org/10.1007/s11355-011-0156-1
Kondolf, G. M., Rubin, Z. K., & Minear, J. T. (2014). Dams on the Mekong: Cumulative sediment starvation. Water Resources Research, 50(6),
5158–5169. https://doi.org/10.1002/2013WR014651
Kondolf, G. M., Schmitt, R., Darby, S., Carling, P., Arias, M., Bizzi, S., etal. (2018). Changing sediment budget of the Mekong: Cumula-
tive threats and management strategies for a large river basin. Science of the Total Environment, 625, 114–134. https://doi.org/10.1016/j.
scitotenv.2017.11.361
Kong, D., Miao, C., Wu, J., Borthwick, A. G., Duan, Q., & Zhang, X. (2017). Environmental impact assessments of the Xiaolangdi Reservoir on
the most hyperconcentrated laden river, Yellow River, China. Environmental Science and Pollution Research, 24(5), 4337–4351. https://doi.
org/10.1007/s11356-016-7975-4
Kotti, F. C., Mahe, G., Habaieb, H., Dieulin, C., Calvez, R., & Ali, H. B. (2016). Etude des pluies et des débits sur le bassin versant de la
Medjerda, Tunisie. Study of rainfall and discharges in the Medjerda watershed, Tunisia (Vol. 38,pp.19–28). Bulletin de l’Institut Scientifique.
Kruk, A., & Penczak, T. (2003). Impoundment impact on populations of facultative riverine fish. Annales de Limnologie-International Journal
of Limnology, 39(3), 197–210. https://doi.org/10.1051/LIMN/2003016
Krull, W., Berry, P.M., Bauduceau, N., Elmqvist, T., Fernandez, M., Hartig, T., etal. (2015). Towards an EU research and innovation policy
agenda for nature-based solutions & Re-naturing cities: Final report of the horizon 2020 expert group on ‘nature-based solutions and
Re-naturing cities’ (full version). European Commission, Directorate-General for Research and Innovation. https://doi.org/10.2777/765301
Kuczynski, L., Chevalier, M., Laffaille, P., Legrand, M., & Grenouillet, G. (2017). Indirect effect of temperature on fish population abundances
through phenological changes. PLoS One, 12(4), e0175735. https://doi.org/10.1371/journal.pone.0175735
Kurita, Y., Meier, S., & Kjesbu, O. (2003). Oocyte growth and fecundity regulation by atresia of Atlantic herring (Clupea harengus) in relation
to body condition throughout the maturation cycle. Journal of Sea Research, 49(3), 203–219. https://doi.org/10.1016/S1385-1101(03)00004-2
Kwak, T. J., Engman, A. C., & Lilyestrom, C. G. (2019). Ecology and conservation of the American eel in the Caribbean region. Fisheries
Management and Ecology, 26(1), 42–52. https://doi.org/10.1111/FME.12300
Laine, A. (1995). Fish swimming behaviour in Finnish fishways. In S. Komura (Ed.), Proceedings of the international symposium on fishways ‘95
in Gifu, Japan. Gifu: Organising committee for international symposium on fishways (pp.323–328).
Langeani, F., Casatti, L., Gameiro, H. S., Carmo, A. B., & Rossa-Feres, D. C. (2005). Riffle and pool fish communities in a large stream of
southeastern Brazil. Neotropical Ichthyology, 3(2), 305–311. https://doi.org/10.1590/S1679-62252005000200009
Larinier, M. (2008). Fish passage experience at small-scale hydro-electric power plants in France. Hydrobiologia, 609(1), 97–108. https://doi.
org/10.1007/s10750-008-9398-9
Lasne, E., Sabatié, M. R., Jeannot, N., & Cucherousset, J. (2015). The effects of dam removal on river colonization by sea lamprey Petromyzon
marinus. River Research and Applications, 31(7), 904–911. https://doi.org/10.1002/rra.2789
Latrubesse, E. M. (2008). Patterns of anabranching channels: The ultimate end-member adjustment of mega rivers. Geomorphology, 101(1–2),
130–145. https://doi.org/10.1016/j.geomorph.2008.05.035
Latrubesse, E. M., Arima, E. Y., Dunne, T., Park, E., Baker, V. R., d’Horta, F. M., etal. (2017). Damming the rivers of the Amazon basin. Nature,
546(7658), 363–369. https://doi.org/10.1038/nature22333
Latrubesse, E. M., d’Horta, F. M., Ribas, C. C., Wittmann, F., Zuanon, J., Park, E., etal. (2020). Vulnerability of the biota in riverine and season-
ally flooded habitats to damming of Amazonian rivers. Aquatic Conservation: Marine and Freshwater Ecosystems, 31(5), 1136–1149. https://
doi.org/10.1002/aqc.3424
Latrubesse, E. M., Stevaux, J. C., & Sinha, R. (2005). Tropical rivers. Geomorphology, 70(3–4), 187–206. https://doi.org/10.1016/j.
geomorph.2005.02.005
Lawrence, E. R., Kuparinen, A., & Hutchings, J. A. (2016). Influence of dams on population persistence in Atlantic salmon (Salmo salar). Cana-
dian Journal of Zoology, 94(5), 329–338. https://doi.org/10.1139/cjz-2015-0195
Le Coarer, Y., Lizée, M. H., Beche, L., & Logez, M. (2022). Horizontal ramping rate framework to quantify hydropeaking stranding risk for fish.
River Research and Applications, 39(3), 1–12. https://doi.org/10.1002/rra.4087
Lees, A., Peres, C., Fearnside, P., Schneider, M., & Zuanon, J. (2016). Hydropower and the future of Amazonian biodiversity. Biodiversity &
Conservation, 25(3), 451–466. https://doi.org/10.1007/s10531-016-1072-3
Lehner, B., Liermann, C., Revenga, C., Vörösmarty, C., Fekete, B., Crouzet, P., etal. (2011). High-resolution mapping of the world's reservoirs
and dams for sustainable river-flow management. Frontiers in Ecology and the Environment, 9(9), 494–502. https://doi.org/10.1890/100125
Leinonen, T., Piironen, J., Koljonen, M. L., Koskiniemi, J., & Kause, A. (2020). Restored river habitat provides a natural spawning area for
a critically endangered landlocked Atlantic salmon population. PLoS One, 15(5), e0232723. https://doi.org/10.1371/journal.pone.0232723
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
54 of 64
Le Luyer, J., Laporte, M., Beacham, T. D., Kaukinen, K. H., Withler, R. E., Leong, J. S., etal. (2017). Parallel epigenetic modifications induced
by hatchery rearing in a Pacific salmon. Proceedings of the National Academy of Sciences, 114(49), 12964–12969. https://doi.org/10.1073/
pnas.1711229114
Lemarie, G., Hosfeld, C. D., Breuil, G., & Fivelstad, S. (2011). Effects of hyperoxic water conditions under different total gas pressures in Euro-
pean sea bass (Dicentrarchus labrax). Aquaculture, 318(1–2), 191–198. https://doi.org/10.1016/j.aquaculture.2011.03.033
Lenhardt, M., Jaric, I., Kalauzi, A., & Cvijanovic, G. (2006). Assessment of extinction risk and reasons for decline in sturgeon. Biodiversity &
Conservation, 15(6), 1967–1976. https://doi.org/10.1007/s10531-005-4317-0
Leopold, L. B., & Wolman, M. G. (1957). River channel patterns: Braided, meandering, and straight. US Government Printing Office. https://
doi.org/10.3133/PP282B
Lessard, J. L., & Hayes, D. B. (2003). Effects of elevated water temperature on fish and macroinvertebrate communities below small dams. River
Research and Applications, 19(7), 721–732. https://doi.org/10.1002/rra.713
Levin, P.S., & Tolimieri, N. (2001). Differences in the impacts of dams on the dynamics of salmon populations. Animal Conservation, 4(4),
291–299. https://doi.org/10.1017/S1367943001001342
Li, C., Wang, H., Chen, H., Zhao, Y., Fu, P., & Ji, X. (2014). Differential expression analysis of genes involved in high-temperature induced
sex differentiation in Nile tilapia. Comparative Biochemistry and Physiology Part B: Biochemistry and Molecular Biology, 177–178, 36–45.
https://doi.org/10.1016/j.cbpb.2014.08.006
Li, D., Lu, X., Overeem, I., Walling, D. E., Syvitski, J., Kettner, A. J., etal. (2021). Exceptional increases in fluvial sediment fluxes in a warmer
and wetter High Mountain Asia. Science, 374(6567), 599–603. https://doi.org/10.1126/science.abi9649
Li, D., Prinyawiwatkul, W., Tan, Y., Luo, Y., & Hong, H. (2021). Asian carp: A threat to American lakes, a feast on Chinese tables. Comprehen-
sive Reviews in Food Science and Food Safety, 20(3), 2968–2990. https://doi.org/10.1111/1541-4337.12747
Li, J., Du, H., Wu, J., Zhang, H., Shen, L., & Wei, Q. (2021). Foundation and prospects of wild population reconstruction of Acipenser dabryanus.
Fishes, 6(4), 55. https://doi.org/10.3390/fishes6040055
Li, J., Qin, H., Pei, S., Yao, L., Wen, W., Yi, L., etal. (2019). Analysis of an ecological flow regime during the Ctenopharyngodon idella spawning
period based on reservoir operations. Water, 11(10), 2034. https://doi.org/10.3390/w11102034
Li, M., Duan, Z., Gao, X., Cao, W., & Liu, H. (2016). Impact of the Three Gorges Dam on reproduction of four major Chinese carps species
in the middle reaches of the Changjiang River. Chinese Journal of Oceanology and Limnology, 34(5), 885–893. https://doi.org/10.1007/
s00343-016-4303-2
Li, M., Gao, X., Yang, S., Duan, Z., Cao, W., & Liu, H. (2013). Effects of environmental factors on natural reproduction of the four major Chinese
carps in the Yangtze River, China. Zoological Science, 30(4), 296–303. https://doi.org/10.2108/zsj.30.296
Li, N., Fu, C., Zhang, J., Liu, X., Shi, X., Yang, Y., & Shi, H. (2019). Hatching rate of Chinese sucker (Myxocyprinus asiaticus Bleeker) eggs
exposed to total dissolved gas (TDG) supersaturation and the tolerance of juveniles to the interaction of TDG supersaturation and suspended
sediment. Aquaculture Research, 50(7), 1876–1884. https://doi.org/10.1111/are.14071
Li, P., Zhu, D. Z., Li, R., Wang, Y., Crossman, J. A., & Kuhn, W. L. (2022). Production of total dissolved gas supersaturation at hydropower
facilities and its transport: A review. Water Research, 223, 119012. https://doi.org/10.1016/j.watres.2022.119012
Li, Q. (2023). Data for 2023RG000819 [Dataset]. figshare. https://doi.org/10.6084/m9.figshare.24419893.v2
Li, T., Chen, Q., Zhang, Q., Feng, T., Zhang, J., Lin, Y., etal. (2022). Transcriptomic analysis on the effects of altered water temperature regime
on the fish ovarian development of Coreius guichenoti under the impact of river damming. Biology, 11(12), 1829. https://doi.org/10.3390/
biology11121829
Li, T., Mo, K., Wang, J., Chen, Q., Zhang, J., Zeng, C., etal. (2021). Mismatch between critical and accumulated temperature following river
damming impacts fish spawning. Science of the Total Environment, 756, 144052. https://doi.org/10.1016/j.scitotenv.2020.144052
Liang, R., Li, B., Li, K., & Tuo, Y. (2013). Effect of total dissolved gas supersaturated water on early life of David’s schizothoracin (Schizothorax
davidi). Journal of Zhejiang University - Science B, 14(7), 632–639. https://doi.org/10.1631/jzus.b1200364
Liang, Y., Liu, D., Shi, X., Xu, Y., Wang, C., Zheng, X., etal. (2014). Overview of fish barging. Journal of Yangtze River Scientific Research
Institute, 31(02), 25–29. https://doi.org/10.3969/j.issn.1001-5485
Liermann, C. R., Nilsson, C., Robertson, J., & Ng, R. Y. (2012). Implications of dam obstruction for global freshwater fish diversity. BioScience,
62(6), 539–548. https://doi.org/10.1525/bio.2012.62.6.5
Limburg, K. E., & Waldman, J. R. (2009). Dramatic declines in North Atlantic diadromous fishes. BioScience, 59(11), 955–965. https://doi.
org/10.1525/bio.2009.59.11.7
Lin, Z., Kitakado, T., Suzuki, N., & Ito, S. (2022). Evaluation of the effects of stock enhancement on population dynamics using a state-space
production model: A case study of Japanese flounder in the Seto inland sea. Fisheries Research, 251, 106299. https://doi.org/10.1016/j.
fishres.2022.106299
Lisle, T. E., Iseya, F., & Ikeda, H. (1993). Response of a channel with alternate bars to a decrease in supply of mixed-size bed load: A flume
experiment. Water Resources Research, 29(11), 3623–3629. https://doi.org/10.1029/93WR01673
Liu, C., Qiu, J., & Li, F. (2019). Recovery degree of the natural flow regimes and the corresponding economic costs for reservoir operation in fish
spawning seasons. International Journal of Environmental Research and Public Health, 16(10), 1699. https://doi.org/10.3390/ijerph16101699
Liu, L., Wu, G., Cao, W., & Wang, Z. (1986). Studies on the ecological effect on spawning of the black carp, the grass carp, the silver carp and
the bighead carp in the Changjiang River after the constructions of the Gezhouba hydroelectric project. Acta Hydrobiologica Sinica, 10(4),
353–364.
Liu, Q., Dai, H., Gui, D., Hu, B. X., Ye, M., Wei, G., etal. (2022). Evaluation and optimization of the water diversion system of ecohydrological
restoration megaproject of Tarim River, China, through wavelet analysis and a neural network. Journal of Hydrology, 608, 127586. https://doi.
org/10.1016/j.jhydrol.2022.127586
Liu, S. (2013). Study on the effect of temperature on the formation and release of supersaturated total dissolved gases. Sichuan University.
Liu, S., Xie, Z., Liu, B., Wang, Y., Gao, J., Zeng, Y., etal. (2020). Global river water warming due to climate change and anthropogenic heat
emission. Global and Planetary Change, 193, 103289. https://doi.org/10.1016/j.gloplacha.2020.103289
Lloyst, M. H., Pratt, T. C., Reid, S. M., & Fox, M. G. (2015). Nearshore habitat associations of stocked American eel, Anguilla rostrata, in Lake
Ontario and the upper St. Lawrence River. Journal of Great Lakes Research, 41(3), 881–889. https://doi.org/10.1016/j.jglr.2015.05.012
Long, L., Ji, D., Liu, D., Yang, Z., & Lorke, A. (2019). Effect of cascading reservoirs on the flow variation and thermal regime in the lower
reaches of the Jinsha River. Water, 11(5), 1008. https://doi.org/10.3390/w11051008
Lu, J., Li, R., Ma, Q., Feng, J., Xu, W., Zhang, F., & Tian, Z. (2019). Model for total dissolved gas supersaturation from plunging jets in high dams.
Journal of Hydraulic Engineering, 145(1), 04018082. https://doi.org/10.1061/(ASCE)HY.1943-7900.0001550
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
55 of 64
Lubejko, M. V., Whitledge, G. W., Coulter, A. A., Brey, M. K., Oliver, D. C., & Garvey, J. E. (2017). Evaluating upstream passage and timing
of approach by adult bigheaded carps at a gated dam on the Illinois River. River Research and Applications, 33(8), 1268–1278. https://doi.
org/10.1002/rra.3180
Lundqvist, H., Rivinoja, P., Leonardsson, K., & McKinnell, S. (2008). Upstream passage problems for wild Atlantic salmon (Salmo salar L.) in a
regulated river and its effect on the population. Hydrobiologia, 602(1), 111–127. https://doi.org/10.1007/s10750-008-9282-7
Luo, X., Yang, S., & Zhang, J. (2012). The impact of the Three Gorges Dam on the downstream distribution and texture of sediments along the
middle and lower Yangtze River (Changjiang) and its estuary, and subsequent sediment dispersal in the East China Sea. Geomorphology, 179,
126–140. https://doi.org/10.1016/j.geomorph.2012.05.034
Lyu, S., Lin, K., Zeng, J., Liu, Y., Chen, Z., & Wang, X. (2021). Fin-spines attachment, a novel external attachment method for the ultrasonic
transmitters on hard fin-spines fish (Sparidae). Journal of Applied Ichthyology, 37(2), 227–234. https://doi.org/10.1111/jai.14164
Ma, C., Xu, R., He, W., & Xia, J. (2020). Determining the limiting water level of early flood season by combining multiobjective optimization
scheduling and copula joint distribution function: A case study of three gorges reservoir. Science of the Total Environment, 737, 139789.
https://doi.org/10.1016/j.scitotenv.2020.139789
Ma, Q., Li, R., Feng, J., Lu, J., & Zhou, Q. (2018). Cumulative effects of cascade hydropower stations on total dissolved gas supersaturation.
Environmental Science and Pollution Research, 25(14), 13536–13547. https://doi.org/10.1007/s11356-018-1496-2
Maavara, T., Chen, Q., Van Meter, K., Brown, L. E., Zhang, J., Ni, J., & Zarfl, C. (2020). River dam impacts on biogeochemical cycling. Nature
Reviews Earth & Environment, 1(2), 103–116. https://doi.org/10.1038/s43017-019-0019-0
MacCrimmon, H. R., & Gots, B. L. (1979). World distribution of Atlantic salmon, Salmo solar. Journal of the Fisheries Board of Canada, 36(4),
422–457. https://doi.org/10.1139/F79-062
MacCrimmon, H. R., Marshall, T. L., & Gots, B. L. (1970). World distribution of brown trout, Salmo trutta: Further observations. Journal of the
Fisheries Board of Canada, 27(4), 811–818. https://doi.org/10.1139/F70-085
Maes, J., & Jacobs, S. (2017). Nature-based solutions for Europe's sustainable development. Conservation Letters, 10(1), 121–124. https://doi.
org/10.1111/conl.12216
Magilligan, F. J., Graber, B. E., Nislow, K. H., Chipman, J. W., Sneddon, C. S., & Fox, C. A. (2016). River restoration by dam removal: Enhanc-
ing connectivity at watershed scales. Elementa: Science of the Anthropocene, 4, 000108. https://doi.org/10.12952/journal.elementa.000108
Magilligan, F. J., & Nislow, K. H. (2005). Changes in hydrologic regime by dams. Geomorphology, 71(1–2), 61–78. https://doi.org/10.1016/j.
geomorph.2004.08.017
Magilligan, F. J., Nislow, K. H., Kynard, B. E., & Hackman, A. M. (2016). Immediate changes in stream channel geomorphology, aquatic
habitat, and fish assemblages following dam removal in a small upland catchment. Geomorphology, 252, 158–170. https://doi.org/10.1016/j.
geomorph.2015.07.027
Maheu, A., St-Hilaire, A., Caissie, D., & El-Jabi, N. (2016a). Understanding the thermal regime of rivers influenced by small and medium size
dams in Eastern Canada. River Research and Applications, 32(10), 2032–2044. https://doi.org/10.1002/rra.3046
Maheu, A., St-Hilaire, A., Caissie, D., El-Jabi, N., Bourque, G., & Boisclair, D. (2016b). A regional analysis of the impact of dams on water
temperature in medium-size rivers in eastern Canada. Canadian Journal of Fisheries and Aquatic Sciences, 73(12), 1885–1897. https://doi.
org/10.1139/cjfas-2015-0486
Maitland, P.S. (2003). Ecology of the river, brook and sea lamprey. In Peterborough: Conserving Natura 2000 rivers ecology series no. 5. English
nature.
Malhi, Y., Doughty, C., Galetti, M., Smith, F., Svenning, J. C., & Terborgh, J. (2016). Megafauna and ecosystem function from the Pleistocene to
the Anthropocene. Proceedings of the National Academy of Sciences, 113(4), 838–846. https://doi.org/10.1073/pnas.1502540113
Mallen-Cooper, M., & Stuart, I. G. (2007). Optimising Denil fishways for passage of small and large fishes. Fisheries Management and Ecology,
14(1), 61–71. https://doi.org/10.1111/j.1365-2400.2006.00524.x
Mameri, D., Rivaes, R., Oliveira, J. M., Pádua, J., Ferreira, M. T., & Santos, J. M. (2019). Passability of Potamodromous species through a fish
lift at a large hydropower plant (Touvedo, Portugal). Sustainability, 12(1), 172. https://doi.org/10.3390/su12010172
Markovic, D., Scharfenberger, U., Schmutz, S., Pletterbauer, F., & Wolter, C. (2013). Variability and alterations of water temperatures across the
Elbe and Danube River basins. Climatic Change, 119(2), 375–389. https://doi.org/10.1007/s10584-013-0725-4
Marohn, L., Prigge, E., & Hanel, R. (2014). Escapement success of silver eels from a German river system is low compared to management-based
estimates. Freshwater Biology, 59(1), 64–72. https://doi.org/10.1111/FWB.12246
Marques, H., Dias, J. H. P., Perbiche-Neves, G., Kashiwaqui, E. A. L., & Ramos, I. P. (2018). Importance of dam-free tributaries for conserving
fish biodiversity in Neotropical reservoirs. Biological Conservation, 224, 347–354. https://doi.org/10.1016/J.BIOCON.2018.05.027
Marriner, B. A., Baki, A. B. M., Zhu, D. Z., Cooke, S. J., & Katopodis, C. (2016). The hydraulics of a vertical slot fishway: A case study
on the multi-species Vianney-Legendre fishway in Quebec, Canada. Ecological Engineering, 90, 190–202. https://doi.org/10.1016/J.
ECOLENG.2016.01.032
Marschall, E. A., Mather, M. E., Parrish, D. L., Allison, G. W., & McMenemy, J. R. (2011). Migration delays caused by anthropogenic
barriers: Modeling dams, temperature, and success of migrating salmon smolts. Ecological Applications, 21(8), 3014–3031. https://doi.
org/10.1890/10-0593.1
Marshall, J., Davison, A. J., Kopf, R. K., Boutier, M., Stevenson, P., & Vanderplasschen, A. (2018). Biocontrol of invasive carp: Risks abound.
Science, 359(6378), 877. https://doi.org/10.1126/science.aar7827
Martin, B. T., Dudley, P.N., Kashef, N. S., Stafford, D. M., Reeder, W. J., Tonina, D., etal. (2020). The biophysical basis of thermal tolerance
in fish eggs: Thermal tolerance in fish eggs. Proceedings of the Royal Society B: Biological Sciences, 287(1937), 20201550. https://doi.
org/10.1098/rspb.2020.1550
Maxwell, S. L., Cazalis, V., Dudley, N., Hoffmann, M., Rodrigues, A. S. L., Stolton, S., etal. (2020). Area-based conservation in the twenty-first
century. Nature, 586(7828), 217–227. https://doi.org/10.1038/s41586-020-2773-z
McDonald, D. G., & Kolar, C. S. (2007). Research to guide the use of lampricides for controlling sea lamprey. Journal of Great Lakes Research,
33(sp2), 20–34. https://doi.org/10.3394/0380-1330(2007)33[20:RTGTUO]2.0.CO;2
McDowall, R. M. (1997). The evolution of diadromy in fishes (revisited) and its place in phylogenetic analysis. Reviews in Fish Biology and
Fisheries, 7(4), 443–462. https://doi.org/10.1023/A:1018404331601
McGarvey, D. J., & Terra, B. F. (2016). Using river discharge to model and deconstruct the latitudinal diversity gradient for fishes of the Western
Hemisphere. Journal of Biogeography, 43(7), 1436–1449. https://doi.org/10.1111/jbi.12618
McIntyre, P., Reidy Liermann, C., & Revenga, C. (2016). Linking freshwater fishery management to global food security and biodi-
versity conservation. Proceedings of the National Academy of Sciences of the United States of America, 113(45), 12880–12885.
https://doi.org/10.1073/pnas.1521540113
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
56 of 64
McLaughlin, R. L., Porto, L., Noakes, D. L., Baylis, J. R., Carl, L. M., Dodd, H. R., etal. (2006). Effects of low-head barriers on stream
fishes: Taxonomic affiliations and morphological correlates of sensitive species. Canadian Journal of Fisheries and Aquatic Sciences, 63(4),
766–779. https://doi.org/10.1139/f05-256
McNabb, C. D., Liston, C. R., & Borthwick, S. M. (2003). Passage of juvenile chinook salmon and other fish species through Archimedes lifts
and a hidrostal Pump at red Bluff, California. Transactions of the American Fisheries Society, 132(2), 326–334. https://doi.org/10.1577/1548
-8659(2003)132<0326:POJCSA>2.0.CO;2
Mcqueen, K., & Marshall, C. T. (2017). Shifts in spawning phenology of cod linked to rising sea temperatures. ICES Journal of Marine Science,
74(6), 1561–1573. https://doi.org/10.1093/icesjms/fsx025
Meister, J., Selz, O. M., Beck, C., Peter, A., Albayrak, I., & Boes, R. M. (2022). Protection and guidance of downstream moving fish with hori-
zontal bar rack bypass systems. Ecological Engineering, 178, 106584. https://doi.org/10.1016/j.ecoleng.2022.106584
Melis, T. S., Walters, C. J., & Korman, J. (2015). Surprise and oppor tunity for learning in Grand Canyon: The Glen Canyon dam adaptive manage-
ment program. Ecology and Society, 20(3), 22. https://doi.org/10.5751/ES-07621-200322
Mensinger, M. A., Brehm, A. M., Mortelliti, A., Blomberg, E. J., & Zydlewski, J. D. (2021). American eel personality and body length influence
passage success in an experimental fishway. Journal of Applied Ecology, 58(12), 2760–2769. https://doi.org/10.1111/1365-2664.14009
Meyers, T., Burton, T., Bentz, C., & Starkey, N. (2008). Common diseases of wild and cultured fishes in Alaska. The Alaska Department of Fish
and Game.
Miao, C., Kong, D., Wu, J., & Duan, Q. (2016). Functional degradation of the water–sediment regulation scheme in the lower Yellow River:
Spatial and temporal analyses. Science of the Total Environment, 551–552, 16–22. https://doi.org/10.1016/j.scitotenv.2016.02.006
Miao, Y., Li, J., Feng, P., Dong, L., Zhang, T., Wu, J., & Katwal, R. (2020). Effects of land use changes on the ecological operation of the
Panjiakou-Daheiting Reservoir system, China. Ecological Engineering, 152, 105851. https://doi.org/10.1016/j.ecoleng.2020.105851
Milardi, M., Chapman, D., Lanzoni, M., Long, J. M., & Castaldelli, G. (2017). First evidence of bighead carp wild recruitment in Western Europe,
and its relation to hydrology and temperature. PLoS One, 12(12), e0189517. https://doi.org/10.1371/journal.pone.0189517
Miller, M. J., & Casselman, J. M. (2014). The American eel: A fish of mystery and sustenance for humans. In Eels and humans (pp.155–169).
Springer. https://doi.org/10.1007/978-4-431-54529-3_11
Miller, M. J., & Tsukamoto, K. (2016). The ecology of oceanic dispersal and survival of Anguillid leptocephali. Canadian Journal of Fisheries
and Aquatic Sciences, 74(6), 958–971. https://doi.org/10.1139/CJFAS-2016-0281
Milliman, J. D. (1997). Blessed dams or damned dams? Nature, 386(6623), 325–327. https://doi.org/10.1038/386325a0
Milliman, J. D., & Syvitski, J. P. (1992). Geomorphic/tectonic control of sediment discharge to the ocean: The importance of small mountainous
rivers. The Journal of Geology, 100(5), 525–544. https://doi.org/10.1086/629606
Milly, P.C., Dunne, K. A., & Vecchia, A. V. (2005). Global pattern of trends in streamflow and water availability in a changing climate. Nature,
438(7066), 347–350. https://doi.org/10.1038/nature04312
Mims, M. C., & Olden, J. D. (2012). Life history theory predicts fish assemblage response to hydrologic regimes. Ecology, 93(1), 35–45. https://
doi.org/10.1890/11-0370.1
Mohamed, N. N. (2019). Negative impacts of Egyptian high aswan dam: Lessons for Ethiopia and Sudan. International Journal of Development
Research, 09(08), 28861–28874.
Moi, D. A., Ernandes-Silva, J., Baumgartner, M. T., & Mormul, R. P. (2020). The effects of river-level oscillations on the macroinvertebrate
community in a river–floodplain system. Limnology, 21(2), 219–232. https://doi.org/10.1007/s10201-019-00605-y
Molina-Moctezuma, A., Peterson, E., & Zydlewski, J. D. (2021). Movement, survival, and delays of Atlantic salmon smolts in the Piscataquis
River, Maine, USA. Transactions of the American Fisheries Society, 150(3), 345–360. https://doi.org/10.1002/tafs.10289
Molony, B. W., Lenanton, R., Jackson, G., & Norriss, J. (2003). Stock enhancement as a fisheries management tool. Reviews in Fish Biology and
Fisheries, 13(4), 409–432. https://doi.org/10.1007/s11160-005-1886-7
Monk, B., Weaver, D., Thompson, C., & Ossiander, F. (1989). Effects of flow and weir design on the passage behavior of Ameri-
can shad and salmonids in an experimental fish ladder. North American Journal of Fisheries Management, 9(1), 60–67. https://doi.
org/10.1577/1548-8675(1989)009<0060:EOFAWD>2.3.CO;2
Moors, E. J., Goot, A., Biemans, H., van Scheltingaa, C. T., Sideriusa, C., Stoffelbc, M., etal. (2011). Adaptation to changing water resources in
the Ganges basin, northern India. Environmental Science & Policy, 14(7), 758–769. https://doi.org/10.1016/j.envsci.2011.03.005
Moran, E., Lopez, M. C., Moore, N., Müller, N., & Hyndman, D. (2018). Sustainable hydropower in the 21st century. Proceedings of the National
Academy of Sciences, 115(47), 11891–11898. https://doi.org/10.1073/pnas.1809426115
Morita, K., Morita, S. H., & Yamamoto, S. (2009). Effects of habitat fragmentation by damming on salmonid fishes: Lessons from white-spotted
charr in Japan. Ecological Research, 24(4), 711–722. https://doi.org/10.1007/s11284-008-0579-9
Morris, G. L., & Fan, J. (2010). Reservoir sedimentation handbook. In Electronic version 1.04. McGraw Hill.
Moser, M. L., Almeida, P.R., King, J. J., & Pereira, E. (2021). Passage and freshwater habitat requirements of anadromous lampreys: Considera-
tions for conservation and control. Journal of Great Lakes Research, 47(1), S147–S158. https://doi.org/10.1016/j.jglr.2020.07.011
Moser, M. L., & Close, D. A. (2003). Assessing pacific lamprey status in the Columbia River basin. Northwest Science, 77(2), 116–125.
Moser, M. L., Corbett, S. C., Keefer, M. L., Frick, K. E., Lopez-Johnston, S., & Caudill, C. C. (2019). Novel fishway entrance modifications for
Pacific lamprey. Journal of Ecohydraulics, 4(1), 71–84. https://doi.org/10.1080/24705357.2019.1604090
Moutinho, S. (2023). A river’s pulse. Science, 379(6627), 18–23. https://doi.org/10.1126/science.adg4958
Moyo, N. A. G., & Rapatsa, M. M. (2021). A review of the factors affecting tilapia aquaculture production in Southern Africa. Aquaculture, 535,
736386. https://doi.org/10.1016/j.aquaculture.2021.736386
Muir, W. D., Smith, S. G., Williams, J. G., & Sandford, B. P. (2001). Survival of juvenile salmonids passing through bypass systems, turbines, and
spillways with and without flow deflectors at Snake River dams. North American Journal of Fisheries Management, 21(1), 135–146. https://
doi.org/10.1577/1548-8675(2001)021<0135:SOJSPT>2.0.CO;2
Naganna, S. R., & Deka, P.C. (2018). Variability of streambed hydraulic conductivity in an intermittent stream reach regulated by vented dams:
A case study. Journal of Hydrology, 562, 477–491. https://doi.org/10.1016/j.jhydrol.2018.05.006
Nagrodski, A., Raby, G. D., Hasler, C. T., Taylor, M. K., & Cooke, S. J. (2012). Fish stranding in freshwater systems: Sources, consequences, and
mitigation. Journal of Environmental Management, 103, 133–141. https://doi.org/10.1016/j.jenvman.2012.03.007
Naish, K. A., Taylor, J. E., III, Levin, P.S., Quinn, T. P., Winton, J. R., Huppert, D., & Hilborn, R. (2007). An evaluation of the effects of conser-
vation and fishery enhancement hatcheries on wild populations of salmon. Advances in Marine Biology, 53, 61–194. https://doi.org/10.1016/
S0065-2881(07)53002-6
Namour, P., Schmitt, L., Eschbach, D., Moulin, B., Fantino, G., Bordes, C., & Breil, P. (2015). Stream pollution concentration in riffle geomor-
phic units (Yzeron basin, France). Science of the Total Environment, 532, 80–90. https://doi.org/10.1016/j.scitotenv.2015.05.057
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
57 of 64
Natri, H. M., Merilä, J., & Shikano, T. (2019). The evolution of sex determination associated with a chromosomal inversion. Nature Communica-
tions, 10(1), 145. https://doi.org/10.1038/s41467-018-08014-y
Nautiyal, H., Singal, S. K., Varun, & Sharma, A. (2011). Small hydropower for sustainable energy development in India. Renewable & Sustaina-
ble Energy Reviews, 15(4), 2021–2027. https://doi.org/10.1016/j.rser.2011.01.006
Neely, B. C., Pegg, M. A., & Mestl, G. E. (2009). Seasonal use distributions and migrations of blue sucker in the Middle Missouri River. Ecology
of Freshwater Fish, 18(3), 437–444. https://doi.org/10.1111/j.1600-0633.2009.00360.x
Nesshöver, C., Assmuth, T., Irvine, K. N., Rusch, G. M., Waylen, K. A., Delbaere, B., etal. (2017). The science, policy and practice of nature-based
solutions: An interdisciplinary perspective. Science of the Total Environment, 579, 1215–1227. https://doi.org/10.1016/j.scitotenv.2016.11.106
Netzband, A., Brils, J., Brauch, H. J., Liska, I., Miloradov, M., Nachtnebel, H. P., etal. (2007). Sediment management: An essential element of
river basin management plans. Journal of Soils and Sediments, 7(2), 117–132. https://doi.org/10.1065/jss2007.02.212
Nguyen, R. M., & Crocker, C. E. (2006). The effects of substrate composition on foraging behavior and growth rate of larval green sturgeon,
Acipenser medirostris. Environmental Biology of Fishes, 76(2), 129–138. https://doi.org/10.1007/s10641-006-9002-y
Nichols, G. (2009). Sedimentology and stratigraphy. John Wiley & Sons. Retrieved from https://www.wiley.com/go/nicholssedimentology
Nieminen, E., Hyytiäinen, K., & Lindroos, M. (2016). Economic and policy considerations regarding hydropower and migratory fish. Fish and
Fisheries, 18(1), 54–78. https://doi.org/10.1111/faf.12167
Nixon, S. W. (2003). Replacing the Nile: Are anthropogenic nutrients providing the fertility once brought to the Mediterranean by a great river?
Ambio, 32(1), 30–39. https://doi.org/10.1579/0044-7447-32.1.30
Noonan, M. J., Grant, J. W. A., & Jackson, C. D. (2012). A quantitative assessment of fish passage efficiency. Fish and Fisheries, 13(4), 450–464.
https://doi.org/10.1111/j.1467-2979.2011.00445.x
Nunes, D. M. F., Magalhães, A. L. B., Weber, A. A., Gomes, R. Z., Normando, F. T., Santiago, K. B., etal. (2015). Influence of a large dam
and importance of an undammed tributary on the reproductive ecology of the threatened fish matrinxã Brycon orthotaenia Günther, 1864
(Characiformes: Bryconidae) in southeastern Brazil. Neotropical Ichthyology, 13(2), 317–324. https://doi.org/10.1590/1982-0224-20140084
O'Connor, J. E., Duda, J. J., & Grant, G. E. (2015). 1000 dams down and counting. Science, 348(6234), 496–497. https://doi.org/10.1126/science.
aaa9204
Oczkowski, A., & Nixon, S. (2008). Increasing nutrient concentrations and the rise and fall of a coastal fishery: A review of data from the Nile
delta. Estuarine, Coastal and Shelf Science, 77(3), 309–319. https://doi.org/10.1016/j.ecss.2007.11.028
Oczkowski, A. J., Nixon, W. S., Granger, S. L., El-Sayed, A. F. M., & McKinney, R. A. (2009). Anthropogenic enhancement of Egypt’s Mediter-
ranean fishery. Proceedings of the National Academy of Sciences, 106(5), 1364–1367. https://doi.org/10.1073/pnas.0812568106
O'Mara, K., Venarsky, M., Stewart-Koster, B., McGregor, G. B., Schulz, C., Kainz, M., etal. (2021). Connectivity of fish communities in a
tropical floodplain river system and predicted impacts of potential new dams. Science of the Total Environment, 788, 147785. https://doi.
org/10.1016/j.scitotenv.2021.147785
O’Reilly, C. M., Sharma, S., Gray, D. K., Hampton, S. E., Read, J. S., Rowley, R. J., etal. (2015). Rapid and highly variable warming of lake
surface waters around the globe. Geophysical Research Letters, 42(24), 10773–10781. https://doi.org/10.1002/2015GL066235
Orr, S., Pittock, J., Chapagain, A., & Dumaresq, D. (2012). Dams on the Mekong River: Lost fish protein and the implications for land and water
resources. Global Environmental Change, 22(4), 925–932. https://doi.org/10.1016/j.gloenvcha.2012.06.002
Ortega-Recalde, O., Goikoetxea, A., Hore, T. A., Todd, E. V., & Gemmell, N. J. (2020). The genetics and epigenetics of sex change in fish. Annual
Review Animal Biosciences, 8(1), 47–69. https://doi.org/10.1146/annurev-animal-021419-083634
Ostberg, C. O., & Chase, D. M. (2022). Ontogeny of eDNA shedding during early development in Chinook Salmon (Oncorhynchus tshawytscha).
Environmental DNA, 4(2), 339–348. https://doi.org/10.1002/edn3.258
Ou, Y., Li, R., Hodges, B. R., Feng, J., & Pu, X. (2016). Impact of temperature on the dissipation process of supersaturated total dissolved gas in
flowing water. Fresenius Environmental Bulletin, 25(6), 1927–1934.
Paillex, A., Dolédec, S., Castella, E., Mérigoux, S., & Aldridge, D. C. (2013). Functional diversity in a large river floodplain: Anticipating the
response of native and alien macroinvertebrates to the restoration of hydrological connectivity. Journal of Applied Ecology, 50(1), 97–106.
https://doi.org/10.1111/1365-2664.12018
Paiva, M. P., Petrere, J. R., Petenate, A. J., & Nepomuceno, F. H. (1994). Relationship between the number of predatory fish species and fish yield
in large north-eastern Brazilian reservoirs. Rehabilitation of freshwater fisheries, 1, 120–129.
Palmer, M., & Ruhi, A. (2019). Linkages between flow regime, biota, and ecosystem processes: Implications for river restoration. Science,
365(6459), eaaw2087. https://doi.org/10.1126/science.aaw2087
Pander, J., Mueller, M., & Geist, J. (2015). Succession of fish diversity after reconnecting a large floodplain to the upper Danube River. Ecolog-
ical Engineering, 75, 41–50. https://doi.org/10.1016/j.ecoleng.2014.11.011
Pankhurst, N. W., & King, H. R. (2010). Temperature and salmonid reproduction: Implications for aquaculture. Journal of Fish Biology, 76(1),
69–85. https://doi.org/10.1111/j.1095-8649.2009.02484.x
Paragamian, V. L., Kruse, G., & Wakkinen, V. (2001). Spawning habitat of Kootenai River white sturgeon, post-Libby Dam. North American
Journal of Fisheries Management, 21(1), 22–33. https://doi.org/10.1577/1548-8675(2001)021<0022:SHOKRW>2.0.CO;2
Park, Y. S., Chang, J., Lek, S., Cao, W., & Brosse, S. (2003). Conservation strategies for endemic fish species threatened by the Three Gorges
Dam. Conservation Biology, 17(6), 1748–1758. https://doi.org/10.1111/j.1523-1739.2003.00430.x
Pavlov, D. S., & Skorobogatov, M. A. (2014). Chapter 5: Fish passages in Russia. In Fish migrations in regulated rivers (pp.154–202). Russian
Academy of Sciences, KMK Scientific Press.
Pedersen, M. I., Jepsen, N., Aarestrup, K., Koed, A., Pedersen, S., & Økland, F. (2012). Loss of European silver eel passing a hydropower station.
Journal of Applied Ichthyology, 28(2), 189–193. https://doi.org/10.1111/j.1439-0426.2011.01913.x
Peng, Q., Liao, W., Li, C., & Yu, X. (2012). Impacts of four major Chinese carps’ natural reproduction in the middle reaches of Changjiang River
by three Gorges project since the impoundment. Journal of Sichuang University (Engineering Science Edition)(S2), 228–232. https://doi.
org/10.15961/j.jsuese.2012.s2.041
Peng, Y., Hong, Y., Liu, D., & Chen, Q. (2022). Quantitative evaluation of fish alternative habitats in the Luosuo River. Water Resources Protec-
tion, 38(2), 190–196. https://doi.org/10.3880/j.issn.1004-6933.2022.02.026
Peng, Y., Lin, Y., Zeng, C., Zha, W., Mao, F., Chen, Q., etal. (2022). Improved model for predicting total dissolved gas generation with the
residence time of the water in the stilling phase. Frontiers in Environmental Science, 9, 770187. https://doi.org/10.3389/fenvs.2021.770187
Péril, E. E. (2004). White sturgeon. The Species at Risk Public Registry.
Perna, C. N., Cappo, M., Pusey, B. J., Burrows, D. W., & Pearson, R. G. (2012). Removal of aquatic weeds greatly enhances fish community rich-
ness and diversity: An example from the Burdekin River floodplain, tropical Australia. River Research and Applications, 28(8), 1093–1104.
https://doi.org/10.1002/rra.1505
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
58 of 64
Peters, D. L., Monk, W. A., & Baird, D. J. (2014). Cold-regions Hydrological Indicators of Change (CHIC) for ecological flow needs assessment.
Hydrological Sciences Journal, 59(3–4), 502–516. https://doi.org/10.1080/02626667.2013.835489
Peterson, M. S., Slack, W. T., Waggy, G. L., Finley, J., Woodley, C. M., & Partyka, M. L. (2006). Foraging in non-native environments: Compar-
ison of Nile Tilapia and three co-occurring native centrarchids in invaded coastal Mississippi watersheds. Environmental Biology of Fishes,
76(2–4), 283–301. https://doi.org/10.1007/s10641-006-9033-4
Petts, G. E., & Gurnell, A. M. (2005). Dams and geomorphology: Research progress and future directions. Geomorphology, 71(1–2), 27–47.
https://doi.org/10.1016/j.geomorph.2004.02.015
Phillips, R., & Rab, P. (2001). Chromosome evolution in the Salmonidae (Pisces): An update. Biological Reviews, 76(1), 1–25. https://doi.
org/10.1017/s1464793100005613
Pieters, R., & Lawrence, G. A. (2012). Plunging inflows and the summer photic zone in reservoirs. Water Quality Research Journal of Canada,
47(3–4), 268–275. https://doi.org/10.2166/wqrjc.2012.143
Pitlick, J., & Wilcock, P. (2001). Relations between streamflow, sediment transport, and aquatic habitat in regulated rivers. Geomorphic Processes
and Riverine Habitat, 4, 185–198. https://doi.org/10.1029/WS004p0185
Pleizier, N. K., Nelson, C., Cooke, S. J., & Brauner, C. J. (2020). Understanding gas bubble trauma in an era of hydropower expansion: How
do fish compensate at depth? Canadian Journal of Fisheries and Aquatic Sciences, 77(3), 556–563. https://doi.org/10.1139/cjfas-2019-0243
Poff, N. L. (2018). Beyond the natural flow regime? Broadening the hydro-ecological foundation to meet environmental flows challenges in a
non-stationary world. Freshwater Biology, 63(8), 1011–1021. https://doi.org/10.1111/fwb.13038
Poff, N. L., Allan, J. D., Bain, M. B., Karr, J. R., Prestegaard, K. L., Richter, B. D., etal. (1997). The natural flow regime. BioScience, 47(11),
769–784. https://doi.org/10.2307/1313099
Poff, N. L., Olden, J. D., Merritt, D. M., & Pepin, D. M. (2007). Homogenization of regional river dynamics by dams and global biodiversity
implications. Proceedings of the National Academy of Sciences of the United States of America, 104(14), 5732–5737. https://doi.org/10.1073/
pnas.0609812104
Pompeu, P.D., & Martinez, C. B. (2007). Efficiency and selectivity of atrap and truck fish passage system in Brazil. Neotropical Ichthyology,
5(2), 169–176. https://doi.org/10.1590/S1679-62252007000200011
Ponomareva, E., Geraskin, P., Sorokina, M., Kovalevа, A., & Grigoriev, V. (2020). Features of the reproductive system development in
the installations of closed water supply system. In E3S web of conferences (Vol. 175,p. 02012). EDP Sciences. https://doi.org/10.1051/
e3sconf/202017502012
Popov, I. (2017). Habitats for the atlantic sturgeons in Russia. Aquatic Conservation: Marine and Freshwater Ecosystems, 27(3), 717–730. https://
doi.org/10.1002/aqc.2716
Pörtner, H. O., & Farrell, A. P. (2008). Physiology and climate change. Science, 322(5902), 690–692. https://doi.org/10.1126/science.1163156
Potter, E. C. E., & Crozier, W. W. (2000). A perspective on the marine survival of Atlantic salmon. In The ocean life of Atlantic salmon: Environ-
mental and biological factors influencing survival (pp.19–36).
Pracheil, B. M., DeRolph, C. R., Schramm, M. P., & Bevelhimer, M. S. (2016). A fish-eye view of riverine hydropower systems: The current
understanding of the biological response to turbine passage. Reviews in Fish Biology and Fisheries, 26(2), 153–167. https://doi.org/10.1007/
s11160-015-9416-8
Pracheil, B. M., McIntyre, P.B., & Lyons, J. D. (2013). Enhancing conservation of large-river biodiversity by accounting for tributaries. Frontiers
in Ecology and the Environment, 11(3), 124–128. https://doi.org/10.1890/120179
Pracheil, B. M., Pegg, M., & Mestl, G. (2009). Tributaries influence recruitment of fish in large rivers. Ecology of Freshwater Fish, 18(4),
603–609. https://doi.org/10.1111/j.1600-0633.2009.00376.x
Prada, A. F., George, A. E., Stahlschmidt, B. H., Jackson, P.R., Chapman, D. C., & Tinoco, R. O. (2020). Influence of turbulence and in-stream
structures on the transport and survival of grass carp eggs and larvae at various developmental stages. Aquatic Sciences, 82(1), 1–16. https://
doi.org/10.1007/s00027-019-0689-1
Prats, J., Val, R., Armengol, J., & Dolz, J. (2010). Temporal variability in the thermal regime of the lower Ebro River (Spain) and alteration due
to anthropogenic factors. Journal of Hydrology, 387(1–2), 105–118. https://doi.org/10.1016/J.JHYDROL.2010.04.002
Prokešová, M., Gebauer, T., Matoušek, J., Lundová, K., Čejka, J., Zusková, E., & Stejskal, V. (2020). Effect of temperature and oxygen regime
on growth and physiology of juvenile Salvelinus fontinalis× Salvelinus alpinus hybrids. Aquaculture, 522, e735119. https://doi.org/10.1016/j.
aquaculture.2020.735119
Prowse, T. D. (2001). River-ice ecology. I: Hydrologic, geomorphic, and water-quality aspects. Journal of Cold Regions Engineering, 15(1),
1–16. https://doi.org/10.1061/(ASCE)0887-381X(2001)15:1(1)
Pulg, U., Barlaup, B. T., Sternecker, K., Trepl, L., & Unfer, G. (2008). Restoration of spawning habitats of brown trout (Salmo trutta) in a regu-
lated chalk stream. River Research and Applications, 29(2), 172–182. https://doi.org/10.1002/rra.1594
Pulg, U., Vollset, K. W., Velle, G., & Stranzl, S. (2016). First observations of saturopeaking: Characteristics and implications. Science of the Total
Environment, 573(23), 1615–1621. https://doi.org/10.1016/j.scitotenv.2016.09.143
Qin, S., Leng, X., Luo, J., Du, H., Liu, Z., Qiao, X., etal. (2020). Growth and physiological characteristics of juvenile Chinese sturgeon (Acip-
enser sinensis) during adaptation to seawater. Aquaculture Research, 51(9), 3813–3821. https://doi.org/10.1111/are.14729
Qu, L., Li, R., Li, J., Li, K., & Wang, L. (2011). Experimental study on total dissolved gas supersaturation in water. Water Science and Engineer-
ing, 4(4), 396–404. https://doi.org/10.3882/j.issn.1674-2370.2011.04.004
Quaranta, E., Katopodis, C., & Comoglio, C. (2019). Effects of bed slope on the flow field of vertical slot fishways. River Research and Applica-
tions, 35(6), 656–668. https://doi.org/10.1002/rra.3428
Ralph, M., Suresh, S., Sebastian, H., Evelyn, C., Stefanie, M., Christian, S., etal. (2009). Gene expression profiling to characterize sediment toxicity–a
pilot study using Caenorhabditis elegans whole genome microarrays. BMC Genomics, 10(1), 160. https://doi.org/10.1186/1471-2164-10-160
Raymond, C. M., Frantzeskaki, N., Kabisch, N., Berry, P., Breil, M., Nita, M. R., etal. (2017). A framework for assessing and implementing the
co-benefits of nature-based solutions in urban areas. Environmental Science & Policy, 77, 15–24. https://doi.org/10.1016/j.envsci.2017.07.008
Rechisky, E. L., Welch, D. W., Porter, A. D., Jacobs-Scott, M. C., & Winchell, P.M. (2013). Influence of multiple dam passage on survival
of juvenile Chinook salmon in the Columbia River estuary and coastal ocean. Proceedings of the National Academy of Sciences, 110(17),
6883–6888. https://doi.org/10.1073/pnas.1219910110
Reid, A. J., Carlson, A. K., Creed, I. F., Eliason, E. J., Gell, P.A., Johnson, P.T., etal. (2019). Emerging threats and persistent conservation
challenges for freshwater biodiversity. Biological Reviews, 94(3), 849–873. https://doi.org/10.1111/brv.12480
Reinfelds, I., Walsh, C., van Der Meulen, D., Growns, I., & Gray, C. (2013). Magnitude, frequency and duration of instream flows to stimulate
and facilitate catadromous fish migrations: Australian bass (Macquaria novemaculeata Perciformes, Percichthyidae). River Research and
Applications, 29(4), 512–527. https://doi.org/10.1002/rra.1611
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
59 of 64
Ren, L., Song, C., Wu, W., Guo, M., & Zhou, X. (2020). Reservoir effects on the variations of the water temperature in the upper Yellow
River, China, using principal component analysis. Journal of Environmental Management, 262, 110339. https://doi.org/10.1016/j.
jenvman.2020.110339
Renaud, C. B. (2011). Lampreys of the world. An annotated and illustrated catalogue of lamprey species known to date. FAO Species Catalogue
for Fishery Purposes, 5, 109. Retrieved from https://www.fao.org/3/i2335e/i2335e00.htm
Ribas, L., Liew, W. C., Díaz, N., Sreenivasan, R., Orbán, L., & Piferrer, F. (2017). Heat-induced masculinization in domesticated zebrafish is
family-specific and yields a set of different gonadal transcriptomes. Proceedings of the National Academy of Sciences, 114(6), E941–E950.
https://doi.org/10.1073/pnas.1609411114
Richards, A. (1982). Egypt’s agricultural development, 1800–1980. Technical and social change. The Journal of Economic History, 43(3),
767–768. https://doi.org/10.1017/S0022050700030539
Richter, B. D., Baumgartner, J. V., Braun, D. P., & Powell, J. (1998). A spatial assessment of hydrologic alteration within a river network. Regulated
Rivers: Research & Management, 14(4), 329–340. https://doi.org/10.1002/(SICI)1099-1646(199807/08)14:4<329::AID-RRR505>3.0.CO;2-E
Richter, B. D., Baumgartner, J. V., Powell, J., & Braun, D. P. (1996). A method for assessing hydrologic alteration within ecosystems. Conserva-
tion Biology, 10(4), 1163–1174. https://doi.org/10.1046/j.1523-1739.1996.10041163.x
Richter, B. D., Baumgartner, J. V., Wigington, R., & Braun, D. P. (1997). How much water does a river need? Freshwater Biology, 37(1),
231–249. https://doi.org/10.1046/j.1365-2427.1997.00153.x
Righton, D., Piper, A., Aarestrup, K., Amilhat, E., Belpaire, C., Casselman, J., etal. (2021). Important questions to progress science and sustain-
able management of anguillid eels. Fish and Fisheries, 22(4), 762–788. https://doi.org/10.1111/faf.12549
Rijnsdorp, A. D., Peck, M. A., Engelhard, G. H., Möllmann, C., & Pinnegar, J. K. (2009). Resolving the effect of climate change on fish popula-
tions. ICES Journal of Marine Science, 66(7), 1570–1583. https://doi.org/10.1093/ICESJMS/FSP056
Rivinoja, P., McKinnell, S., & Lundqvist, H. (2001). Hindrances to upstream migration of Atlantic salmon (Salmo salar) in a northern Swed-
ish river caused by a hydroelectric power-station. Regulated Rivers: Research & Management: An International Journal Devoted to River
Research and Management, 17(2), 101–115. https://doi.org/10.1002/rrr.607
Roberts, C. G., Bašić, T., Britton, J. R., Rice, S., & Pledger, A. G. (2020). Quantifying the habitat and zoogeomorphic capabilities of spawning
European barbel Barbus barbus, a lithophilous cyprinid. River Research and Applications, 36(2), 259–279. https://doi.org/10.1002/rra.3573
Rodeles, A. A., Galicia, D., & Miranda, R. (2017). Recommendations for monitoring freshwater fishes in river restoration plans: A wasted oppor-
tunity for assessing impact. Aquatic Conservation: Marine and Freshwater Ecosystems, 27(4), 880–885. https://doi.org/10.1002/aqc.2753
Rowland, D. E., & Jensen, J. O. T. (1988). The effect of gas supersaturated water on Juvenile Chinook (Oncorhynchus tshawytscha) held in cages
in the nechako river, British Columbia. Department of Fisheries and Oceans, Biological Sciences Branch.
Rytwinski, T., Kelly, L. A., Donaldson, L. A., Taylor, J. J., Smith, A., Drake, D. A. R., etal. (2021). What evidence exists for evaluating the
effectiveness of conservation-oriented captive breeding and release programs for imperilled freshwater fishes and mussels? Canadian Journal
of Fisheries and Aquatic Sciences, 78(9), 1332–1346. https://doi.org/10.1139/cjfas-2020-0331
Rzóska, J. (1976). A controversy reviewed. Nature, 261(5560), 444–445. https://doi.org/10.1038/261444a0
Saadatpour, M., Javaheri, S., Afshar, A., & Solis, S. S. (2021). Optimization of selective withdrawal systems in hydropower reservoir considering
water quality and quantity aspects. Expert Systems with Applications, 184, 115474. https://doi.org/10.1016/j.eswa.2021.115474
Sabo, J., Ruhi, A., Holtgrieve, G., Elliott, V., Arias, M., Ngor, P.B., etal. (2017). Designing river flows to improve food security futures in the
Lower Mekong Basin. Science, 358(6368), eaao1053. https://doi.org/10.1126/science.aao1053
Saha, T. K., Pal, S., & Sarda, R. (2022). Impact of river flow modification on wetland hydrological and morphological characters. Environmental
Science and Pollution Research, 29(50), 75769–75789. https://doi.org/10.1007/s11356-022-21072-6
Santos, J. M., Amaral, S. D., & Pádua, J. (2021). Can fish lifts aid upstream dispersal of the invasive red swamp crayfish (Procambarus clarkii)
past high-head hydropower plants? River Research and Applications, 38(8), 1519–1523. https://doi.org/10.1002/rra.3856
Sawyer, A. H., Cardenas, M. B., Bomar, A., & Mackey, M. (2009). Impact of dam operations on hyporheic exchange in the riparian zone of a
regulated river. Hydrological Processes, 23(15), 2129–2137. https://doi.org/10.1002/hyp.7324
Schilt, C. R. (2007). Developing fish passage and protection at hydropower dams. Applied Animal Behaviour Science, 104(3–4), 295–325. https://
doi.org/10.1016/j.applanim.2006.09.004
Schisler, G. J., Bergersen, E. P., & Walker, P.G. (2000). Effects of multiple stressors on morbidity and mortality of fingerling Rainbow Trout
infected with Myxobolus cerebralis. Transactions of the American Fisheries Society, 129(3), 859–865. https://doi.org/10.1577/1548-8659(20
00)129<0859:EOMSOM>2.3.CO;2
Schlueter, U. (1971). Ueberlegungen zum naturnahen Ausbau von Wasseerlaeufen. Landschaft und Stadt, 9(2), 72–83.
Schmidt, J. (1923). Breeding places and migrations of the eel. Nature, 111(2776), 51–54. https://doi.org/10.1038/111051a0
Schmitt, R. J. P., Bizzi, S., Castelletti, A., Opperman, J. J., & Kondolf, G. M. (2019). Planning dam portfolios for low sediment trapping shows
limits for sustainable hydropower in the Mekong. Science Advances, 5(10), eaaw2175. https://doi.org/10.1126/sciadv.aaw2175
Secor, D., Arefjev, V., Nikolaev, A., & Sharov, A. (2000). Restoration of sturgeons: Lessons from the Caspian Sea sturgeon ranching programme.
Fish and Fisheries, 1(3), 215–230. https://doi.org/10.1111/j.1467-2979.2000.00021.x
Servili, A., Canario, A. V. M., Mouchel, O., & Muñoz-Cueto, J. A. (2020). Climate change impacts on fish reproduction are mediated at multiple
levels of the brain-pituitary-gonad axis. General and Comparative Endocrinology, 291, 113439. https://doi.org/10.1016/j.ygcen.2020.113439
Shaffer, J. A., Munsch, S., & Juanes, F. (2018). Functional diversity responses of a nearshore fish community to restoration driven by large-scale
dam removal. Estuarine, Coastal and Shelf Science, 213, 245–252. https://doi.org/10.1016/j.ecss.2018.08.030
Shahabi, M., Ghomeshi, M., Ahadiyan, J., Mohammadian, T., & Katopodis, C. (2021). Do fishways stress fish? Assessment of physiological and
hydraulic parameters of rainbow trout navigating a novel W-weir fishway. Ecological Engineering, 169, 106330. https://doi.org/10.1016/j.
ecoleng.2021.106330
Shen, Y., Yang, N., Liu, Z., Chen, Q., & Li, Y. (2020). Phylogenetic perspective on the relationships and evolutionary history of the Acipenseri-
formes. Genomics, 112(5), 3511–3517. https://doi.org/10.1016/j.ygeno.2020.02.017
Shrimpton, J. M., Randall, D. J., & Fidler, L. E. (1990a). Assessing the effects of positive buoyancy on rainbow trout (Oncorhynchus mykiss) held
in gas supersaturated water. Canadian Journal of Zoology, 68(5), 969–973. https://doi.org/10.1139/z90-139
Shrimpton, J. M., Randall, D. J., & Fidler, L. E. (1990b). Factors affecting swim bladder volume in rainbow trout (Oncorhynchus mykiss) held in
gas supersaturated water. Canadian Journal of Zoology, 68(5), 962–968. https://doi.org/10.1139/z90-138
Siddique, M. A. M., Psenicka, M., Cosson, J., Dzyuba, B., Rodina, M., Golpour, A., & Linhart, O. (2016). Egg stickiness in artificial reproduction
of sturgeon: An overview. Reviews in Aquaculture, 8(1), 18–29. https://doi.org/10.1111/raq.12070
Sieiro, P., Otero, J., & Aubourg, S. P. (2020). Biochemical composition and energy strategy along the reproductive cycle of female Octopus
vulgaris in Galician waters (NW Spain). Frontiers in Physiology, 11, 760. https://doi.org/10.3389/fphys.2020.00760
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
60 of 64
Skoglund, H., Einum, S., Forseth, T., & Barlaup, B. T. (2011). Phenotypic plasticity in physiological status at emergence from nests as a
response to temperature in Atlantic salmon (Salmo salar). Canadian Journal of Fisheries and Aquatic Sciences, 68(8), 1470–1479. https://
doi.org/10.1139/f2011-056
Słowik, M., Dezső, J., Marciniak, A., Tóth, G., & Kovács, J. (2018). Evolution of river planforms downstream of dams: Effect of dam construction
or earlier human-induced changes? Earth Surface Processes and Landforms, 43(10), 2045–2063. https://doi.org/10.1002/esp.4371
Smith, D. P., Kortman, S. R., Caudillo, A. M., Kwan-Davis, R. L., Wandke, J. J., Klein, J. W., etal. (2020). Controls on large boulder mobility
in an ‘auto-naturalized’ constructed step-pool river: San Clemente Reroute and dam removal project, Carmel River, California, USA. Earth
Surface Processes and Landforms, 45(9), 1990–2003. https://doi.org/10.1002/esp.4860
Sokolov, D. I., Erina, O. N., Tereshina, M. A., & Puklakov, V. V. (2020). Impact of Mozhaysk dam on the Moscow river sediment transport.
Geography, Environment, Sustainability, 13(4), 24–31. https://doi.org/10.24057/2071-9388-2019-150
Soleimani, S., Bozorg-Haddad, O., Saadatpour, M., & Loáiciga, H. A. (2019). Simulating thermal stratification and modeling outlet water temper-
ature in reservoirs with a data-mining method. Journal of Water Supply: Research & Technology Aqua, 68(1), 7–19. https://doi.org/10.2166/
aqua.2018.036
Song, H., Xu, S., Luo, K., Hu, M., Luan, S., Shao, H., etal. (2022). Estimation of genetic parameters for growth and egg related traits in Russian
sturgeon (Acipenser gueldenstaedtii). Aquaculture, 546, 737299. https://doi.org/10.1016/j.aquaculture.2021.737299
Song, Y. A., Kirkwood, N., Maksimovic, C., Zhen, X. D., O'Connor, D., Jin, Y. L., & Hou, D. Y. (2019). Nature based solutions for contaminated
land remediation and brownfield redevelopment in cities: A review. Science of the Total Environment, 663, 568–579. https://doi.org/10.1016/j.
scitotenv.2019.01.347
Soukhaphon, A., Baird, I., & Hogan, Z. (2021). The impacts of hydropower dams in the Mekong River Basin: A review. Water, 13(3), 265. https://
doi.org/10.3390/w13030265
Steffensen, S. M., Thiem, J. D., Stamplecoskie, K. M., Binder, T. R., Hatry, C., Langlois-Anderson, N., & Cooke, S. J. (2013). Biological effec-
tiveness of an inexpensive nature-like fishway for passage of warmwater fish in a small Ontario stream. Ecology of Freshwater Fish, 22(3),
374–383. https://doi.org/10.1111/eff.12032
Stevaux, J. C., Martins, D. P., & Meurer, M. (2009). Changes in a large regulated tropical river: The Paraná River downstream from the Porto
Primavera dam, Brazil. Geomorphology, 113(3–4), 230–238. https://doi.org/10.1016/j.geomorph.2009.03.015
Stich, D. S., Kinnison, M. T., Kocik, J. F., & Zydlewski, J. D. (2015). Initiation of migration and movement rates of Atlantic salmon smolts in
fresh water. Canadian Journal of Fisheries and Aquatic Sciences, 72(9), 1339–1351. https://doi.org/10.1139/cjfas-2014-0570
Stoffels, R. J., Humphries, P., Bond, N. R., & Price, A. E. (2022). Fragmentation of lateral connectivity and fish population dynamics in large
rivers. Fish and Fisheries, 23(3), 680–696. https://doi.org/10.1111/faf.12641
Stoffers, T., Buijse, A. D., Geerling, G. W., Jans, L. H., Schoor, M. M., Poos, J. J., etal. (2022). Freshwater fish biodiversity restoration in flood-
plain rivers requires connectivity and habitat heterogeneity at multiple spatial scales. Science of the Total Environment, 838(4), 156509. https://
doi.org/10.1016/j.scitotenv.2022.156509
Stoffers, T., Collas, F.P.L., Buijse, A. D., Geerling, G. W., Jans, L. H., van Kessel, N., etal. (2021). 30 years of large river restoration: How long
do restored floodplain channels remain suitable for targeted rheophilic fishes in the lower river Rhine. Science of the Total Environment, 755,
142931. https://doi.org/10.1016/j.scitotenv.2020.142931
Stone, R. (2008). Three Gorges dam: Into the unknown. Science, 321(5889), 628–632. https://doi.org/10.1126/science.321.5889.628
Su, G. H., Logez, M., Xu, J., Tao, S. L., Villéger, S., & Brosse, S. (2021). Human impacts on global freshwater fish biodiversity. Science,
371(6531), 835–838. https://doi.org/10.1126/science.abd3369
Su, X., Nilsson, C., Pilotto, F., Liu, S., Shi, S., & Zeng, B. (2017). Soil erosion and deposition in the new shorelines of the Three Gorges Reservoir.
Science of the Total Environment, 599, 1485–1492. https://doi.org/10.1016/j.scitotenv.2017.05.001
Suen, J., & Eheart, J. W. (2006). Reservoir management to balance ecosystem and human needs: Incorporating the paradigm of the ecological f low
regime: Reservoir management to balance ecosystem. Water Resources Research, 42(3), W03417. https://doi.org/10.1029/2005WR004314
Sun, Z., & Wang, D. (2020). Practical innovation of nature reserve management under hydropower development: Case of the national rare and
endemic fishes nature reserve in upper reaches of Yangtze River. Journal of Yangtze River Scientific Research Institute, 37(11), 1–7. https://
doi.org/10.11988/ckyyb.20201065
Sykes, G. E., Johnson, C. J., & Shrimpton, J. M. (2009). Temperature and flow effects on migration timing of Chinook salmon smolts. Transac-
tions of the American Fisheries Society, 138(6), 1252–1265. https://doi.org/10.1577/T08-180.1
Syvitski, J., Cohen, S., Miara, A., & Best, J. (2019). River temperature and the thermal-dynamic transport of sediment. Global and Planetary
Change, 178, 168–183. https://doi.org/10.1016/j.gloplacha.2019.04.011
Syvitski, J. P., Cohen, S., Kettner, A. J., & Brakenridge, G. R. (2014). How important and different are tropical rivers? An overview. Geomorphol-
ogy, 227, 5–17. https://doi.org/10.1016/j.geomorph.2014.02.029
Tang, C., Yan, Q., Li, W., Yang, X., & Zhang, S. (2022). Impact of dam construction on the spawning grounds of the four major Chinese carps in
the Three Gorges Reservoir. Journal of Hydrology, 609, 127694. https://doi.org/10.1016/j.jhydrol.2022.127694
Tang, L., Mo, K., Zhang, J., Wang, J., Chen, Q., He, S., etal. (2021). Removing tributary low-head dams can compensate for fish habitat losses
in dammed rivers. Journal of Hydrology, 598, 126204. https://doi.org/10.1016/j.jhydrol.2021.126204
Taniwaki, R. H., Borghi, T. C., Magrin, A. G. E., Calijuri, M. D. C., Bottino, F., & Moschini-Carlos, V. (2013). Structure and dynamics of the
community of periphytic algae in a subtropical reservoir (state of São Paulo, Brazil). Acta Botanica Brasilica, 27(3), 551–559. https://doi.
org/10.1590/S0102-33062013000300013
Tao, J., Qiao, Y., Tan, X., & Chang, J. (2009). Species identification of Chinese sturgeon using acoustic descriptors and ascertaining their
spatial distribution in the spawning ground of Gezhouba Dam. Chinese Science Bulletin, 54(21), 3972–3980. https://doi.org/10.1007/
s11434-009-0557-9
Tao, Y., Wang, Y., Rhoads, B., Wang, D., Ni, L., & Wu, J. (2020). Quantifying the impacts of the three gorges reservoir on water temperature in
the middle reach of the Yangtze River. Journal of Hydrology, 582, 124476–124483. https://doi.org/10.1016/j.jhydrol.2019.124476
Terra, B. F., Santos, A. B. I., & Araújo, F. G. (2010). Fish assemblage in a dammed tropical river: An analysis along the longitudinal and temporal
gradients from river to reservoir. Neotropical Ichthyology, 8(3), 599–606. https://doi.org/10.1590/S1679-62252010000300004
Tesch, F. (2003). The eel. Blackwell Science Ltd. https://doi.org/10.1002/9780470995389.ch2
Thompson, M. S. A., Brooks, S. J., Sayer, C. D., Woodward, G., Axmacher, J. C., Perkins, D. M., & Gray, C. (2018). Large woody debris “rewild-
ing” rapidly restores biodiversity in riverine food webs. Journal of Applied Ecology, 55(2), 895–904. https://doi.org/10.1111/1365-2664.13013
Thoms, M. C. (2003). Floodplain–river ecosystems: Lateral connections and the implications of human interference. Geomorphology, 56(3–4),
335–349. https://doi.org/10.1016/S0169-555X(03)00160-0
Thorncraft, G., & Harris, J. H. (2000). Fish passage and fishways in new south wales: A status report. Technical report 1/2000. Cooperative
Research Centre for Freshwater Ecology.
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
61 of 64
Tiffan, K. F., Haskell, C. A., & Kock, T. J. (2010). Quantifying the behavioral response of spawning chum salmon to elevated discharges from
Bonneville Dam, Columbia River, USA. River Research and Applications, 26(2), 87–101. https://doi.org/10.1002/rra.1248
Timpe, K., & Kaplan, D. (2017). The changing hydrology of a dammed Amazon. Science Advances, 3(11), e1700611. https://doi.org/10.1126/
sciadv.1700611
Tomie, J., Cairns, D. K., Hobbs, R. S., Desjardins, M., Fletcher, G. L., & Courtenay, S. C. (2016). American eel (Anguilla rostrata) substrate
selection for daytime refuge and winter thermal sanctuary. Marine and Freshwater Research, 68(1), 95–105. https://doi.org/10.1071/MF15102
Topping, D. J., Rubin, D. M., & Vierra, L. E., Jr. (2000). Colorado river sediment transport: 1. Natural sediment supply limitation and the influ-
ence of Glen Canyon dam. Water Resources Research, 36(2), 515–542. https://doi.org/10.1029/1999WR900285
Toscano, B. J., & Griffen, B. D. (2013). Predator size interacts with habitat structure to determine the allometric scaling of the functional
response. Oikos, 122(3), 454–462. https://doi.org/10.1111/j.1600-0706.2012.20690.x
Tranmer, A. W., Benjankar, R., & Tonina, D. (2020). Post-wildfire riparian forest recovery processes along a regulated river corridor. Forest
Ecology and Management, 478, 118513. https://doi.org/10.1016/j.foreco.2020.118513
Triano, B., Kappenman, K. M., McMahon, T. E., Blank, M., Heim, K. C., Parker, A. E., etal. (2022). Attraction, entrance, and passage efficiency
of Arctic grayling, trout, and suckers at Denil fishways in the Big Hole River basin, Montana. Transactions of the American Fisheries Society,
151(4), 453–473. https://doi.org/10.1002/tafs.10362
Tsukamoto, K., Chow, S., Otake, T., Kurogi, H., Mochioka, N., Miller, M. J., etal. (2011). Oceanic spawning ecology of freshwater eels in the
western North Pacific. Nature Communications, 2(1), 1–9. https://doi.org/10.1038/ncomms1174
Tucker, D. W. (1959). A new solution to the Atlantic eel problem. Nature, 183(4660), 495–501. https://doi.org/10.1038/183495a0
Tummers, J. S., Winter, E., Silva, S., O’Brien, P., Jang, M. H., & Lucas, M. C. (2016). Evaluating the effectiveness of a Larinier super active
baffle fish pass for European river lamprey Lampetra fluviatilis before and after modification with wall-mounted studded tiles. Ecological
Engineering, 91, 183–194. https://doi.org/10.1016/j.ecoleng.2016.02.046
Tytell, E. D., & Lauder, G. V. (2004). The hydrodynamics of eel swimming: I. Wake structure. Journal of Experimental Biology, 207(11),
1825–1841. https://doi.org/10.1242/jeb.00968
van Puijenbroek, P.J. T. M., Buijse, A. D., Kraak, M. H. S., & Verdonschot, P.F. M. (2018). Species and river specific effects of river fragmenta-
tion on European anadromous fish species. River Research and Applications, 35(1), 68–77. https://doi.org/10.1002/rra.3386
van Vliet, M. T. H., Franssen, W. H. P., Yearsley, J. R., Ludwig, F., Haddeland, I., Lettenmaier, D. P., & Kabat, P. (2013a). Global river discharge
and water temperature under climate change. Global Environmental Change, 23(2), 450–464. https://doi.org/10.1016/j.gloenvcha.2012.11.002
van Vliet, M. T. H., Ludwig, F., & Kabat, P. (2013). Global streamflow and thermal habitats of freshwater fishes under climate change. Climatic
Change, 121(4), 739–754. https://doi.org/10.1007/s10584-013-0976-0
Vié, J., Hilton-Taylor, C., & Stuart, S. N. (2009). Wildlife in a changing world: An analysis of the 2008 IUCN red list of threatened species. IUCN.
Retrieved from https://www.iucn.org/content/wildlife-a-changing-world-analysis-2008-iucn-red-list-threatened-speciestm
Vilizzi, L. (2012). Abundance trends in floodplain fish larvae: The role of annual flow characteristics in the absence of overbank flooding. Funda-
mental and Applied Limnology, 181(3), 215–227. https://doi.org/10.1127/1863-9135/2012/0394
Vine, J. R., Holbrook, S. C., Post, W. C., & Peoples, B. K. (2019). Identifying environmental cues for Atlantic sturgeon and shortnose stur-
geon spawning migrations in the Savannah River. Transactions of the American Fisheries Society, 148(3), 671–681. https://doi.org/10.1002/
tafs.10163
Vitule, J., Freire, C., & Simberloff, D. (2009). Introduction of non-native freshwater fish can certainly be bad. Fish and Fisheries, 10(1), 98–108.
https://doi.org/10.1111/j.1467-2979.2008.00312.x
Wagner, B., Hauer, C., Schoder, A., & Habersack, H. (2015). A review of hydropower in Austria: Past, present and future development. Renewa-
ble and Sustainable Energy Reviews, 50, 304–314. https://doi.org/10.1016/j.rser.2015.04.169
Walter, R. C., & Merritts, D. J. (2008). Natural streams and the legacy of water-powered mills. Science, 319(5861), 299–304. https://doi.
org/10.1126/science.1151716
Wan, H., Li, J., Li, R., Feng, J., & Sun, Z. (2020). The optimal power generation operation of a hydropower station for improving fish shelter area
of low TDG level. Ecological Engineering, 147, 105749. https://doi.org/10.1016/j.ecoleng.2020.105749
Wan, H., Tan, Q., Li, R., Cai, Y., Shen, X., Yang, Z., etal. (2021). Incorporating fish tolerance to supersaturated total dissolved gas for generating
flood pulse discharge patterns based on a simulation-optimization approach. Water Resources Research, 57(9), e2021WR030167. https://doi.
org/10.1029/2021WR030167
Wang, H., Tao, J., & Chang, J. (2019). Endangered levels and conservation options evaluations for Chinese sturgeon, Acipenser sinensis Gary.
Resources and Environment in the Yangtze Basin, 28(9), 2100–2108.
Wang, J., Li, C., Duan, X., Chen, D., Feng, S., Luo, H., etal. (2014). Variation in the significant environmental factors affecting larval abundance
of four major Chinese carp species: Fish spawning response to the three Gorges dam. Freshwater Biology, 59(7), 1343–1360. https://doi.
org/10.1111/fwb.12348
Wang, J., Li, C., Duan, X., Luo, H., Feng, S., Peng, Q., & Liao, W. (2014). The relationship between thermal regime alteration and spawning
delay of the four major Chinese carps in the Yangtze River below the Three Gorges Dam. River Research and Applications, 30(8), 987–1001.
https://doi.org/10.1002/rra.2691
Wang, L. (2016). Impacts on natural stock genetic diversity and fish composition by Sebastes schlegelii enhancement in Lidao bay, (Doctoral
dissertation). Institutional Repository of Key Laboratory of Experimental Marine Biology, Institute of Oceanology, Chinese Academy of
Sciences. Retrieved from http://ir.qdio.ac.cn/handle/337002/116986
Wang, L., & Huang, Z. (2020). What is actually the main cause for the survival crisis of Chinese sturgeon? Journal of Lake Science, 32(4),
924–940. https://doi.org/10.18307/2020.0403
Wang, N., Teletchea, F., Kestemont, P., Milla, S., & Fontaine, P. (2010). Photothermal control of the reproductive cycle in temperate fishes.
Reviews in Aquaculture, 2(4), 209–222. https://doi.org/10.1111/J.1753-5131.2010.01037.X
Wang, W., Tian, S., Meng, Z., & Lai, R. (2012). Evolution processes of river pattern in lower Yellow River after commissioning of Xiaolangdi
reservoir. Journal of Sediment Research, 1, 23–31.
Wang, X., Dong, Z., Ai, X., Dong, X., & Li, Y. (2020). Multi-objective model and decision-making method for coordinating the ecological bene-
fits of the Three Gorger Reservoir. Journal of Cleaner Production, 270, 122066. https://doi.org/10.1016/j.jclepro.2020.122066
Wang, Y., An, R., Li, Y., & Li, K. (2017). Swimming performance of rock carp Procypris rabaudi and Prenant’s Schizothoracin Schizothorax
prenanti acclimated to total dissolved gas supersaturated water. North American Journal of Fisheries Management, 37(6), 1183–1190. https://
doi.org/10.1080/02755947.2017.1353558
Wang, Y., Dai, H., Wang, B., & Dai, L. (2013). Study of the eco-scheduling for optimization Chinese sturgeon spawning habitats. Shuili Xuebao,
44(3), 319–326. https://doi.org/10.13243/j.cnki.slxb.2013.03.007
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
62 of 64
Wang, Y., Lei, X., Wen, X., Fang, G., Tan, Q., Tian, Y., etal. (2019). Effects of damming and climatic change on the eco-hydrological system: A
case study in the Yalong River, southwest China. Ecological Indicators, 105, 663–674. https://doi.org/10.1016/j.ecolind.2018.07.039
Wang, Y., Li, K., Li, J., Li, R., & Deng, Y. (2015). Tolerance and avoidance characteristics of Prenant's Schizothoracin Schizothorax prenanti to
total dissolved gas supersaturated water. North American Journal of Fisheries Management, 35(4), 827–834. https://doi.org/10.1080/027559
47.2015.1052160
Wang, Y., Rhoads, B. L., & Wang, D. (2016). Assessment of the flow regime alterations in the middle reach of the Yangtze River associated with
dam construction: Potential ecological implications. Hydrological Processes, 30(21), 3949–3966. https://doi.org/10.1002/hyp.10921
Wang, Y., Zhang, N., Wang, D., & Wu, J. (2020). Impacts of cascade reservoirs on Yangtze River water temperature: Assessment and ecological
implications. Journal of Hydrology, 590, 125240. https://doi.org/10.1016/j.jhydrol.2020.125240
Wang, Z., Lee, J. H. W., & Xu, M. (2013). Eco-hydraulics and eco-sedimentation studies in China. Journal of Hydraulic Research, 51(1), 19–32.
https://doi.org/10.1080/00221686.2012.753554
Ward, D. L., Boyce, R. R., Young, F. R., & Olney, F. E. (1997). A review and assessment of transportation studies for juvenile Chinook salmon in the
Snake River. North American Journal of Fisheries Management, 17(3), 652–662. https://doi.org/10.1577/1548-8675(1997)017<0652:ARAA
OT>2.3.CO;2
Ward, E. J., Anderson, J. H., Beechie, T. J., Pess, G. R., & Ford, M. J. (2015). Increasing hydrologic variability threatens depleted anadromous
fish populations. Global Change Biology, 21(7), 2500–2509. https://doi.org/10.1111/gcb.12847
Warrick, J. A., Bountry, J. A., East, A. E., Magirl, C. S., Randle, T. J., Gelfenbaum, G., etal. (2015). Large-scale dam removal on the Elwha
River, Washington, USA: Source-to-sink sediment budget and synthesis. Geomorphology, 246, 729–750. https://doi.org/10.1016/j.
geomorph.2015.01.010
Watson, S., Schneider, A., Santen, L., Deters, K. A., Mueller, R., Pflugrath, B., et al. (2022). Safe passage of American eels through a novel
hydropower turbine. Transactions of the American Fisheries Society, 151(6), 711–724. https://doi.org/10.1002/tafs.10385
Webb, B. W., Hannah, D. M., Moore, R. D., Brown, L. E., & Nobilis, F. (2008). Recent advances in stream and river temperature research. Hydro-
logical Processes: International Journal, 22(7), 902–918. https://doi.org/10.1002/hyp.6994
Webb, T., & Meyer, A. (2019). The mystery and value of place names: Can history help us get to the bottom of Sturgeon Pool? Fisheries, 44(7),
314–320. https://doi.org/10.1002/fsh.10259
Wei, Q., He, J., Yang, D., Zheng, W., & Li, L. (2004). Status of sturgeon aquaculture and sturgeon trade in China: A review based on two recent
nationwide surveys. Journal of Applied Ichthyology, 20(5), 321–332. https://doi.org/10.1111/j.1439-0426.2004.00593.x
Weitkamp, D. E., & Katz, M. (1980). A review of dissolved gas supersaturation literature. Transactions of the American Fisheries Society, 109(6),
659–702. https://doi.org/10.1577/1548-8659(1980)109<659:ARODGS>2.0.CO;2
Welch, D. W., Rechisky, E. L., Melnychuk, M. C., Porter, A. D., Walters, C. J., Clements, S., etal. (2008). Survival of migrating salmon smolts
in large rivers with and without dams. PLoS Biology, 6(10), e265. https://doi.org/10.1371/journal.pbio.0060265
Whalen, K. G., Parrish, D. L., & McCormick, S. D. (1999). Migration timing of Atlantic salmon smolts relative to environmental and
physiological factors. Transactions of the American Fisheries Society, 128(2), 289–301. https://doi.org/10.1577/1548-8659(1999)128
%3C0289:MTOASS%3E2.0.CO;2
Whitledge, G. W., Knights, B., Vallazza, J., Larson, J., Weber, M. J., Lamer, J. T., etal. (2019). Identification of bighead carp and silver carp early-
life environments and inferring lock and dam 19 passage in the upper Mississippi river: Insights from otolith chemistry. Biological Invasions,
21(3), 1007–1020. https://doi.org/10.1007/s10530-018-1881-2
Williams, J. G., & Katopodis, C. (2016). Commentary-incorrect application of data negates some meta-analysis results in Bunt etal. (2012). River
Research and Applications, 32(10), 2109–2115. https://doi.org/10.1002/rra.3076
Wilson, C. C., Haxton, T., Wozney, K. M., & Friday, M. (2022). Historical genetic connectivity of lake sturgeon in a dammed Great Lakes tribu-
tary. Journal of Great Lakes Research, 48(3), 798–805. https://doi.org/10.1016/j.jglr.2022.03.007
Wilson, P.H. (2003). Using population projection matrices to evaluate recovery strategies for Snake River spring and summer Chinook salmon.
Conservation Biology, 17(3), 782–794. https://doi.org/10.1046/j.1523-1739.2003.01535.x
Winemiller, K., McIntyre, P., Castello, L., Fluet-Chouinard, E., Giarrizzo, T., Nam, S., etal. (2016). Balancing hydropower and biodiversity in
the Amazon, Congo, and Mekong. Science, 351(6269), 128–129. https://doi.org/10.1126/science.aac7082
Witt, A., Stewart, K., & Hadjerioua, B. (2017). Predicting total dissolved gas travel time in hydropower reservoirs. Journal of Environmental
Engineering, 143(12), 06017011. https://doi.org/10.1061/(ASCE)EE.1943-7870.0001281
Wittmann, M. E., Cooke, R. M., Rothlisberger, J. D., & Lodge, D. M. (2014). Using structured expert judgment to assess invasive species preven-
tion: Asian carp and the Mississippi-Great Lakes hydrologic connection. Environmental Science & Technology, 48(4), 2150–2156. https://doi.
org/10.1021/es4043098
Wofford, J. E. B., Gresswell, R. E., & Banks, M. A. (2005). Influence of barriers to movement on within-watershed genetic variation of coastal
cutthroat trout. Ecological Applications, 15(2), 628–637. https://doi.org/10.1890/04-0095
World Bank. (2008). Biodiversity, climate change, and adaptation: Nature-based solutions from the World Bank portfolio. World Bank. Retrieved
from https://openknowledge.worldbank.org/handle/10986/6216
Wotton, R. (1995). Temperature and lake-outlet communities. Journal of Thermal Biology, 20(1–2), 121–125. https://doi.
org/10.1016/0306-4565(94)00042-H
Wright, P.J., Orpwood, J. E., & Boulcott, P. (2017). Warming delays ovarian development in a capital breeder. Marine Biology, 164(4), 80. https://
doi.org/10.1007/s00227-017-3116-y
Wu, F., Du, K., & Liu, L. (2020). Effects of gas supersaturation on fertilized eggs, larvae and juveniles of grass carp and silver carp. Freshwater
Fishery, 50(4), 91–98. https://doi.org/10.1186/s12302-020-00330-9
Wu, Y., Fang, H., Huang, L., & Cui, Z. (2020). Particulate organic carbon dynamics with sediment transport in the upper Yangtze River. Water
Research, 184, 116193. https://doi.org/10.1016/j.watres.2020.116193
Xia, X., Yang, Z., & Wu, Y. (2009). Incorporating eco-environmental water requirements in integrated evaluation of water quality and quantity-a
study for the Yellow River. Water Resources Management, 23(6), 1067–1079. https://doi.org/10.1007/s11269-008-9315-z
Xia, Y., Feng, Z., Niu, W., Qin, H., Jiang, Z., & Zhou, J. (2019). Simplex quantum-behaved particle swarm optimization algorithm with application
to ecological operation of cascade hydropower reservoirs. Applied Soft Computing, 84, 105715. https://doi.org/10.1016/j.asoc.2019.105715
Xiao, F. (2021). Research on the evaluation framework of hydropower station removal of tributaries in Yunnan section of Lancang River basin.
Yunnan University. https://doi.org/10.27456/d.cnki.gyndu.2021.000697
Xie, P. (2003). Three-Gorges dam: Risk to ancient fish. Science, 302(5648), 1149–1151. https://doi.org/10.1126/science.302.5648.1149b
Xu, H., & Pittock, J. (2018). Limiting the effects of hydropower dams on freshwater biodiversity: Options on the Lancang River, China. Marine
and Freshwater Research, 70(2), 169–194. https://doi.org/10.1071/MF17394
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
63 of 64
Xu, W., Yang, Z., Yi, R., Yao, J., & Chen, X. (2020). Relationship between environmental variables and egg abundance of the four major Chinese
carps, downstream of the Three Gorges Reservoir. River Research and Applications, 37(8), 1191–1200. https://doi.org/10.1002/rra.3750
Xu, Z., Yin, X., Sun, T., Cai, Y., Ding, Y., Yang, W., & Yang, Z. (2017). Labyrinths in large reservoirs: An invisible barrier to fish migration and
the solution through reservoir operation. Water Resources Research, 53(1), 817–831. https://doi.org/10.1002/2016WR019485
Xue, H., Ma, Q., Li, R., Diao, M. J., Zhu, D., & Lu, J. (2018). Experimental study on the dissipation of supersaturated TDG during the jet breakup
process. Journal of Hydrodynamics, 31(5), 760–766. https://doi.org/10.1007/s42241-018-0106-6
Xue, S., Wang, Y., Liang, R., Li, K., & Li, R. (2019). Effects of total dissolved gas supersaturation in fish of different sizes and species. Interna-
tional Journal of Environmental Research and Public Health, 16(13), 2444. https://doi.org/10.3390/ijerph16132444
Yang, D., Liu, B., & Ye, B. (2005). Stream temperature changes over Lena River basin in Siberia. Geophysical Research Letters, 32(5). https://
doi.org/10.1029/2004GL021568
Yang, J., Pan, X., Chen, X., Wang, X., Zhao, Y., Li, J., etal. (2013). Overview of the artificial enhancement and release of endemic freshwater
fish in China. Zoological Research, 34(4), 267–280. https://doi.org/10.11813/j.issn.0254-5853.2013.4.0267
Yang, Q. H., Wei, Q., Qiao, Y., Zhang, L., Wang, J. S., Dai, J., etal. (2014). The response of the spawning of the four major Chinese carps to the
flow re-operation of the Three Gorges Reservoir in the Yangtze River, China. Advanced Materials Research, 1065–1069, 3223–3234. https://
doi.org/10.4028/www.scientific.net/AMR.1065-1069.3223
Yang, W., Yang, H., & Yang, D. (2020). Classifying floods by quantifying driver contributions in the Eastern Monsoon Region of China. Journal
of Hydrology, 585, 124767. https://doi.org/10.1016/j.jhydrol.2020.124767
Yang, Y., Zhang, M., Zhu, L., Liu, W., Han, J., & Yang, Y. (2017). Influence of large reservoir operation on water-levels and flows in reaches
below dam: Case study of the Three Gorges Reservoir. Scientific Reports, 7(1), 1–14. https://doi.org/10.1038/s41598-017-15677-y
Yang, Z., Zhu, Q., Cao, J., Jin, Y., Zhao, N., Xu, W., et al. (2021). Using a hierarchical model framework to investigate the relationships
between fish spawning and abiotic factors for environmental flow management. Science of the Total Environment, 787, 147618. https://doi.
org/10.1016/j.scitotenv.2021.147618
Yao, W., Rutschmann, P., & Sudeep (2015). Three high flow experiment releases from Glen Canyon Dam on rainbow trout and flannelmouth
sucker habitat in Colorado River. Ecological Engineering, 75, 278–290. https://doi.org/10.1016/j.ecoleng.2014.11.024
Ye, H., Deyle, E. R., Gilarranz, L. J., & Sugihara, G. (2015). Distinguishing time-delayed causal interactions using convergent cross mapping.
Scientific Reports, 5(1), 14750. https://doi.org/10.1038/srep14750
Yi, B., Yu, Z., & Liang, C. (1988). Gezhouba water control project and the four major Chinese carps in the Yangtze River. Hubei Science &
Technology Press.
Yi, Y., & Wang, Z. (2009). Impact from dam construction on migration fishes in Yangtze River Basin. Water Resources and Hydropower Engi-
neering, 40(1), 29–33. https://doi.org/10.13928/j.cnki.wrahe.2009.01.024
Yi, Y., Wang, Z., & Yang, Z. (2010). Impact of the Gezhouba and three Gorges dams on habitat suitability of carps in the Yangtze River. Journal
of Hydrology, 387(3–4), 283–291. https://doi.org/10.1016/j.jhydrol.2010.04.018
Yin, X., Yang, Z., & Petts, G. E. (2011). Reservoir operating rules to sustain environmental flows in regulated rivers. Water Resources Research,
47(8), W08509. https://doi.org/10.1029/2010WR009991
YoosefDoost, A., & Lubitz, W. D. (2020). Archimedes screw turbines: A sustainable development solution for green and renewable energy
generation—A review of potential and design procedures. Sustainability, 12(18), 7352. https://doi.org/10.3390/su12187352
Young, P.S., Cech, J. J., & Thompson, L. C. (2011). Hydropower-related pulsed-flow impacts on stream fishes: A brief review, conceptual model,
knowledge gaps, and research needs. Reviews in Fish Biology and Fisheries, 21(4), 713–731. https://doi.org/10.1007/s11160-011-9211-0
Yuan, Y., Huang, J., Wang, Z., Gu, Y., Li, R., Li, K., & Feng, J. (2020). Experimental investigations on the dissipation process of supersaturated
total dissolved gas: Focus on the adsorption effect of solid walls. Water Research, 183, 116087. https://doi.org/10.1016/j.watres.2020.116087
Yuan, Y., Huang, Y., Feng, J., Li, R., An, R., & Huang, J. (2018). Numerical model of supersaturated total dissolved gas dissipation in a channel
with vegetation. Water, 10(12), 1–22. https://doi.org/10.3390/w10121769
Yuan, Y., Wang, Y., Zhou, C., An, R., & Li, K. (2018). Tolerance of Prenant's Schizothoracin Schizothorax prenanti to total dissolved gas super-
saturated water at varying temperature. North American Journal of Aquaculture, 80(1), 107–115. https://doi.org/10.1002/naaq.10007
Yuan, Y., Wei, Q., Yuan, Q., Wang, Y., Liang, R., Li, K., & Zhu, D. Z. (2021). Impact of TDG supersaturation on native fish species under differ-
ent hydropower flood discharge programs. Aquatic Toxicology, 237, 105898. https://doi.org/10.1016/j.aquatox.2021.105898
Zarfl, C., Lumsdon, A., Berlekamp, J., Tydecks, L., & Tockner, K. (2015). A global boom in hydropower dam construction. Aquatic Sciences,
77(1), 161–170. https://doi.org/10.1007/s00027-014-0377-0
Zarri, L. J., Danner, E. M., Daniels, M. E., & Palkovacs, E. P. (2019). Managing hydropower dam releases for water users and imperiled fishes
with contrasting thermal habitat requirements. Journal of Applied Ecology, 56(11), 2423–2430. https://doi.org/10.1111/1365-2664.13478
Zeng, C., Mo, K., Guan, T., Chen, Q., Li, T., & Zhang, H. (2020). Effect of total dissolved gas supersaturation on fish in the reser-
voir between cascade hydropower stations. Hydro-Science and Engineering, 6, 32–41. Retrieved from http://slsy.nhri.cn/en/article/
id/1bcf850a-54f4-4777-8c25-6ce052fe4bbf
Zhang, H., Jarić, I., Roberts, D. L., He, Y., Du, H., Wu, J., etal. (2020). Extinction of one of the world's largest freshwater fishes: Lessons for
conserving the endangered Yangtze fauna. Science of the Total Environment, 710, 136242. https://doi.org/10.1016/j.scitotenv.2019.136242
Zhang, H., Kang, M., Shen, L., Wu, J., Li, J., Du, H., etal. (2020). Rapid change in Yangtze fisheries and its implications for global freshwater
ecosystem management. Fish and Fisheries, 21(3), 601–620. https://doi.org/10.1111/faf.12449
Zhang, H., Kang, M., Wu, J., Wang, C., Li, J., Du, H., etal. (2019). Increasing river temperature shifts impact the Yangtze ecosystem: Evidence
from the endangered Chinese sturgeon. Animals, 9(8), 583. https://doi.org/10.3390/ani9080583
Zhang, J., Shang, Y., Liu, J., Lu, J., Wei, S., Wan, Z., etal. (2021). Optimisation of reservoir operation mode to improve sediment transport
capacity of silt-laden rivers. Journal of Hydrology, 594, 125951. https://doi.org/10.1016/j.jhydrol.2020.125951
Zhang, L., Wang, H., Gessner, J., Congiu, L., Haxton, T. J., Jeppesen, E., etal. (2023). To save sturgeons, we need river channels around hydro-
power dams. Proceedings of the National Academy of Sciences, 120(13), e221786120. https://doi.org/10.1073/pnas.2217386120
Zhang, P., Qiao, Y., Grenouillet, G., Lek, S., Cai, L., & Chang, J. (2021). Responses of spawning thermal suitability to climate change and
hydropower operation for typical fishes below the Three Gorges Dam. Ecological Indicators, 121, 107186. https://doi.org/10.1016/j.
ecolind.2020.107186
Zhang, W., Hu, P., Hou, J., Jia, Y., & Xu, F. (2020). A research of the impacts of hydropower development on the survival and reproduction of fish
and related conservation in the Hengduan Mountains. Journal of China Institute of Water Resources and Hydropower Research, 18(1), 12–20.
https://doi.org/10.13244/j.cnki.jiwhr.2020.01.002
Zhao, J., Cao, Y., Li, S., Li, J., Deng, Y., & Lu, G. (2011). Population genetic structure and evolutionary history of grass carp Ctenopharyngodon
idella in the Yangtze River, China. Environmental Biology of Fishes, 90(1), 85–93. https://doi.org/10.1007/s10641-010-9720-z
Reviews of Geophysics
CHEN ETAL.
10.1029/2023RG000819
64 of 64
Zheng, L., Lv, Z., Li, F., Zhang, L., & Yu, W. (2014). Fish community structure in the Yellow River estuary: Effect of water and sediment
discharge regulations. Journal of Fishery Sciences of China, 21(3), 602–610. https://doi.org/10.3724/SP.J.1118.2014.00602
Zheng, Y., Zhang, Y., Xie, Z., Shin, P.K. S., Xu, J., Fan, H., etal. (2022). Seasonal changes of growth, immune parameters and liver function in
wild Chinese sturgeons under indoor conditions: Implication for artificial rearing. Frontiers in Physiology, 13, 894729. https://doi.org/10.3389/
fphys.2022.894729
Zhou, X., Wang, K., Chen, D., Liu, S., Duan, X., Wang, D., etal. (2019). Effects of ecological operation of Three Gorges Reservoir on larval
resources of the four major Chinese carps in Jianli section of the Yangtze River. Journal of Fisheries of China, 43(8), 1781–1789. https://doi.
org/10.11964/jfc.20180711365
Zhu, Z., Soong, D. T., Garcia, T., Behrouz, M. S., Butler, S. E., Murphy, E. A., etal. (2018). Using reverse-time egg transport analysis for predict-
ing Asian carp spawning grounds in the Illinois River. Ecological Modelling, 384, 53–62. https://doi.org/10.1016/j.ecolmodel.2018.06.003
Zhuang, P., Ke, F., Wei, Q., He, X., & Cen, Y. (1997). Biology and life history of Dabry’s sturgeon, Acipenser dabryanus, in the Yangtze River.
Environmental Biology of Fishes, 48(1), 257–264. https://doi.org/10.1023/A:1007399729080
Zhuang, P., Zhao, F., Zhang, T., Chen, Y., Liu, J., Zhang, L., & Kynard, B. (2016). New evidence may support the persistence and adaptability of
the near-extinct Chinese sturgeon. Biological Conservation, 193, 66–69. https://doi.org/10.1016/j.biocon.2015.11.006
Zielinski, D. P., McLaughlin, R., Castro-Santos, T., Paudel, B., Hrodey, P., & Muir, A. (2019). Alternative sea lamprey barrier technologies:
History as a control tool. Reviews in Fisheries Science & Aquaculture, 27(4), 438–457. https://doi.org/10.1080/23308249.2019.1625300
Zielinski, D. P., Miehls, S., & Lewandoski, S. (2022). Test of a screw-style fish lift for introducing migratory fish into a selective fish passage
device. Water, 14(15), 2298. https://doi.org/10.3390/w14152298
Ziv, G., Baran, E., So, N., Rodríguez-Iturbe, I., & Levin, S. (2012). Trading-off fish biodiversity, food security, and hydropower in the Mekong
River Basin. Proceedings of the National Academy of Sciences, 109(15), 5609–5614. https://doi.org/10.1073/pnas.1201423109
References From the Supporting Information
Chen, A., Wang, D., Sui, X., Chen, K., & Wu, M. (2017). Retrospective of research on water temperature impacts of Xiaolangdi reservoir. Yellow
River, 39(8), 63–66.
Gray, R., Jones, H. A., Hitchcock, J. N., Hardwick, L., Pepper, D., Lugg, A., etal. (2019). Mitigation of cold-water thermal pollution downstream
of a large dam with the use of a novel thermal curtain. River Research and Applications, 35(7), 855–866. https://doi.org/10.1002/rra.3453
Jiang, B., Wang, F., & Ni, G. (2018). Heating impact of a tropical reservoir on downstream water temperature: A case study of the Jinghong dam
on the Lancang River. Water, 10(7), 951. https://doi.org/10.3390/w10070951
Ling, T., Gerunsin, N., Soo, C., Nyanti, L., Sim, S., & Grinang, J. (2017). Seasonal changes and spatial variation in water quality of a large young
tropical reservoir and its downstream river. Journal of Chemistry, 2017, 8153246. https://doi.org/10.1155/2017/8153246
Nikulenkov, A. M., Dvornikov, A. Y., Rumynin, V. G., Ryabchenko, V. A., & Vereschagina, E. A. (2017). Assessment of allowable thermal load
for a river reservoir subject to multi-source thermal discharge from operating and designed Beloyarsk NPP units (South Ural, Russian Federa-
tion). Environmental Modeling & Assessment, 22(6), 609–623. https://doi.org/10.1007/s10666-017-9562-6
Ren, Y., Zhao, L., Cao, H., & Ruan, Y. (2020). Influence of ecological regulation of cascade reservoirs in the lower Jinsha River. In Ecology and
environmental monitoring of three Gorges (Vol. 5,pp.8–13).
Seyedhashemi, H., Moatar, F., Vidal, J. P., Diamond, J. S., Beaufort, A., Chandesris, A., & Valette, L. (2021). Thermal signatures identify
the influence of dams and ponds on stream temperature at the regional scale. Science of the Total Environment, 766, 142667. https://doi.
org/10.1016/J.SCITOTENV.2020.142667
Wiejaczka, Ł., Kijowska-Strugała, M., Pierwoła, P., & Nowak, M. (2018). Water temperature dynamics in a complex of reservoirs and its effect
on the temperature patterns of a mountain river. Water Resources, 45(6), 861–872. https://doi.org/10.1134/S0097807818060167
Xing, Z., Fong, D. A., Yat-Man Lo, E., & Monismith, S. G. (2014). Thermal structure and variability of a shallow tropical reservoir. Limnology
& Oceanography, 59(1), 115–128. https://doi.org/10.4319/lo.2014.59.1.0115
Yang, R., Wu, S., Wu, X., Ptak, M., Li, X., Sojka, M., etal. (2022). Quantifying the impacts of climate variation, damming, and flow regulation
on river thermal dynamics: A case study of the Włocławek reservoir in the Vistula river, Poland. Environmental Sciences Europe, 34(3), 2–11.
https://doi.org/10.1186/s12302-021-00583-y
... The reduction in Cuban finfish landings has also been attributed to the decrease in the nutrient load in coastal areas because of the damming of rivers and the reduction in fertilizers in agriculture [5,6,11,12]. Similar impacts caused by dam construction have been recorded in the Black Sea [47], cited in [12], the Mediterranean Sea [48,49], and the Gulf of Mexico [50]. In the 1960s, Cuban water resources were limited, and the government established a dam construction program to mitigate the intense and prolonged drought in 1961-1962 and the severe flooding caused by hurricanes [15]. ...
Article
Full-text available
Marine finfish landings in Cuba have decreased during the last 30 years. However, in Cuba’s most productive fishing region, certain species, including rays, herrings, and snappers, have had increased landings over the past decade. Despite these anomalies, no comprehensive analysis of the interactions among multispecies landing dynamics, environmental factors, and fishing efforts has been carried out. This study estimates the dynamics of multispecies finfish landings between 1981 and 2017 on the southeastern coast of Cuba. A log-normal generalized additive model (GAM) was fit to evaluate the effects of various environmental and effort-related variables on the total landings. During the period analyzed, the finfish landings and fishing effort decreased by 46% and over 80%, respectively. Despite concerns about overfishing, landings per unit of effort (LPUE) increased by 2.8 times. The total fish landings were significantly related to changes in the fishing effort, coastal vegetation, rainfall, chlorophyll-a, and the Southern Oscillation Index (SOI). This study highlights the changing relationship between the landings and fishing effort, suggesting that LPUE may not accurately reflect true stock abundance. The findings of this study will assist in integrating the dynamics of finfish species, ecosystem status, and management actions for Cuba’s most productive fishing zone.
... 1 Among these activities, dam construction, dredging, and shipping have a profound effect on the survival and reproduction of aquatic organisms. 2 The Yangtze River, the world's third-longest river and a hotspot for fish diversity has suffered extensive anthropogenic interventions. Construction of over 50,000 dams and the frequent dredging affecting 70% of the Yangtze River's length has fragmented and degraded aquatic habitats, leading to severe fish depletion. ...
... Even though barriers are known to fragment habitats, cut off habitat connections, and alter the hydrological, thermal, solute, and sediment regime (Chen et al., 2023), the ecological importance of an uninterrupted river continuum is only gradually recognized by decisionmakers. Nevertheless, legislation such as the European Water Framework Directive (European Commission, 2000), the European Biodiversity Strategy (European Commission, 2020) or the recently proposed European Nature Restoration Law (European Commission, 2023;Stoffers et al., 2024;Hering et al., 2023) has been or will accelerate efforts towards restoring longitudinal connectivity in the future. ...
Article
In this study, we explored how barriers such as dams have affected the longitudinal connectivity of riverine habitats from the perspective of potamodromous fish. For this purpose, connectivity changes are investigated in the central part of the Austrian Danube system, where the national reporting for the EU Water Framework Directive provides detailed information on the position and characteristics of barriers as well as the distribution range of native fish species. This assessment is based on an estimation regarding the quantitative upstream and downstream passability of individual barriers, where we further investigate three different passability scenarios to account for uncertainties. We then apply several combinations of passability scenarios and assumptions on dispersal distances to calculate a series of network-based reach and catchment connectivity indices. On average, the estimation of barrier passability indicated a high downstream passability, while upstream passability was substantially lower across scenarios. Furthermore, existing fish passes were estimated to have increased passability on average between 20 % and 24 %. Overall, the results indicated a strong effect of barriers on the longitudinal connectivity of the investigated river network. Catchment scale indices revealed a loss of connectivity, which increased with dispersal distance. Reach connectivity indices displayed a strong disruption of the natural connectivity gradient along the river network and indicated that individual river reaches have, on average, become more isolated in addition to the overall decrease in connectivity. The average loss of connectivity across scenarios was estimated at 72 % (SD = 16 %) when taking into account all connections to other reaches and 66 % (SD = 7 %) when only connections to upstream reaches were considered. We conclude that longitudinal connectivity in the Austrian Danube system is still severely compromised, making it increasingly challenging for potamodromous fish species to complete their life cycle. This issue is further amplified by the severe loss of fish habitats as a consequence of river engineering.
... On the other hand, alterations in river flow patterns might have profound impacts on riverine ecosystems [52,53], affecting biodiversity [54], water quality [55], and the overall health of aquatic habitats. Changes in flow regimes may disrupt the life cycles of native aquatic species, including fish and macroinvertebrates, by affecting spawning cues, nursery habitats, and food availability [56][57][58]. Additionally, reduced flow volumes can concentrate pollutants, elevate temperatures, and decrease oxygen levels, further stressing aquatic life [59,60]. ...
Article
Full-text available
Salmonids create an egg nest, called a redd, in hyporheic streambed gravel to bury their eggs, characterized by a pit and a hump topography, resembling a dune. Embryos' survival depends on downwelling oxygen‐rich stream water fluxes, influenced by the redd shape, stream hydraulics, and the hydraulic conductivity of the redd sediment, K. We hypothesize that downwelling fluxes increase with stream discharge and redd aspect ratio (AR = A/L, with A, the redd amplitude and L, its length), and they can be predicted using a set of dimensionless numbers, which include the stream flow Reynolds (Re) and Froude (Fr) numbers, AR, and the redd relative submergence (A/Y0, with Y0, the water depth). We address our goal by simulating the surface and subsurface flows with numerical hydraulic models linked through the near‐bed pressure distribution quantified with a two‐phase (air‐water) two‐dimensional surface water computational fluid dynamics model, validated with experiments. We apply the modeling approach to five redd sizes, which span the field observed range (from ∼1 to ∼4 m long), and by increasing discharge from shallow (0.1 m) and slow (0.15 m/s) to deep (8 m) and fast (3.3 m/s). Results confirm that downwelling fluxes increase with discharge and redd aspect ratio due to the increased near‐bed head gradient over the redd. Averaged downwelling fluxes are a function of Re, Fr, AR, and hydraulic conductivity of the redd and not of the undisturbed streambed material. The derived regression equation may help evaluate regulated and unregulated flow impacts on hyporheic flows during embryo incubation.
Article
Full-text available
Sturgeon, a flagship taxon for functional rivers and adjacent seas, have existed for more than 200 million years (1). Around since the time of dinosaurs, they have survived at least two mass extinction events (2). However, all contemporary sturgeon species are now endangered, the result of human activities on rivers that have taken place during the last two centuries. The consequences of these effects are now becoming apparent.
Article
Full-text available
Field investigation indicated that the reduction in fish spawning was associated with the alteration in water temperatures, even a 2–3 °C monthly difference due to reservoir operations. However, the physiological mechanism that influences the development of fish ovary (DFO) remains unclear. Thus, experiments of Coreius guichenoti were conducted at three different temperatures, optimal temperature (~20 °C, N) for fish spawning, lower (~17 °C, L), and higher (~23 °C, H), to reveal the effects of altered water temperature on the DFO. Comparisons were made between the L and N (LvsN) conditions and H and N (HvsN) conditions. Transcriptomic analysis differentially expressed transcripts (DETs) related to heat stress were observed only in LvsN conditions, indicating that the DFO showed a stronger response to changes in LvsN than in HvsN conditions. Upregulation of DETs of vitellogenin receptors in N temperature showed that normal temperature was conducive to vitellogenin entry into the oocytes. Other temperature-sensitive DETs, including microtubule, kinesin, dynein, and actin, were closely associated with cell division and material transport. LvsN significantly impacted cell division and nutrient accumulation in the yolk, whereas HvsN only influenced cell division. Our results highlight the impact of altered water temperature on the DFO, thereby providing insights for future reservoir operations regarding river damming and climate change and establishing fish conservation measures.
Article
Full-text available
Intermittent water releases from hydropower plants, called hydropeaking, negatively affect river biota. The impacts mainly depend on hydrological alterations, but changes in physical habitat conditions are suspected to be co‐responsible. For example, hydropeaking accompanied by a sudden change of water temperature in the downstream river—called thermopeaking—is also presumed to impair aquatic ecosystems. Still, knowledge about these thermopeaking impacts on aquatic species and life‐stages is limited. We performed flume experiments under semi‐natural conditions to fill this knowledge gap, simulating single hydropeaking events with a change in water temperature. As response parameters, we quantified the drift and stranding of early life‐stages of European grayling (Thymallus thymallus L.), a key fish species of Alpine hydropeaking rivers. Hydropeaking events with a decrease in water temperature (“cold thermopeaking”) led to significantly higher downstream drift (mean = 51%) than events with increasing water temperature (“warm thermopeaking”, mean = 27%). Moreover, during cold thermopeaking, a comparably high fish drift was recorded up to 45 min after the start of peak flows. In contrast, drift rates quickly decreased after 15 min during warm thermopeaking. Remarkably, the spatial distribution of downstream drift along gravel bars during cold thermopeaking showed the opposite pattern compared to those triggered by warm thermopeaking events indicating different behavioral responses. Furthermore, the stranding rates of the cold thermopeaking trials were twice as high (mean = 31%) as those of the warm thermopeaking experiments (mean = 14%). The outcomes present vital information for improving mitigation measures and adapting environmental guidelines.
Article
Flow and water surface elevation recession rate Regulated and unregulated flows Dam operation A B S T R A C T Fish can be stranded in pools during flow recessions that may occur naturally or as a result of dam operations. Past studies have recommended dam operation guidelines that only consider water surface elevation (WSE) recession rate to reduce fish stranding potential. Here, we holistically consider the role of flow, WSE recession rate, and river-floodplain topography under unregulated and regulated flow regimes to predict the potential of stranding pool formation. The proposed methodology leverages advances in two-dimensional (2D) surface hydraulic (HD) modeling and high-resolution topobathymetric LiDAR surveys. We applied this method to the South Fork Boise River (SFBR), Idaho, USA. We developed a 2D HD model supported by meter-resolution top-obathymetric data to simulate channel and floodplain hydraulics and stranding pool formation for different flow scenarios. Comparison between modeled and fieldsurveyed stranding pool locations shows good match, supporting the use of this method to predict stranding pool potential. Results show that potential for stranding pool formation (number of pools and their surface area) increases with flow for the same flow recession rate. Contrary to expectations , our model predicts that unregulated flows have a considerably higher number of potential fish stranding pools than for regulated flows in the SFBR. The reach-average WSE recession rates were higher for unregulated (0.07 to 0.09 cm/h) than for regulated (0.02 to 0.04 cm/h) flow based on daily mean flows. Our study has shown the importance of river and floodplain morphology and hydraulics to predict fish stranding pool locations and analyze impacts of flow regulations on fish stranding. This approach may provide helpful information to identify potential fish stranding pools and mitigation and to manage for fish stranding issues in regulated and unregulated river systems.
Article
In the Brazilian Amazon, Indigenous people and researchers join hands to monitor the impacts of a controversial dam.
Article
Hydropeaking due to hydropower production can have negative impacts on aquatic fauna. One of the mechanisms for causing impacts on fish and aquatic macroinvertebrates is linked to the rapid dewatering of habitats, which can result in stranding or trapping. The magnitude of these impacts depends both on the characteristics of the flow variations and of the river morphology, as well as biological parameters (species, behavior, etc). When discharge is rapidly reduced, the risk of impacts on fishes (and notably the risk of fish stranding in dewatered zones along the riverbank) is frequently assessed by calculations of vertical ramping velocity among other methods. However, to assess fish stranding risks, the lateral ramping velocity calculated as a horizontal ramping rate (HRR) appears to be a more relevant indicator as it directly measures shoreline drawdown rates. HRR has the advantage of integrating river morphology, but it remains challenging to calculate HRRs in complex situations such as braided rivers. Using hydraulic simulations of the Durance, a gravel bed braided river, we have developed an innovative approach for HRR calculation. Considering two simulated flows, the algorithms for the calculations require partitioning the finite elements into wet and drying meshes. To recommend rates of lowering discharges during hydropeaking events, further studies are required to evaluate more precisely HRR limits for fish stranding regarding biotic and abiotic parameters: species, sizes, nychthemeral cycles, temperature, substrate, and so forth.
Article
With the continuous development of hydropower on a global scale, stranding of freshwater fishes is of growing concern, and an understanding of the mechanisms and variables affecting fish stranding in hydropeaking rivers is urgently needed. In particular, a methodology is required to identify the magnitude and timing at which fish stranding occurs in relation to environmental conditions. Here, we studied fish stranding in three reaches downstream of a hydropeaking generation station in the Saskatchewan River, Saskatchewan, Canada, using an innovative remote photography approach with 45 trail cameras and traditional transect monitoring, conducting 323 transects. We observed that juvenile sport and commercial fish species are stranding at a higher proportion than small bodied fish species. The remote photography approach provided more precise fish stranding timing and associated the environmental and physical conditions with a given stranding event, but captured fewer fish and only rarely allowed species identification. The comparison of the two methodologies resulted in similar stranded fish densities, but the remote photography allowed for continuous observations whereas the transect monitoring was limited by the observer availability in the field. Remote photography allowed for additional information on the scavenging of stranded fish, with scavenging occurring on average within 240 minutes of the fish being stranded. The probability of fish stranding increased significantly with increasing water temperature and substrate particle size resulted in greater stranding on finer substrates. Our findings have important implications for hydroelectric flow management by introducing an innovative, standardized method to study the effects of hydropeaking events on fish stranding that can be applied to increase our understanding of the impacts of hydropeaking on fish communities.
Article
Since the early 1900s,more than 1700 dams have been removed from rivers in the United States. Native American Tribes have played a key role in many significant removals,bringing cultural,economic,and legal resources to bear on the process. Their involvement contrasts with the displacement and marginalisation that have historically characterised the relationship between Native Americans and the dams built by settler – colonial governments on their rivers. Our research investigates Tribal involvement in dam removals,with examples from the Ottaway,Penobscot,and Elwha rivers. We ask the following: what roles have Tribes played in successful removals? How do dam removals affect and reflect shifting relations between Tribal governments and non-Tribal actors? Our research finds that Tribal involvement provides opportunities for inserting underacknowledged values and resource claims into dam removal efforts,and that it facilitates new collaborations and alliances. We also find evidence of Tribal involvement affecting the nature and practice of river restoration through dam removal. We conclude that the involvement of Tribes in dam removal contributes to important shifts in environmental politics in the US,and that it also creates opportunities for restorative environmental justice for Native Americans and their rivers. © THIS ARTICLE IS DISTRIBUTED UNDER THE TERMS OF THE CREATIVE COMMONS ATTRIBUTION-NONCOMMERCIAL-SHAREALIKE LICENSE WHICH PERMITS ANY NON COMMERCIAL USE,DISTRIBUTION,AND REPRODUCTION IN ANY MEDIUM,PROVIDED THE ORIGINAL AUTHOR(S) AND SOURCE ARE CREDITED. SEE HTTPS://CREATIVECOMMONS.ORG/LICENSES/BY-NC-SA/3.0/FR/DEED.EN