Open access peer-reviewed chapter

Characteristics and Process Interactions in Natural Fluvial–Riparian Ecosystems: A Synopsis of the Watershed-Continuum Model

Written By

Lawrence E. Stevens, Raymond R. Johnson and Christopher Estes

Reviewed: 19 August 2022 Published: 13 October 2022

DOI: 10.5772/intechopen.107232

From the Edited Volume

River Basin Management - Under a Changing Climate

Edited by Ram L. Ray, Dionysia G. Panagoulia and Nimal Shantha Abeysingha

Chapter metrics overview

234 Chapter Downloads

View Full Metrics

Abstract

The watershed-continuum model (WCM) describes fluvial-riparian ecosystems (FREs) as dynamic reach-based ecohydrogeological riverine landscapes linking aquatic, riparian, and upland domains within watersheds. FRE domains include aquatic (channels, hyporheic zones, springs, other groundwater zones and in-channel lakes), riparian, and adjacent upland zones, all of which can interact spatio-temporally. Occupying only a minute proportion of the terrestrial surface, FREs contain and process only a tiny fraction of the Earth’s freshwater, but often are highly productive, flood-disturbed, and ecologically interactive, supporting diverse, densely-packed biotic assemblages and socio-cultural resource uses and functions. FRE biodiversity is influenced by hydrogeomorphology, ecotonal transitions, and shifting habitat mosaics across stage elevation. Thus, the WCM integrates physical, biological, and socio-cultural characteristics, elements, and processes of FREs. Here, we summarize and illustrate the WCM, integrating diverse physical and ecological conceptual models to describe natural (unmanipulated) FRE dynamics. We integrate key processes affecting FRE forms and functions, and illustrate reach-based organization across temporal and spatial scales. Such a holistic approach into natural FRE structure and functions provides a baseline against which to measure and calibrate ecosystem alteration, management, and rehabilitation potential. Integration of groundwater, fluvial, and lacustrine ecological interactions within entire basins supports long-term, seasonally-based sustainable river management, which has never been more urgently needed.

Keywords

  • conceptual model
  • continuum
  • ecohydrogeology
  • ecosystem
  • fluvial
  • riparian
  • rivers
  • springs
  • streams
  • watershed

1. Introduction

Fluvial-riparian ecosystems (FREs) are watershed- (catchment-, drainage area-) based riverine landscape systems that integrate aquatic, riparian, and upland domains within watersheds, linking physical, biological, and cultural-economic processes [1, 2, 3]. From the context of a system FREs consist of “… a structured set of objects and/or attributes… (,) components or variables …that exhibit discernible relationships with one another and operate together as a complex whole …” ([4], 1–2; [5, 6, 7]). FREs include all sources of water that contribute to the basin’s riverine ecosystem, including springs, surface runoff, lakes, and atmospheric sources such as humidity and fog. Only an average of 2120 km3 (0.0002 percent) of the world’s water exists in river systems at any given time [8] (Figure 1). But while rivers process only a tiny fraction of the Earth’s fresh water and occupy only a minute proportion of the Earth’s terrestrial surface, FREs are highly productive and ecologically interactive, often supporting complex landforms and diverse, densely packed biotic assemblages that change across fine to coarse spatial and temporal scales [9, 10] and burgeoning human populations. FRE physical and biological characteristics and processes among aquatic and riparian domains step, intergrade, and may interact through reaches within the watershed, from the headwaters to the terminus in an endorheic basin or the sea, and can extend far out into the submarine environment (e.g., [11]; Figure 1). Physically, FREs are “complex adaptive process–response system(s) with …morphological system(s) of channels, floodplains, hillslopes, deltas, … and cascading system(s) of …water and sediment” [12].

Figure 1.

Surface hydrological cycle and fluvial-riparian landscape within the watershed. Numbers represent the percent of freshwater storage 6 (redrawn from [1]).

Within FREs, the riparian domain is a zone of “transition between the aquatic ecosystem and the adjacent terrestrial ecosystem [13]. Riparian zones function as filters that reduce the impacts of flooding and surface runoff, as habitats that support vegetation, fish, and wildlife populations and habitat, and often provide critically important ecosystem goods and services [13, 14, 15, 16, 17, 18, 19]. Elevated FRE biodiversity is linked to, and influenced by factors including tectonics, geology, climate, hydrology, geomorphology, and latitude, in the context of shifting habitat mosaic heterogeneity and ecotonal dynamics [20, 21, 22, 23]. Human reliance on FREs, and our species’ evolutionary history and modern demography clearly demonstrate that reliance. As human domination of the Earth has progressed, rivers have been subjected to a host of anthropogenic alterations, including resource extraction, groundwater withdrawal, flow diversion and regulation, water quality degradation, and introduction of non-native species. The natural dimensions and human impacts on FREs have stimulated deep interest, concern, and much basic and applied research, generating a vast literature and prompting development of a suite of interrelated, but not necessarily integrated ecohydrogeological models. Focus on particular aspects of FRE channel development, geomorphology, ecology, or sociology has sometimes diminished wholistic integration. Also, graphic representation of FRE ecology can be improved to enhance conceptualization, and improve educational outreach.

Here, we provide an overview description and illustrated summary of the watershed-continuum model (WCM) [1], which couples interdisciplinary physical and ecological conceptualization of FRE ecology. The WCM links conceptual models of fluvial spatio-temporal development and geomorphology across stream order and reaches FRE ecology, trophic energy and matter dynamics, biodiversity, and evolutionary interactions from the river’s source to its mouth. We provide a chronological analysis of major concepts in FRE ecohydrogeology (Table 1) and illustrate the WCM with an improved spatio-temporal, reach-linked conceptual diagram that integrates “bottom-up” physical factors, including geology, hydrology, geochemistry, geomorphology, sedimentology, and fluvial climate, with aquatic and riparian biotic assemblages and ecosystem structure within the watershed, and the potential for trophic cascade effects [26, 27, 36, 37, 99, 100, 101, 102]. Due to the brevity of this manuscript, we emphasize here integration of physical and ecological conceptual elements and processes in natural, unmanipulated FREs, recognizing that such an integration is needed as a basis to improve watershed stewardship.

ModelDescription
Trophic-dynamic aspect of ecology [24, 25]Consistency of energy dynamics across trophic levels within ecosystems (e.g., Cedar Bog, Silver Springs)
Stream order classification [26, 27]Classification of dendritic hierarchy
Lentic ecosystem ecology (e.g. [28])Limnology of fresh waters
Stream channel development [29, 30, 31, 32, 33]Depth, width, velocity, slope, discharge, and sediment load interactively control channel geometry
Dynamic equilibrium concept [34]Channel geomorphology and energy moves toward equilibrium, never reaching it due to subsequent perturbation
Perpetual Riparian Succession [35]Regular flood scouring of floodplains keeps riparian vegetation in a state of perpetual or suspended succession
Lotic ecosystem ecology [36, 37]Limnology of moving fresh waters
Riparian ecosystem ecology [13, 14, 15, 16, 17, 18, 19, 20, 21, 23, 38, 39, 40, 41, 42, 43, 44]Riparian ecosystems are biologically diverse, structured, and highly ecologically interactive landscapes
River Continuum Concept [6, 45]Rivers as flow-integrated ecosystems; invertebrate feeding guilds vary longitudinally
Dynamic Equilibrium Model (species richness) [46, 47, 48, 49]Intermediate levels of disturbance intensity and productivity maximize the biodiversity of passively dispersing organisms
Nutrient spiraling concept [50, 51]Autochthonous (endogenic) and allochthonous (exogenic) nutrients and matter are transported through helical ecological pathways through FREs
Serial discontinuity concept [52, 53]Relationships between natural as well as anthropogenic dams and tributaries regulate downstream FRE structure and function
Ecological and land use history [54, 55]FRE ecology requires detailed and long-term understanding of geologic, hydrographic, biota, and land use history of the basin
Flood Pulse Concept [56, 57], e.g. [58]Flooding regulates developmental cyclicity of rivers
Stream channel classification [59, 60, 61, 62, 63, 64]Systematic analysis of reach-based channel geometry
River Productivity Model [65, 66]Fluvial productivity is spatially heterogeneous, affecting FRE ecological function
Process domain concept [67, 68]Tributary and/or bedrock-controlled reaches generate fluvial geomorphic discontinuities
Telescoping material spiral model [69]Material spirals tend to lengthen with stream order
Link discontinuity concept [70, 71]Large tributaries create abrupt discontinuities, generating multi-reach alteration downstream
Riparian eco-hydrogeomorphology [13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 72, 73, 74, 75, 76]Channel geomorphology and stage shape riparian zonation and ecology
Network Dynamics Hypothesis [77, 78, 79]The complexity of the overall basin shapes tributary contributions to the mainstream
Variation in solar radiation influences FRE ecology [80, 81]Cliff shading in canyon-bound FRE reaches strongly influences aquatic and riparian productivity
Top predators affect fluvial geomorphology [82]Trophic cascades can influence channel geomorphology at a reach scale
Riparian vegetation Niche Construction Perspective and Niche-Box Model [83, 84]Riparian plant species life history traits interact successionally with fluvial landform dynamics
Biogeochemical retention and processing network [85, 86, 87]All parts of the fluvial system (mainstreams, floodplains, lakes, and wetlands) form a fluvial bio-geochemical retention and processing network
FRE species life history strategies interactions [86, 87, 88, 89, 90]FRE guilds of plant and fish include competitive, ruderal, and stress-tolerant species
River wave concept [91]FRE aquatic domain processes can be viewed as waves that determine or regulate production and transport of organic matter
Ephemeral stream ecology [42, 43, 92, 93]Seasonally ephemeral streams are punctuated, rapidly functioning biogeochemical systems
Biological stream width theory [94]Resource subsidies from the FRE aquatic domain extend well into the surrounding upland terrain
Least Action Principle (LAP) [95]FRE teleomatic change through the LAP to achieve maximum energy efficiency
Spring ecosystem classification and ecology [96, 97]As “zero-order streams”, spring contributions to FREs vary by geomorphic type
Integrated Metasystems Theory [98]Regional processes act across spatial scales to control FRE form and function

Table 1.

Concepts in physical and biological FRE ecology, presented in approximate chronological order of publication. Rows in gray indicate concepts that are primarily focused on physical processes, whereas unshaded rows indicate concepts that are more strongly focused on eco-biological issues.

We reserve more detailed discussions on the details of riparian and aquatic community ecology related to the WCM for subsequent summaries [1] but focus on Integration and clear depiction of FRE domain interactions among reaches and over time within the basin. We discuss understudied issues and opportunities, the resolution of which will help advance FRE ecology in the future. Such an objective is essential for sustainable management of rivers.

Advertisement

2. FRE elements, processes, and interactions

2.1 Overview

Natural FREs are terrestrial, dendritic surface-water flow paths transporting matter and energy both downslope and upslope through their channels within watersheds, with flow contributed by groundwater and multiple surface water, as well as atmospheric sources (Figures 1 and 2). Rivers play a disproportionately large role in geochemical cycling, biodiversity, human culture, esthetics, and socio-economic appropriation [98, 103]. Collectively, physical state variables regulate flow, hydrography, water quality, and sediment transport in a “bottom-up” ecosystem fashion, generating FRE habitat templates at a local scale and within reaches that are created by parent rock geology and geologic structure, as well as tributary influences [6, 67, 69, 98]. FREs are structured in relation to latitude, the extent of geological constraint, flooding, sediment transport, potential productivity, and many other factors. FRE aquatic and riparian habitats are colonized through both active and passive biogeographic processes and change across stage and over time through disequilibrial or successional processes [34, 46, 47, 48, 74, 75]. We present a rough chronology of advances in physical and biological conceptual models (Table 1), with additional references in the WCM [1], and illustrate this complex, “bottom-up” array of physical influences and responses on FRE structure (Figure 2).

Figure 2.

Conceptual FRE model depicting interactions among independent and dependent physical and ecological variables and processes, across stream order (zero headwaters to X- mouth) and time (T1 to TX). Reaches depicted reflect general patterns of width. Line thickness indicates strength of effect; black arrow points indicate relative impacts of tributaries of different sizes. C – Consumers, P – Productivity or producers, PAR – Photosynthetically-active radiation, PNAPP – Potential net annual primary productivity R – Respiration, WQ – Water quality. Redrawn from [1].

2.2 Physical conceptual models

Basin and consequent FRE development vary in relation to complex interactions among tectonic terrain, geologic structure, parent bedrock, and climate over time (Table 1; Figure 1). Tectonic setting and geologic history ultimately control regional aquifer and surface watershed development (e.g., [104, 105]). Large, high-order rivers develop over geologic time frames, particularly in interior continental settings. Basin integration across complex landscapes for many large rivers has largely occurred during Neogene time (<23.5 million yr.; e.g., the Amazon, Colorado, Mississippi, Nile, Rhine, Rio Grande, and Yangtze rivers) [106, 107, 108, 109, 110, 111], although some river basins are far older (e.g., [112]).

River basins exist in a continuous state of development, expanding and capturing adjacent drainage basins at both gradual and punctuated rates (e.g., through periods of relatively rapid crustal deformation or vulcanism) and achieving transverse integration through at least five mechanisms (antecedence, superposition, anteposition, spillover, and piracy) [113, 114]. River channels respond to perturbations such as flood events by moving toward an equilibrium state that, due to recurring perturbation, may never be reached [101]. Although poorly synthesized, the three main tectonic interactions (convergence, divergence, and transform) are likely to produce different groundwater and surface water basin configurations. Aquifers that source rivers in these landscapes often are assumed to be constrained by surface catchment boundaries (e.g., [115]); however, such is not always the case (e.g., [116]). Regional climate also affects landscape geomorphology, and sometimes reciprocally influences tectonic processes (e.g., [117]). With tectonism as a driver, river course integration within a basin has been analogized to the process of organic evolution, in which drainage-head erosion allows individual tributaries to “competitively” integrate increasingly larger catchment areas, ultimately securing throughflow to the terminus [118].

Drainage network complexity is related to stream order, which increases when two channels of the same magnitude meet [26, 27]. Many of the world’s major rivers arise from discrete springs, spring-fed marshes, or groundwater-fed lakes. Headwater springs function as “zero order” streams and springbrooks are commonly regarded as first-order channels. Rivers with spring-sourced baseflows include the Amazon, Colorado, Mississippi, Rhine, Volga, Murray, and many others [119]. Springs often exhibit strikingly different temperature and geochemical characteristics from those in the adjacent higher order streams into which they are confluent (e.g., [120]). Their unique water quality may influence mainstream processes, such as imprinting among larval fish [121] but see [122]. The ecological transition from headwater springs into first-order streams is highly individualistic, often occurring at a chemically and thermally discrete distance from the source [123]. The quantity and quality of riverside or in-stream springs, as well as seasonal flow changes driven by precipitation, also can affect stream channel geomorphology through travertine deposition (e.g., [124]), persistence of in-channel woody vegetation [125], and river water quality (e.g., geothermal spring influences; e.g., [126]).

Stream order increases erratically downstream, with middle-order streams often having the greatest productivity and species richness. FRE lacustrine reaches can occur at any order within a basin (Figures 1 and 2) and lake limnological processes like thermal or chemical stratification can influence downstream flow and geochemistry. The largest rivers include the Amazon and Volga, which are regarded as 12th or 13th order streams. Such large rivers buffer temporal and spatial changes in water temperature, geochemistry, and the timing of dynamic equilibration.

River drainage networks often are subdivided into segments and reaches, which function as organizational units within FREs. Reaches lie within segments and are best distinguished geomorphologically on the basis of differences in parent rock geology, shoreline erodibility, slope (gradient), and thalweg position (e.g., [127]). River segments (sensu [128]) include one or more river reaches that are collectively subject to distinctive changes in one or more ecosystem characteristics (e.g., water temperature, geochemistry, suspended sediment load, or gamma diversity in an assemblage). Such changes often are introduced by a tributary, thereby affecting downstream FRE ecology [70, 71, 79, 129, 130].

The frequency, duration, magnitude, and timing of both high and low flows are critical determinants of FRE ecology. Discharge and flood frequency and magnitude are increasingly monitored and evaluated in relation to human activities within watershed. Magilligan [131] described variation in channel boundary shear stress and unit stream power on an array of stream channels across 2- to 500-year floods. She noted three-fold order of magnitude variation in flood power through the basin due to valley width, with broad, alluvial channels in wide valleys subject to lower flood power. In contrast, she reported increased stream power in narrow valleys with constrained channels, a pattern influenced both by basin size and by local controls. She also suggested that maximum flood impacts on channel geomorphology occur at discrete points within reaches. Such focal points are likely to shift over time, suggesting that drainage evolution may occur most intensively at the local scale. Antecedent events are critically important, as prior high flows exert long-lasting impacts on FRE structure and ecology (e.g., [34, 132, 133], many others). Detection of such events through dendrochronology is becoming more frequently used, helping to determine long-term drought frequency and duration, and providing insights into adaptive strategic options for FRE management in the face of climate change (e.g., [134, 135, 136, 137, 138, 139]).

The impacts of natural, regular, short-term stage fluctuations in rivers are generally poorly known, but are of great consequence in management of rivers impounded for hydroelectric power production (e.g., [140, 141]). Natural semi-daily tidal bores are common in the lowermost reaches of low-gradient rivers that reach the sea. Daily variation in flow stage in such settings may desiccate or freeze macrophytes, macroinvertebrate habitats and eggs, or interrupt aquatic and riparian faunal feeding and other behaviors, leading to reduced or fluctuating primary and secondary consumer production. Understanding the effects of natural fluctuating flows remains a potentially important topic for future natural and regulated FRE research.

River water quality varies across lithology, latitude, elevation, humidity province, season, and stream order within basins and among reaches, and springs, lakes, and glacial melt influence river waters. Water quality characteristics can transition markedly over stream order and are important determinants of macrophyte structure and composition, and life history and feeding guild distributions of aquatic macroinvertebrates, fish, and amphibians, in turn influencing food web linkage ([141], but see [142]) and riparian groundwater quality and quantity. Surface flow geochemical and sedimentological changes occur at tributary confluences (e.g., [129, 130]), abruptly as discontinuities, or more gradually and to a lesser extent in side channels and other shallow, low-velocity shoreline habitats. Limnologically-influenced water quality dominates lake-sourced rivers, but we know of little research on natural downriver responses to such alteration. River water generally trends toward a universal quality across stream order, generating relatively similar geochemistry among the world’s major rivers at their mouths; however, the contributions and evolution of FRE water quality depend in large measure on subbasin geology and the relative contributions of tributaries (e.g., [143, 144]), as well as anthropogenic impacts.

The erosion, and deposition of bed, suspended, dissolved loads, and flotsam sediments are related to watershed geology, aquifer properties, flow dynamics, channel configuration, and other factors like glacial influences [29, 32, 33, 59, 127, 145, 146]. Cumulatively, including anthropogenic materials, the world’s rivers deposit about 20 billion mt of solids into the sea each year [147]. Rather than being a sole function of basin area, this deposition is largely the result of discharge from many thousands of small basins (<10,000 km2) with relatively high-gradient rivers that mouth directly into the seas [148]. Large rivers deposit proportionally less sediment due to subaqueous storage in deltas.

Recent research and stewardship attention in fluvial geomorphology has shifted to temporally based based spatial scales of reaches in the continuum of alluvial to constrained rivers (e.g., [10, 53, 67, 68]). Alluvial reaches often have relatively uniform bed materials and channel landform configuration, and often are closer to equilibrium than geologically constrained channels. Models of sediment deposition and erosion are diverse (reviewed in [149], among many others) and can provide adequate two-dimensional prediction of suspended sediment transport through channels with varying bed roughness, channel steepness, and sediment transport. However, most rivers have insufficient historical flow, sedimentological, and hydrographic data to permit high-precision modeling [150]. Variation in turbulence, sheer stress, transport capacity, and bed and suspended sediment loading are likely to increase channel landform and FRE habitat diversity among reaches. In comparison with constrained reaches, seasonally dynamic alluvial reaches often support broader riparian zones, with increasing filtering, storage, and processing of matter from upstream and local sources [75, 76, 151].

Many watershed factors Influence FRE functions. Upland wildfire, forest pest insect outbreaks, coarse wood debris loading, and overgrazing can affect fluvial FRE sedimentation, geomorphology, and nutrient and nutrient transport (e.g., [152, 153]). Ice-related impacts on FREs in temperate and boreal streams involve ice formation, “shoving”, and black-ice melting, resulting in severe scouring of shorelines and bed surfaces, damming of channels, uplifting/redeposition of fine to coarse substrata (including boulders and coarse woody debris) [153, 154].

Fluvial climate is influenced by global- to local-scale conditions, the latter including nocturnal cool air subsidence and upriver mountain valley wind patterns (e.g., [155, 156]). However, few meso-scale studies of river influences on basin, reach, or local microclimate have been conducted. Riparian and in-stream interception of photosynthetically-active radiation (PAR) varies temporally and by reach in canyon-bound rivers, influencing in-canyon air temperature, relative humidity, and aquatic and riparian production (e.g., [80, 81]). Although not yet studied, variation in PAR flux also may influence erosion in temperate cliff-dominated channels through increased slope failure frequency, cliff retreat, and canyon landform evolution. At local, cross-sectional scales, discharge, cliff shading, and channel aspect influence microclimate and fluvial solar energy flux, in turn influencing FRE air temperature and relative humidity [157], which has been positively associated with temperate avian species diversity [158]. Thus, along with riparian soil ecology, fluvial microclimate may contribute to the biodiversity and productivity of riparian zones.

Habitat complexity at tributary confluences increases ecological productivity and biodiversity, and sustains habitat spatio-temporal connectivity [10, 39, 65, 70, 71159]. Tributary impacts on mainstream water quality are greatest when flows of the former exceed those of the latter; however, differences in biota may follow the opposite pattern. Aquatic macroinvertebrate assemblages in small, spring-fed tributaries may substantially differ from those in the large, adjacent mainstream (e.g., [130]). The WCM depicts the magnitude of tributary influences as dark triangles of varying size (Figure 2), but we note that large or influential tributaries can exist and interact with the mainstream anywhere in the watershed. The process domain concept (PDC) posits that variation among such geophysical processes at the reach scale shapes channel form, disturbance responses, and ecosystem structure and dynamics [68].

Many FRE physical models have focused on particular aspects of FRE channel development, form, and function. Leopold summarized much of his research on alluvial channels, through which he was able to describe the negative relationship between sinuosity and slope [160]. He lamented [33] that no comprehensive channel model adequately encompassed the self-regulating nature of alluvial channels. Existing alluvial channel studies and models have been criticized for generating: (1) untested qualitative unifying theories; (2) empirical and statistical analyses that support focused semi-quantitative models; (3) reductionist applications of Newtonian theory; and (4) theoretical resolutions of primary flow eqs. [95]. Based on Morisawa’s [34] articulation of the dynamic equilibrium concept, Nanson and Huang proposed focus on the least action principle, as represented in alluvial channels through the trend toward maximum flow efficiency [95]. While the WCM does not presume that all FRE component models (particularly ecological models) will have clear, predictive mathematical solutions, placing these many concepts in a logical order and visually representing them is an important step forward (Table 1; Figure 2).

2.3 Ecological conceptual models

Ecosystem ecology developed through the diverse contributions of Linnaeus, von Humboldt, Mobius, Darwin, Forbes, Warming, Cowles, Elton, and many others. Tansley defined the ecosystem concept as involving interrelated physical and biological elements and processes [161], and Lindeman [24] and Odum [25] initiated analysis of trophic-dynamic aspects of ecology (Table 1). FREs are primarily driven by physical factors, generating “bottom-up” ecosystem structure, with dependent biotic composition, structure, function, and trophic interactions (Figures 2 and 3). Lower stream order FRE changes often occur in a punctuated, stepped, or reach-bounded fashion as FREs receive tributary contributions of sediment, water temperature, and nutrients. As with fluvial water quality, the ecology of higher-order streams generally changes more gradually, both spatially and over time.

Figure 3.

(a) Expanded detail of foodweb linkages in FREs, contrasting allochthonous (uplands and tributary) vs. autochthonous (mainstream) ecosystem energy inputs with aquatic vs. riparian domain interactions. Arrows indicate common energy pathways among trophic levels in the four FRE arenas. Not all interactions occur in every FRE, and other trophic interactions not depicted here may exist in some FREs. (b) Differential spatial or functional change in reach-based FRE structure and function can occur in response to watershed changes. For example, upland fire can result in sediment, ash, and nutrient loading through tributaries, processes that may diminish FRE productivity and ecological role in the watershed. Similarly, reduction in precipitation or groundwater alterations through climate change or aquifer depletion may reduce mainstream and riparian function. Redrawn from. [1].

Hutchinson emphasized lake ecosystem limnology (e.g., [28]), while Hynes described stream limnology [36, 37], including the spatial scale and groundwater influences on the watershed, but with somewhat less attention to the FRE riparian domain. Hynsian (lotic) versus Hutchinsonian (lentic) emphases created long-standing differences in interpretation of the roles of habitat and biotic factors on FRE research [162]. Nonetheless, combining these lines of inquiry initiated a plethora of subsequent integrative research on FRE ecology, which continues today.

The most prominent post-Hynesian FRE conceptual advance was the river continuum concept (RCC) [6, 45, 163]. The RCC described a river ecosystem as “…a continuum of biotic adjustments and consistent pattern of loading, transport, utilization, and storage of organic matter along the (ir) length” ([6], 130). The RCC regarded “…the entire fluvial system as a continuously integrated series of physical gradients and associated biotic adjustments as the river flows from headwater to mouth”, with “…maintenance of longitudinal, lateral, and vertical pathways for biological, hydrological, and physical processes” ([99], 9–10); see [12]. The RCC lent support many patterns observed in studies of low-medium order streams and ichthyological studies, primarily in mesic regions [159, 164, 165, 166]. However, it has been criticized for not fully recognizing the roles of: (1) fluvial discontinuities; (2) groundwater and spring sourcing [167]; (3) river-sourced lakes, lentic zones, and productivity hot spots (e.g., [65]); (4) hyporheic refugia [99, 101, 168, 169]; (5) riparian ecology, except as subservient to the aquatic domain; (6) the role of temporal scale in FRE development and function, including dynamic seasonal and interannual geomorphic perturbation and adjustment [34]; (7) ephemeral and intermittent stream FRE ecology [170]; as well as (8) its applicability to higher order streams [171, 172].

Subsequent to the RCC, many FRE syntheses have been undertaken, including comprehensive edited volumes and reviews (including but not limited to [1, 23, 41, 42, 43, 44, 66, 69, 73, 75, 91, 99, 101, 102, 151, 164, 165, 166, 167, 171, 173, 174, 175, 176, 177, 178]). Below we briefly describe some of the major biologically-based conceptual components of the WCM.

Ward clarified four dimensions of spatial and temporal scale operating in most lotic ecosystems, including (1) the “longitudinal” dimension up- and downstream through rivers; (2) across-channel, riparian-aquatic domain interactions; (3) vertical interactions with hyporheic habitats and groundwater; and (4) a broad temporal dimension [102]. Dynamic interactions among these dimensions contribute to the individuality in the character of FREs. Focusing on riparian stage zones and related to Ward’s considerations, Nilsson and Svedmark [76] recognized four major processes or characteristics interactively functioning in FREs. (a) Flow regime (hydrographic) dynamics regulate FRE ecological and geomorphological processes (including riparian succession through Connell and Slayters’ three modes – facilitation, inhibition, and tolerance [179]). (b) The channel provides a corridor for organic and organic transport, primarily downstream but also upstream via anemochorous and zoochorous dispersal of propagules (Figure 2). (c) The riparian zone functions as a filter and boundary between upland and riverine processes (e.g., [85, 180]). (d) Many have recognized the high levels of biodiversity and ecological interactivity of FREs, related to elevated productivity and disturbance and high levels of habitat heterogeneity (e.g., [38, 39, 181]).

The flood pulse concept emphasized the importance of high flow pulses to FRE ecology by regularly restructuring channel and riparian landscapes [56, 57]. Regular seasonal flooding accounts for the state of suspended (or perpetual) succession in natural riparian vegetation zones, particularly in constrained channels [35], with riparian vegetation zonation in belts parallel to the mainstream [182] and with composition controlled by physiological and life history characteristics (e.g., riparian response guilds models [83, 183]).

Advances in nearshore marine ecology patch and disturbance dynamics concepts (e.g., [184, 185, 186, 187]) contributed to Thorp and DeLong’s [70] riverine productivity model (RPM). The RPM posited that production, as well as decomposition, recruitment, and other important river processes are related to niche diversity, occurring at specific points within the channel, such as at tributary confluences, along shorelines, or in specific depositional settings. Thus, FREs function as microhabitat mosaics.

The serial discontinuity concept (SDC) initially was developed to describe the impacts of impoundments on regulated rivers [52, 53], but also by extension to the roles and impacts of natural dams that form lacustrine reaches, and affect natural FRE channel geomorphology, flow, and population dynamics, both upstream and downstream. Lacustrine reaches can occur anywhere in a basin as a result of tectonism, lava dams, slope failure, or glacier development, and natural dams may persist for evolutionarily significant durations. The SDC posits that the location and size of a dam reset and influence downstream recovery of FRE characteristics through tributary contributions of flow, water quality, and biota [129], and through link discontinuity [70, 71] and network dynamics [77, 78, 79]. Examples of natural dams include Lake Victoria in Uganda, which formed as a result of tectonic rifting and interruption of flow in the Kagera and other Nile River headwater streams; Lago de Nicaragua (L. Cocibolca) in Central America, which formed as a result of tectonic uplift in the lower Tipitapa and San Juan River basins; and many basalt dams ni the southwestern USA [188, 189, 190, 191]. Six types of slope failure dams that can affect rivers are globally recognized, ranging from relatively common single events with partial valley impoundment to rare, simultaneous impoundment of multiple valleys that create multiple natural lakes [191] (e.g., [192, 193, 194]). Ice dam failure also is a well-known phenomenon (e.g., the collapse of Pleistocene Lake Missoula [195], and fjord ice dam failures [196]). These natural impoundments and their failures change downstream channel geomorphology, water quality and flow, hydrography, stage relations, velocity, habitat quality and distribution, and FRE biogeography.

The RCC did not adequately integrate the ecology of ephemeral and intermittent FREs or groundwater-surface water interactions [101]. Colloquially known as dry washes, arroyos, or wadis, ephemeral channels are extremely abundant, comprising far more than half of the global stream channel network [170], and are becoming increasingly abundant as humans and climate change dewater rivers (e.g., [197]). Flooding releases CO2 sequestered by seasonal or erratic burial of organic matter and invertebrates, such as leaf litter or clams [198]. Benthic invertebrates that shred, graze, or collect organic debris often are generally absent or rare in ephemeral FREs, reducing decomposition rates and transferring those functions to microbial and physical molar actions when the stream floods. Terrestrially, ephemeral versus intermittent riparian zones are bordered by distinctive suites of xeroriparian (dry riparian) to mesoriparian perennial plant species that provide cover and food resources [43]. Analysis of an ephemeral stream in Pakistan revealed deeply rooted woody perennial shrubs in the channel, and a bed dominated by weeds after winter rains, with drought-resistant species occurring on terraces [199]. Aquatic productivity and trophic energetics of arid-land ephemeral streams are reduced and interrupted during dry seasons (e.g., [92]), generating temporal discontinuities of stream processes. Nonetheless, ephemeral channels commonly provide essential wildlife habitat connectivity, and function as punctuated, rapidly changing biogeochemical reactors [93], and warranting additional research.

FRE productivity (P) and disturbance (D) intensity interactively influence aquatic and riparian domain biodiversity through habitat and resource availability, organism size distributions, niche specialization, assemblage composition, competition, and other factors. P and D are related to colonization (C) and extinction (E) processes in insular biogeographic models of species richness, with high levels of P related to C, and high levels of D related to E (Figure 2) [46, 47, 48, 200, 201, 202, 203, 204, 205, 206, 207]. Riparian and shallow aquatic domain P and D, as well as the depth, velocity, and transparency of water, and soil moisture and nutrient content are generally negatively related to stage. FRE riparian biodiversity tends to be maximal at intermediate levels of P and D, at intermediate flood return frequencies, and terrace stages with the maximal “ecological hospitability” to potential colonists. Our observations suggest that this pattern appears to be reversed in the aquatic domain, where shallow shorelines are most biologically diverse; however, more research is needed to understand this FRE “mirror effect” between the two domains. Both the intermediate disturbance hypothesis and insular biogeographic theory [46, 47, 48, 207] were developed to describe the biodiversity of sessile taxa, such as plants and corals, not vagile species like many larger macroinvertebrates, fish, and other vertebrates, which often actively cue on hydrographic disturbances (e.g., flood avoidance by belostomatid giant water bugs [208], or hydrograph-cued spawning among fish species). Such relationships may result, in some cases, in FRE riparian trophic cascades in which top predators can reciprocally influence channel geomorphology at reach scales [82]. Such cascades are regularly observed in fish-dominated ecosystems and in some low-order fishless systems, but often are limited in FRE aquatic domains by bottom-up physical processes (e.g., hydrology, sediment transport, ice impacts), with average sheer stress/unit area often negatively related to stream order [131]. Nonetheless, the commonly observed phenomenon of elevated FRE biodiversity, particularly at intermediate stream orders, is at least somewhat related to these coupled gradient interactions and convergent life history strategies.

FREs receive and transmit multidimensional exchanges of ecological matter and energy subsidies in the watershed. Muehlbauer et al. [94] reported that avenues of exchange may be relatively narrow (50% of exchanges occurred within 1.5 m of the stream edge), and 10% of the exchanges occurred 0.5 km into the adjacent uplands. Exchanges are reported to disproportionally influence primary producers and predators, potentially affecting both bottom-up and trophic cascades (e.g., [209]). FRE subsidy exchange involves five spatially directional processes over time, including (a) gravity-driven downslope flow and material allochthonous transport; (b) downstream flow; (c) river-to-uplands eolian and zoochorous transport; (d) lateral and downward surface to hyporheic transport; and (e) upwelling artesian groundwater influences. In addition, downstream main channel or tributary flooding in very low gradient reaches can initiate upstream-directed flow [102, 210]. Like channel geometry, these FRE ecological processes respond dynamically and temporally to climate and other factors, moving toward, but never achieving equilibrium in form, function, boundary conditions, matter transport, or trophic energy dynamics described for channel adjustment by [1, 34, 93].

In the riparian domain, FRE riparian vegetation is “…a complex of vegetation units along the river network that is functionally related to the other components of the fluvial system and surrounding area” [211, 212], which interacts in a reach- and segment-dependent fashion with the aquatic domain and its associated processes)” [213]. Like the aquatic domain, the riparian domain is interactively influenced by regional climate (e.g., [214, 215, 216]) through direct forcing effects on channel roughness, bedform morphology, and sediment transport during peak flows and seasonally changing rainfall and snowmelt [217, 218, 219], and by drought [220, 221]. Groundwater availability also can affect channel and floodplain stability and riparian vegetation [222, 223, 224, 225, 226]; (e.g., [227]), [228]. Climate influences on FRE groundwater vary spatiotemporally but can provide recharge that affects reach and segment scales through precipitation and infiltration, with potentially strong seasonal variation, as demonstrated through isotopic studies [229, 230, 231, 232, 233, 234, 235, 236].

Coupling the calculation of the standardized precipitation index (SPI; [237]) with the standardized groundwater level index (SGI; [238, 239]) can be used to relate precipitation to groundwater recharge [240, 241]. These metrics affect FRE riparian productivity [242, 243] (e.g., [244, 245, 246, 247, 248]) through groundwater recharge in relation to river stage [249], and are affected by air temperature [250, 251, 252] and extreme precipitation events [253, 254, 255], which in turn affect stream discharge [256], groundwater recharge and availability [257, 258], and phreatic zone and riparian rooting depth [259]. Inorganic sediment transport and turbidity generally (but not always) increase with stream order, reducing downstream PAR availability and 1° through 3° aquatic production [81] and strongly influencing aquatic macroinvertebrate feeding guild structure [6] and riparian nutrient availability. Complex trophic relationships can develop in riparian zones, directly and indirectly influencing primary producer structure and composition. For example, leaf-beetles, grasshoppers, beavers, and ungulates all can strongly influence riparian vegetation composition, structure, and decomposition/soil formation [260, 261, 262].

FRE biogeography involves colonization, recruitment, and population establishment overland by volant and other highly vagile species, as well as passive dispersal through gravity, aerial drift, or zoochorous transport of propagules through both overland and dendritic stream corridors (Figure 2). Regardless of the pathway, FRE population persistence and assemblage resilience are predicated on the ability of a species to remain in, or disperse-recover their position in the watershed. Therefore, persistence of all FRE species requires some form of upstream dispersal, with eviction or extirpation the inevitable consequence of failed in situ or headwater recruitment strategies. Larval aquatic macroinvertebrates may drift downstream, while adult aquatic insects often fly or are blown upstream as aerial drift. Dragonflies, salmonids, and many other fish taxa migrate upstream to spawn, against the dominant current direction, and some fish transport larval unionid mussel larvae upstream. Migratory western North American warblers and other passerine birds intensively use arid-land riparian habitat as stop-over habitat during migration [263, 264], a pattern not strongly evident in mesic eastern North America [265]. Although front-based migration also occurs among some shorebirds, many waterbird species follow FRE corridors, particularly through complex landscapes [80, 266]. In addition, many non-volant vertebrate species follow river corridors as dendritic pathways, although terrestrial faunal movements can be thwarted by steep cliffs, perilous crossings, and anthropogenic landscape interruptions [80].

The riparian plant niche construction perspective [83] and niche-box model (NBM) [84, 183] classified guilds of riparian plants in relation to similarities among life history traits. The NBM incorporates and compares autecological elements for each plant species to improve prediction of vegetation assemblage development in relation to hydrography and riparian conditions. While successfully grouping some species, the large amount of variation in the NBM multivariate plots is a reminder that life history strategies vary tremendously among species, variance that is highly adaptive but which does not readily lend itself to simple classification systems.

Trajectories of vegetation succession through modes of facilitation, inhibition, or tolerance [179] differ temporally among reaches, between humidity provinces and in relation to stream order, fluvial hydrodynamics (disturbance frequency and across stage), geomorphic setting, grain-size distribution, depth to water table [267], and biological effects, such as mycorrhizal succession [268], selective vertebrate [260] or invertebrate herbivory [261, 262], in relation to plant diseases and the presence of some bird species [204]. Surrounding upland assemblages also strongly interact with upper riparian terrace vegetation (e.g., [182, 269]). Due to increased riparian soil water availability and regular flood disturbance, FRE riparian vegetation structure is not well represented by the upland-centric Holdridge [270] global vegetation model [1].

Advertisement

3. Conclusions and research recommendations

Fluvial-riparian ecosystems are hierarchically and dynamically influenced by physical and biotic processes that vary spatially among reaches, over stream order and time within the watershed, approaching but rarely achieving equilibrium in channel geometry, fluid and matter transport, biotic composition, and ecosystem energy dynamics and structure. A wide array of conceptual physical and ecological models has described aspects of FRE ecology and responses to natural and anthropogenic perturbations. However, most models have focused on single or a reduced suite of variables at site-specific, within-reach, alluvial or constrained channels, other watershed scales, and most often on anthropogenically-altered streams. The WCM emphasizes the importance of understanding temporal and spatial scaling across the entire basin in natural systems to provide guidance for improving FRE stewardship.

Despite much progress, a wide array of important ecohydrological processes, questions, and issues remain to be addressed or more fully integrated. Not presented in prioritized order, this list of additional research topics includes but is not limited to: (1) corresponding convergence toward dynamic equilibrium of physical and biological processes; (2) extent of self-similarity among physical and biological processes across reach and stream order, space, and time; (3) groundwater-surface water interactions and connectivity under changing climates; (4) the significance, extent, and roles of groundwater and headwater springs as zero order streams in FRE ecology [120, 271]; (5) natural inter-relationships among lentic and downstream lotic reaches; (6) ephemeral stream ecosystem ecology; (7) the limiting effects of photosynthetically-active radiation in canyon-bound streams; (8) stream microclimate ecology; (9) the interactive effects of flooding, ice, and glacial effects in boreal and high elevation streams; (10) multidirectional subsidy and gene exchange in dendritic pathways; (11) the role of plant physiology in riparian vegetation zonation; (12) the significance of corridors, barriers/filter, and refugial biogeographic effects in dendritic river ecosystems [80]; (13) stream order-driven and cross-sectional spatial impacts on biodiversity; (14) population and successional models among FRE biota across trophic levels; (15) FRE ecosystem genetics and the evolution of endemism across latitude, longitude, and among tectonic landscapes; and (16) the role of noise on FRE faunal assemblages. Adequately examining and incorporating these and other topics will more fully expand the WCM model through future research and will enhance collaborative discussion among hydrogeological, ecological, and socio-cultural disciplines [272]. Such data and integration efforts will improve understanding, modeling, and stewardship of FREs at local, regional, and global spatial and temporal scales.

FREs are complex, continually changing, and vital to life on Earth. Although informative and elegant, all FRE models remain incomplete, and even the most comprehensive FRE conceptual syntheses fall short of adequately representing these remarkable, important, and dynamic ecosystems. Furthermore, discipline-based or regional specificity has often limited the applicability of some models (e.g., [273, 274, 275, 276, 277, 278]). Here we present and illustrate a synthesis of FRE knowledge through the WCM, and suggest topics for further investigation. However, FREs cannot be readily, adequately, or usefully reduced to a single suite of equations or simple illustrations. For example, non-Judeo-Christian-Islamic cultures commonly view rivers as living entities, supporting divine spirits, and essential to cultural well-being. Integrating indigenous traditional ecological concepts and knowledge into improved stewardship has rarely been achieved. We suggest that improved comprehension of FREs may require consideration of other socio-cultural dimensions. Enhanced understanding of the complex, multidimensional inter-relationships among Physical, biological, cultural, and socio-economic elements and processes within watersheds is essential to improving FRE stewardship and sustaining ecological functions vital to nature, human life, and societal integrity.

Advertisement

Acknowledgments

We thank the Springs Stewardship Institute for institutional support.

Advertisement

Conflicts of interest

We declare no conflicts of interest related to this manuscript.

References

  1. 1. Stevens LE, Johnson RR, Estes C. The watershed continuum: A conceptual fluvial-riparian ecosystem model. In: Johnson RR, Carothers SW, Finch DM, Kingsley KJ, Stanley JT, editors. Riparian Research and Management: Past, Present, Future. Ft. Collins: U.S. Department of Agriculture, Forest Service; 2020. pp. 80-137
  2. 2. Allan JD, Castillo MM, Capps KA. Fluvial Geomorphology: Stream Ecology. Cham: Springer; 2021. pp. 45-73
  3. 3. Tockner K, Ward JV, Edwards PJ, Kollmann J. Riverine landscapes: An introduction. Freshwater Biology. 2002;47:497-500
  4. 4. Chorley RJ, Kennedy BA. Physical Geography: A Systems Approach. London: Prentice-Hall; 1971
  5. 5. Huggett R. A history of the systems approach in geomorphology Une histoire de l’ approche systémique en géomorphologie. Géomorphologie Reli. Process. Environ. 2007;2:145-158. DOI: 10.4000/geomorphologie.1031
  6. 6. Vannote RL, Minshall GW, Cummins KW, Sedell JR, Cushing CE. The river continuum concept. Canadian Journal of Fisheries and Aquatic Sciences. 1980;37:130-137
  7. 7. Ward ND, Sawakuchi HO, Richey JE. The Amazon River’s ecosystem: Where land meets the sea. Eos. 2018;99. DOI: 10.1029/2018EO088573
  8. 8. Shiklomanov I. World fresh water resources. In: Gleick PH, editor. Water in Crisis: A Guide to the World’s Fresh Water Resources. New York: Oxford University Press; 1993. pp. 13-24
  9. 9. Behrensmeyer AK, Damuth JA, Dimichele WA, Potts R, Sues H-D, Wing SL. Terrestrial Ecosystems through Time: Evolutionary Paleoecology of Terrestrial Plants and Animals. Chicago, IL: University of Chicago Press; 1992. pp. 205-325
  10. 10. Sabo JL, Hagen EM. A network theory for resource exchange between rivers and their watersheds. Water Resources Research. 2012;48:W04515
  11. 11. Stanford JA, Ward JV. An ecosystem perspective of alluvial rivers: Connectivity and the hyporheic corridor. Journal of North American Benthological Society. 1993;12(1):48-60. DOI: 10.2307/1467685
  12. 12. Piégay H. System approaches in fluvial geomorphology. In: Kondolf H, Piegay GM, editors. Tools in Fluvial Geomorphology. West Sussex, UK: Wiley; 2003. pp. 79-101
  13. 13. Verry ES, Dolloff CA, Manning ME. Riparian ecotone: A functional definition and delineation for resource assessment. Water, Air, Soil Pollution Focus. 2004;4(1):67-94. DOI: 10.1023/B:WAFO.0000012825.77300.08
  14. 14. Zarnott DH, Tavares VEQ , Giardina C, Alba JMF. The influence of riparian areas on direct surface runoff of precipitation events. Floresta e Ambient. 2020;27(4)
  15. 15. Vogel HF, Zawadzki CH, Metri R. Florestas ripárias: importância e principais ameaças. Rev. Saúde e Biol. 2009;4(1):24-30
  16. 16. Pusey B, Arthington A. Importance of the riparian zone to the conservation and management of freshwater fish: a review. Marine Freshwater Research. 2003;54:1-6
  17. 17. Meehan WR, Swanson FJ, Sedell JR. lnfluences of riparian vegetation on aquatic ecosystems with particular references to salmonid fishes and their food supply. In: Johnson RR, Jones DA, editors. Importance Preservation and Management of Riparian Habitat: A Symposium. Rocky Mountain Forest and Range Experiment Station. Fort Collins, CO, USA: USDA Forest Service; 1977. pp. 137-145
  18. 18. Swanson FJ, Gregory SV, Sedell JR, Campbell AG. Land-Water Interaction: The Riparian Zone. In: Analysis of Coniferous Forest Ecosystems in the Western United States US/IBP Synthesis Series 14. Stroudsburg, PA, USA: Hutchinson Ross Publishing Co; 1982. pp. 267-291
  19. 19. Dwire KA, Mellmann-Brown S, Gurrieri JT. Potential effects of climate change on riparian areas, wetlands, and groundwater-dependent ecosystems in the Blue Mountains, Oregon, USA. Climate. 2018;10:44-52
  20. 20. Gregory SV, Swanson FJ, McKee WA, Cummins KW. An ecosystem perspective of riparian zones. BioScience. 1991;41:540-551
  21. 21. Naiman RJ, Décamps H. The ecology of interfaces: Riparian zones. Annual Review of Ecology and Systematics. 1997;28:621-658
  22. 22. Naiman RJ, Melillo JM, Lock MA, Ford TE, Reice SR. Longitudinal patterns of ecosystem processes and community structure in a subarctic river continuum. Ecology. 1987;68:1139-1156
  23. 23. Naiman RJ, DéCamps H, McClaim ME. Riparia: Ecology, Conservation, and Management of Streamside Communities. New York: Wiley; 2005
  24. 24. Lindeman RL. The tropic-dynamic aspect of ecology. Ecology. 1942;23:399-417
  25. 25. Odum HT. Trophic structure and productivity of Silver Springs, Florida. Ecological Monographs. 1957;27:55-112
  26. 26. Horton RE. Erosional development of streams and their drainage basins: Hydrophysical approach to quantitative morphology. Bulletin of the Geological Society of America. 1945;56:275-370
  27. 27. Strahler AN. Quantitative analysis of watershed geomorphology. Transactions of the American Geophysical Union. 1957;38:913-920
  28. 28. Hutchinson GE. A Treatise on Limnology, Volume II. Introduction to Lake Biology and the Limnoplankton. New York: John Wiley & Sons; 1967
  29. 29. Leopold LB, Maddock R Jr. The Hydraulic Geometry of Stream Channels and Some Physiographic Implications. Washington: U.S. Geological Survey; 1953
  30. 30. Schumm SA. Geomorphic thresholds and complex response of drainage systems. In: Morisawa M, editor. Fluvial Geomorphology. Binghamton: SUNY Binghamton Publication in Geomorphology; 1973. pp. 299-310
  31. 31. Schumm SA. Patterns of alluvial rivers. Annual Review of Earth and Planetary Sciences. 1985;13:5-27
  32. 32. Schumm SA, Chorley RJ, Sugden DE. Geomorphology. New York: Methuen; 1985. p. 413
  33. 33. Leopold LB. A View of the River. Cambridge: Harvard University Press; 1994
  34. 34. Morisawa M. Streams: Their Dynamics and Morphology. New York: McGraw-Hill; 1968
  35. 35. Campbell CJ, Green W. Perpetual succession of stream-channel vegetation in a semiarid region. Journal of the Arizona Academy of Sciences. 1968;5:86-98
  36. 36. Hynes HBN. The Ecology of Running Waters. Toronto: University of Toronto Press; 1970
  37. 37. Hynes HBN. The stream and its valley. Verhandlunger der Internationalen Vereinigung für Theoretische und Angewandte Limnologie. 1975;91:1-15
  38. 38. Carothers SW, Johnson RR, Aitchison SW. Population structure and social organization in southwestern riparian birds. American Zoologist. 1974;14:97-108
  39. 39. Naiman RJ, DéCamps H, Pollock MM. The role of riparian corridors in maintaining regional biodiversity. Ecological Applications. 1993;3:209-212
  40. 40. Nilsson C, Grelsson G, Johansson ME, Sperens U. Patterns of plant species richness along riverbanks. Ecology. 1989;70:77-84
  41. 41. Johnson RR, Jones DA. Importance, Preservation and Management of Riparian Habitat: A Symposium. Fort Collins, CO; USA: Department of Agriculture Forest Service Rocky Mountain Research Station; 1977
  42. 42. Johnson RR, Carothers SW, Finch DM, Kingsley KJ, Stanley JT, editors. Riparian Research and Management: Past, Present, Future. Ft. Collins: Department of Agriculture, Forest Service; 2019-2020
  43. 43. Johnson RR, Carothers SW, Simpson JM. A riparian classification system. In: Warner RE, Hendrix KM, editors. California Riparian Systems. Berkeley, CA: University of California Press; 1984. pp. 375-382
  44. 44. Johnson RR, Siebel CD, Patton DR, Ffolliott PF, Hamre RH. Riparian ecosystems and their management: Reconciling conflicting uses. In: First North American Riparian Conference. Tucson, AZ: U.S. Department of Agriculture Forest Service Rocky Mountain Research Station; 1985
  45. 45. Cummins KW. The ecology of running water: Theory and practice. In: Baker DB, Prater BL, editors. Proceedings of the Sandusky River Basin Symposium. Washington, DC: U.S. Government Printing Office; 1976. pp. 277-293
  46. 46. Connell JH. Diversity in tropical rain forests and coral reefs. Science. 1978;199:1302-1310
  47. 47. Huston MA. A general hypothesis of species diversity. American Naturalist. 1979;113:81-101
  48. 48. Huston MA. Biological Diversity: The Coexistence of Species on Changing Landscapes. Cambridge: Cambridge University Press; 1994
  49. 49. Pollock MM, Naiman RJ, Hanley TA. Plant species richness in riparian wetlands—A test of biodiversity theory. Ecology. 1998;79:94-105
  50. 50. Newbold JD, Elwood JW, O’Neill RV, Winkle WV. Measuring nutrient spiraling in streams. Canadian Journal of Fisheries and Aquatic Sciences. 1981;38:860-863
  51. 51. Ensign SH, Doyle MW. Nutrient spiraling in streams and river networks. Journal of Geophysical Research. 2006;111:G04009
  52. 52. Ward JV, Stanford JA. The serial discontinuity concept of river ecosystems. In: Fontaine TD, Bartell SM, editors. Dynamics of Lotic Ecosystems. Ann Arbor: Science Publications; 1983. pp. 29-43
  53. 53. Ward JV, Stanford JA. The serial discontinuity concept: Extending the model to floodplain rivers. Regulated Rivers: Research & Management. 1995;10:159-168
  54. 54. Décamps H, Fortuné M, Gazelle F, Pautou G. Historical influence of man in the riparian dynamics of a fluvial landscape. Landscape Ecology. 1988;1:163-173
  55. 55. Petts GE, Möller H, Roux AL, editors. Historical Change in Large Alluvial Rivers. Chichester: John Wiley; 1989
  56. 56. Junk WJ, Bayley PB, Sparks RE. The flood pulse concept in river continuum systems. Canadian Special Publication on Fisheries and Aquatic Sciences. 1989;106:89-109
  57. 57. Tockner K, Malard F, Ward JV. An extension of the flood pulse concept. Hydrological Processes. 2000;14:2861-2883
  58. 58. Stella JC, Hayden M, Battles JJ, Fremier A. The role of abandoned channels as refugia for sustaining pioneer riparian forest ecosystems. Ecosystems. 2011;14:776-790
  59. 59. Tinkler KJ, Wohl EE. Rivers over rock: Fluvial processes in bedrock channels. American Geophysical Union Geophysical Monograph Series. 1998;107:336
  60. 60. Rosgen DL. Applied River Morphology. Pagosa: Wildland Hydrology; 1996
  61. 61. Rosgen DL. Discussion of Critical evaluation of how the Rosgen classification and associated natural channel design methods fail to integrate and quantify fluvial processes and channel responses. Journal of the American Water Resources Association. 2008;44:782-792
  62. 62. Griffiths RE, Anderson DE, Springer AE. The morphology and hydrology of small spring-dominated channels. Geomorphology. 2008;102:511-521
  63. 63. Buffington JM, Montgomery DR. Geomorphic classification of rivers. In: Shroder J, Wohl EE, editors. Treatise on Geomorphology. San Diego, CA: Academic Press; 2013. pp. 730-767
  64. 64. McManamay RA, DeRolph CR. A stream classification system for the coterminous United States. Scientific Data. 2019;190:17
  65. 65. Thorp JH, DeLong MD. The riverine productivity model: An heuristic view of carbon sources and organic processing in large river ecosystems. Oikos. 1994;70:305-308
  66. 66. Thorp JH, Thoms MC, Delong MD, editors. The Riverine Ecosystem Synthesis: Toward Conceptual Cohesiveness in River Science. London: Academic Press; 2008
  67. 67. Montgomery DR. Process domains and the river continuum. Journal of the American Water Resources Association. 1999;35:397-410
  68. 68. Bellmore JR, Baxter CV. Effects of geomorphic process domains on river ecosystems: A comparison of floodplain and confined valley segments. River Research and Applications. 2014;30:617-630
  69. 69. Fisher SG, Grimm NB, Marti E, Holmes RM, Jones JB. Material spiraling in stream corridors: A telescoping ecosystem model. Ecosystems. 1998;1:19-34
  70. 70. Rice SP, Greenwood MT, Joyce CG. Tributaries, sediment sources and the longitudinal organization of macroinvertebrate fauna along river systems. Canadian Journal of Fisheries and Aquatic Science. 2001;58:824-840
  71. 71. Rice SP, Kiffney P, Greene C, Pess GR. The ecological importance of tributaries and confluences. In: Rice SP, Roy AG, Rhoads BL, editors. River Confluences, Tributaries and the Fluvial Network. Hoboken, New Jersey: John Wiley & Sons; 2008. pp. 209-242
  72. 72. Bennett SJ, Simon A. Riparian Vegetation and Fluvial Geomorphology. Washington, DC: American Geophysical Union; 2013
  73. 73. Committee on Riparian Zone Functioning and Strategies for Management (CRZFSM). Riparian Areas: Functions and Strategies for Management. Washington, DC: National Academy Press; 2002
  74. 74. Hupp CR, Osterkamp WR. Bottomland vegetation distribution along Passage Creek, Virginia, in relation to fluvial landforms. Ecology. 1985;66:670-681
  75. 75. Hupp CR, Osterkamp WR. Riparian vegetation and fluvial geomorphic processes. Geomorphology. 1996;14:277-295
  76. 76. Nilsson C, Svedmark M. Basic principles and ecological consequences of changing water regimes: Riparian plant communities. Environmental Management. 2002;30:468-480
  77. 77. Benda L, Andras K, Miller D, Bigelow P. Confluence effects in rivers: Interactions of basin scale, network geometry, and disturbance regimes. Water Resources Research. 2004;40:W05402
  78. 78. Benda L, Poff NL, Miller D, Dunne T, Reeves G, Pess G, et al. The network dynamics hypothesis: How channel networks structure riverine habitats. Bioscience. 2004;4:413-427
  79. 79. Clay PA, Muehlbauer JD, Doyle MW. Effect of tributary and braided confluences on aquatic macroinvertebrate communities and geomorphology in an alpine river watershed. Freshwater Science. 2015;34:845-856
  80. 80. Stevens LE. The biogeographic significance of a large, deep canyon: Grand Canyon of the Colorado River, southwestern USA. In: Stevens LE, editor. Global Advances in Biogeography. Rijeka: InTech Publications; 2012. pp. 169-208
  81. 81. Yard MD, Bennett GE, Mietz SN, Coggins LN, Stevens LE, Hueftle S, et al. Influence of topographic complexity on solar insolation estimates for the Colorado River, Grand Canyon, AZ. Ecological Modeling. 2005;183:157-172
  82. 82. Beschta RL, Ripple WJ. River channel dynamics following extirpation of wolves in northwestern Yellowstone National Park, USA. Earth Surface Processes and Landforms. 2006;31:1525-1539
  83. 83. Corenblit D, Seiger J, Gurnell AM, Naiman RJ. Plants intertwine fluvial landform dynamics with ecological succession and natural selection: A niche construction perspective for riparian systems. Global Ecology and Biogeography. 2009;18:507-520
  84. 84. Merritt DM, Scott ML, Poff NL, Auble GT, Lytle DA. Theory, methods, and tools for determining environmental flows for riparian vegetation: Riparian vegetation-flow response guilds. Freshwater Biology. 2010;55:206-225
  85. 85. Bouwman AF, Giffioen MFP, Hefting MM, Middelburg JJ, Middelkoop H, Slomp CP. Nutrient dynamics, transfer and retention along the aquatic continuum from land to ocean: Toward integration of ecological and biogeochemical models. Biogeosciences. 2013;10:1-23
  86. 86. Jacobs SM, Bechtold JS, Biggs HC, Grimm NB, Lorentz S, McClain MR, et al. Nutrient vectors and riparian processing: A review with special reference to African semiarid savanna ecosystems. Ecosystems. 2007;10:1231-1249
  87. 87. Meybeck M. Carbon, nitrogen and phosphorous transport by world rivers. American Journal of Science. 1982;282:401-450
  88. 88. Bennett MG. Effects of Flow Regime on Fishes and Fisheries: From Life Histories to Ecosystem Services. Carbondale, IL: Southern Illinois University; 2015
  89. 89. Grime JP. Evidence for the existence of three primary strategies in plants and its relevance to ecological and evolutionary theory. American Naturalist. 1977;111:1169-1194
  90. 90. Mims MC, Olden JD, Shattuck Z, Poff NL. Life history trait diversity of native freshwater fishes in North America. Ecology of Freshwater Fish. 2010;19:390-400
  91. 91. Humphries P, Keckeis H, Finlayson B. The river wave concept: Integrating river ecosystem models. BioScience. 2014;64:870-882
  92. 92. Jenkins KM, Boulton AJ. Connectivity in a dryland river: Short-term aquatic microinvertebrate recruitment following floodplain inundation. Ecology. 2003;84:2708-2723
  93. 93. Larned ST, Datry T, Arscott DB, Tockner K. Emerging concepts in temporary-river ecology. Freshwater Biology. 2010;55:717-738
  94. 94. Muehlbauer JD, Collins SF, Doyle MW, Tockner K. How wide is a stream? Spatial extent of the potential “stream signature” in terrestrial food webs using meta-analysis. Ecology. 2014;95:44-55
  95. 95. Nanson GC, Huang H. A philosophy of rivers: Equilibrium states, channel evolution, teleomatic change and least action principle. Geomorphology. 2018;302:3-19
  96. 96. Springer AE, Stevens LE. Spheres of discharge of springs. Hydrogeology Journal. 2009;17:83-93
  97. 97. Stevens LE, Schenk ER, Springer AE. Springs ecosystem classification. Ecological Applications. 2021;31:e2218
  98. 98. Cid N, Erős T, Heino J, Singer G, Thibault D. From meta-system theory to the sustainable management of rivers in the Anthropocene. Frontiers in Ecology and the Environment. 2022;20:49-57
  99. 99. Annear T, Chisholm I, Beecher H, Locke A, Aarrestad P, Coomer C, Estes C, Hunt J, Jacobson R, Jöbsis G, Kauffman J, Marshall J, Mayes K, Smith G, Wentworth R, Stalnaker C. Instream Flows for Riverine Resource Stewardship, Revised ed. Cheyenne, WY: Instream Flow Council; 2004. p. 268
  100. 100. Annear T, Lobb D, Coomer C, Woythal M, Hendry C, Estes C, Williams K. International Instream Flow Program Initiative, A Status Report of State and Provincial Fish and Wildlife Agency Instream Flow Activities and Strategies for the Future, Final Report for Multi-State Conservation Grant Project WY M-7-T. Cheyenne, WY: Instream Flow Council; 2009
  101. 101. Stanford JA. Rivers in the landscape: Introduction to the special issue on riparian and groundwater ecology. Freshwater Biology. 1998;40:402-406
  102. 102. Ward JV. The four-dimensional nature of lotic ecosystems. Journal of the North American Benthological Society. 1989;8:2-8
  103. 103. Reid AJ, Carlson AK, Creek IF, et al. Emerging threats and persistent conservation challenges for freshwater biodiversity. Biological Reviews. 2019;94:849-873
  104. 104. Holbrook J, Schumm SA. Geomorphic and sedimentary response of rivers to tectonic deformation: A brief review and critique of a tool for recognizing subtle epeirogenic deformation in modern and ancient settings. Tectonophysics. 1999;305:287-306
  105. 105. Duvall A, Kirby E, Burbank D. Tectonic and lithologic controls on bedrock channel profiles and processes in coastal California. Journal of Geophysical Research. 2004;109:F03002
  106. 106. Böhme M, Aiglstorfer M, Uhl D, Kullmer O. The antiquity of the Rhine River: Stratigraphic coverage of the Dinotherien-sande (Eppelsheim Formation) of the Mainz Basin (Germany). PLoS ONE. 2012. DOI: 10.1371/journal.pone.0036817
  107. 107. Figueiredo JP, Hoorn C, van der Ven P, Soards E. Late Miocene onset of the Amazon River and Amazon deep-sea fan: Evidence from the Foz do Amazonas Basin. Geology. 2009;37:619-622
  108. 108. Mack GH, Seager WR, et al. Pliocene and Quaternary history of the Rio Grande, the axial river of the southern Rio Grande rift. New Mexico Earth-Science Reviews. 2006;79:141-162
  109. 109. Wickert AD, Mitrovica JX, Williams C, Anderson RS. Gradual demise of a thin southern Laurentide ice sheet recorded by Mississippi drainage. Nature. 2013;502:668-671
  110. 110. Woodward JC, Macklin MG, Krom MD, Williams MAJ. The Nile: Evolution, Quaternary river environments and material fluxes. In: Gupta A, editor. Large Rivers: Geomorphology and Management. London: John Wiley & Sons, Ltd; 2007. pp. 261-292
  111. 111. Zheng H. Birth of the Yangtze River: Age and tectonic geomorphic implications. National Science Review. 2015;2:438-453
  112. 112. Heinsalu H, Bednarczyk W. Tremadoc of the East European Platform: Lithofacies and paleogeography. Proceedings of the Estonian Academy of Sciences Geo!ogy. 1997;46:59-74
  113. 113. Larson PH, Dorn RI, Gootee BF, Seong YB. Drainage integration in extensional tectonic settings. Geomorphology. 2022;399:108082
  114. 114. Ranney W. Carving Grand Canyon: Evidence, Theories, and Mystery. 2nd ed. Grand Canyon: Grand Canyon Association; 2012
  115. 115. Springer AE, Boldt EM, Junghans KM. Local vs. regional groundwater flow delineationfrom stable isotopes at western North America springs. Groundwater. 2016;55:100-109
  116. 116. Winograd IJ, Thordarson W. Hydrogeologic and Hydrochemical Framework, South-central Great Basin, Nevada-California, with Special Reference to the Nevada Test Site. Washington, DC: U.S. Geological Survey; 1975
  117. 117. Montgomery DR, Balco G, Willett SD. Climate, tectonics, and the morphology of the Andes. Geology. 2001;29:579-582
  118. 118. Lucchitta I. History of the Grand Canyon and of the Colorado River in Arizona. In: Beus SS, Morales M, editors. Grand Canyon Geology. Oxford: Oxford University Press; 1990. pp. 311-332
  119. 119. Junghans K. Sources of Perennial Stream Flow in Arizona, USA. Flagstaff: Northern Arizona University; 2016
  120. 120. Lowe WH, Likens GE. Moving headwater streams to the head of the class. BioScience. 2005;55:196-197
  121. 121. Teears TD. The Effect of Water Quality on the Survival and Fitness of Brook Trout (Salvelinus fontinalis) Eggs, Alevins and Fry in Aquaculture and Deep Springs along the South River in Wayesboro, VA. Harrisonburg: James Madison University; 2016
  122. 122. Muller-Schwarze D. Chemical Ecology of Vertebrates. Cambridge: Cambridge University Press; 2006
  123. 123. Morrison RR, Stone MC, Sada DW. Environmental response of a desert springbrook to incremental discharge reductions, Death Valley National Park, California, USA. Journal of Arid Environments. 2013;99:5-13
  124. 124. Cantonati M, Segadelli S, Ogata K, Tran H, Sanders D, Gerecke R, et al. A global review on ambient limestone-precipitating springs (LPS): Hydrogeological setting, ecology, and conservation. Science of the Total Environment. 2016;15:568-624
  125. 125. Stevens LE, Jenness J, Ledbetter JD. Springs and springs-dependent taxa in the Colorado River Basin, southwestern North America: Geography, ecology, and human impacts. Water. 2020;12:1501
  126. 126. Ruzo A. The Boiling River: Adventure and Discovery in the Amazon. New York: TED Books; 2016
  127. 127. Schmidt JC, Graf JB. Aggradation and Degradation of Alluvial Sand Deposits, 1965-1986, Colorado River, Grand Canyon National Park, Arizona. Washington, DC: United States Geological Survey; 1990
  128. 128. Stevens LE, Shannon JP, Blinn DW. Benthic ecology of the Colorado River in Grand Canyon: Dam and geomorphic influences. Regulated Rivers: Research & Management. 1997;13:129-149
  129. 129. Dye A. Contribution of Unregulated Tributaries to the Ecological Functioning of the Main Channel of Rivers. Snowy River Recovery: Snowy Flow Response Monitoring and Modeling. Sydney: New South Wales Office of Water; 2010
  130. 130. Stevens LE, Holway JH, Ellsworth C. Benthic discontinuity between an unregulated tributary and the dam-controlled Colorado River, Grand Canyon, Arizona, USA. Annals of Ecology and Environmental Science. 2020;4:33-48
  131. 131. Magilligan FJ. Thresholds and the spatial variability of flood power during extreme floods. Geomorphology. 1992;5:373-390
  132. 132. Foster DR, Knight DH, Franklin JF. Landscape patterns and legacies resulting from large, infrequent disturbances. Ecosystems. 1998;1:497-510
  133. 133. Parsons M, McLoughlin CA, Kotschy KA, Rogers KH, Rountree MW. The effects of extreme floods on the biophysical heterogeneity of river landscapes. Frontiers in Ecology and the Environment. 2005;3:487-494
  134. 134. Akkemik U, D’Arrigo R, Cherubini P, Kose N, Jacoby GC. Tree-ring reconstructions of precipitation and streamflow for northwestern Turkey. International Journal of Climatology. 2004;28:173-183
  135. 135. Case RA, MacDonald GM. Tree ring reconstructions of streamflow for three Canadian prairie rivers. Journal of the American Water Resouces Association. 2003;39:703-716
  136. 136. D’Arrigo R, Abram N, Ummenhofer C, Palmer J, Mudelsee M. Reconstructed Streamflow for Citarum River. Java, Indonesia: Climate Dynamics; 2009
  137. 137. Melo DM, Woodhouse CA, Morino K. Dendrochronology and links to streamflow. Journal of Hydrology. 2012;412-413:200-209
  138. 138. Therrell MD, Bialecki MB. A multi-century tree-ring record of spring flooding on the Mississippi River. Journal of Hydrology. 2015;529:490-498
  139. 139. US Bureau of Reclamation. Colorado River Basin Water Supply and Demand Study. Washington, DC: US Department of the Interior; 2012
  140. 140. Kennedy TA, Muehlbauer JD, Yackulic CB, Lytle DA, Miller SW, Dibble KL, et al. Flow management for hydropower extirpates aquatic insects, undermining river food webs. BioScience. 2016;66:561-575
  141. 141. Merritt RW, Cummins KW, Berg MB. An Introduction to the Aquatic Insects of North America. 4th ed. Dubuque: Kendall-Hunt; 2008
  142. 142. Heino J. A comparative analysis reveals weak relationships between ecological factors and beta diversity of stream insect metacommunities at two spatial scales. Ecology and Evolution. 2015;5:1235-1248
  143. 143. Giller S, Malmqvist B. The Biology of Streams and Rivers. Oxford: Oxford University Press; 1998
  144. 144. Kabede S, Travi Y, Alemayehu T, Ayenew T. Groundwater recharge, circulation and geochemical evolution in the source region of the Blue Nile River, Ethiopia. Applied Geochemistry. 2005;20:1658-1676
  145. 145. Hey RD, Bathurst JC, Thorne CR. Gravel-bed Rivers: Fluvial Processes, Engineering, and Management. New York: John Wiley & Sons; 1982
  146. 146. Vogel S. Life in Moving Fluids: The Physical Biology of Flow. Boston: Willard Grant Press; 1981
  147. 147. Gray JR, Simões FJM. Estimating sediment discharge. In: Garcia M, editor. Sedimentation Engineering – Processes, Measurements, Modeling, and Practice. Reston, Virginia: American Society of Civil Engineers; 2008. pp. 1065-1086
  148. 148. Milliman JD, Syvitski JPM. Geomorphic/tectonic control of sediment discharge to the ocean: The importance of small mountainous rivers. Journal of Geology. 1992;100:525-544
  149. 149. Merritt WS, Letcher RA, Jakeman AJ. A review of erosion and sediment transport models. Environmental Modeling & Software. 2003;18:761-799
  150. 150. Alley WM, Evenson EJ, Barber NL, Bruce BW, Dennehy KF, Freeman MC, et al. Progress toward Establishing a National Assessment of Water Availability and Use. Washington, DC: U.S. Geological Survey; 2013
  151. 151. Naiman RJ, Bilby RE, editors. River Ecology and Management. New York: Springer-Verlag; 1998
  152. 152. Bormann FH, Likens GE. Pattern and Process in a Forested Ecosystem. New York, NY: Springer-Verlag; 1979
  153. 153. Stevens V. The Ecological Role of Coarse Woody Debris: An Overview of the Ecological Importance of CWD in BC Forests. Victoria: British Columbia Ministry of Forests; 1997
  154. 154. Prowse TD, Culp JM. Ice breakup: A neglected factor in river ecology. Canadian Journal of Civil Engineering. 2003;30:128-144
  155. 155. Draught AE, Rubin DM. Measurements of Wind, Eolian sand Transport, and Precipitation in the Colorado River Corridor. Grand Canyon, Arizona: U.S. Geological Survey; 2006
  156. 156. Whiteman CD. Observations of thermally developed wind systems in mountainous terrain. Meteorological Monographs. 1990;45:5-42
  157. 157. Stanitski-Martin D. Seasonal Energy Balance Relationships over the Colorado River and Adjacent Riparian Habitat: Glen Canyon, Arizona. Tempe: Arizona State University; 1999
  158. 158. Selwood K, Clark RH, McGeoch MA, MacNally R. Green tongues into the arid zone: River floodplains extend the distribution of terrestrial bird species. Ecosystems. 2016;20:745-756
  159. 159. Merriam G. Connectivity: A fundamental ecological characteristic of landscape pattern. In: Brandt J, Agger P, editors. Proceedings of the First International Seminar on Methodology in Landscape Ecology Research and Planning. Vol. 1. Roskikde: Roskilde Universitessforlag GeoRue; 1984. pp. 5-15
  160. 160. Leopold LB, Wolman MG, Miller JP. Fluvial Processes in Geomorphology. San Francisco: Freeman; 1964
  161. 161. Tansley AG. The use and abuse of vegetational concepts and terms. Ecology. 1935;16:284-307
  162. 162. Wurtsbaugh WA, Heredia NA, Laub BG, Meredith CS, Mohn HE, Null SE, et al. Approaches for studying fish production: Do river and lake researchers have different perspectives? Canadian Journal of Fisheries and Aquatic Sciences. 2015;72:149-160
  163. 163. Meehan WR, Swanson FJ, Sedell JR. Influences of riparian vegetation on aquatic ecosystems with particular reference to salmonid fishes and their food supply. In: Johnson RR, Jones DA, editors. Importance, Preservation and Management of Riparian Habitat: A Symposium. Fort Collins, CO: U.S. Department of Agriculture Forest Service Rocky Mountain Research Station; 1977. pp. 137-145
  164. 164. Minshall GW. Stream ecosystem theory: A global perspective. Journal of the North American Benthological Society. 1988;7:263-288
  165. 165. Minshall GW, Peterson RC, Cummins KW, Bott TL, Sedell JR, Cushing CE, et al. Interbiome comparison of stream ecosystem dynamics. Ecological Monographs. 1983;53:1-25
  166. 166. Minshall GW, Cummins KW, Petersen RC, Cushing CE, Bruns DA, Sedell JR, et al. Developments in stream ecosystem theory. Canadian Journal of Fisheries and Aquatic Sciences. 1985;42:1045-1055
  167. 167. National Research Council. Riparian Areas: Functions and Strategies for Management. Washington, DC: National Academy Press; 2002
  168. 168. Palmer MA, Bely AE, Berg KE. Response of invertebrates to lotic disturbance: A test of the hyporheic refuge hypothesis. Oecologia. 1992;89:182-194
  169. 169. Stanford JA, Ward JV. The hyporheic habitat of river ecosystems. Nature. 1988;355:64-66
  170. 170. Datry T, Larned ST, Tocknerr K. Intermittent rivers: A challenge for freshwater ecology. BioScience. 2014;64:229-235
  171. 171. Sedell JR, Richey JE, Swanson RJ. The river continuum concept: A basis for the expected ecosystem behavior of very large rivers? Canadian Special Publications in Fisheries and Aquatic Sciences. 1989;106:49-55
  172. 172. Johnson BL, Richardson WB, Naimo TJ. Past, present, and future concepts in large river ecology. BioScience. 1995;45:134-141
  173. 173. Johnson RR, Lowe CH. On the development of riparian ecology. In: Johnson RR, Ziebell CD, Patton DR, Ffolliott PF, Hamre RH, editors. Riparian Ecosystems and Their Management: Reconciling Conflicting Uses. Tucson, AZ: Department of Agriculture Forest Service; 1985. pp. 112-116
  174. 174. Karr JR. Biological integrity: A long-neglected aspect of water resource management. Ecological Applications. 1991;1:66-84
  175. 175. Karr JR. Defining and measuring river health. Freshwater Biology. 1999;41:221-234
  176. 176. Karr JR, Chu EW. Restoring Life in Running Waters: Better Biological Monitoring. Washington, DC: Island Press; 1999
  177. 177. Malanson GP. Riparian Landscapes. Cambridge: Cambridge University Press; 1993
  178. 178. Thorp JH, Thoms MC, DeLong MD. The riverine ecosystem synthesis: Biocomplexity in river networks across space and time. River Research & Applications. 2006;22:123-147
  179. 179. Connell JH, Slayter RO. Mechanisms of succession in natural communities and their role in community stability and organization. American Naturalist. 1977;111:1119-1144
  180. 180. Hough-Snee N, Laub BG, Merritt DM, Long AL, Nackley LL, Roper BB, et al. Multi-scale environmental filters and niche partitioning govern the distributions of riparian vegetation guilds. Ecosphere. 2015;6:1-22
  181. 181. Nilsson C. Frequency distributions of vascular plants in the geolittoral vegetation along two rivers in northern Sweden. Journal of Biogeography. 1983;10:351-369
  182. 182. Carothers SW, Aitchison SW, Johnson RR. Natural resources, white-water recreation and river management alternatives on the Colorado River, Grand Canyon National Park, Arizona. Proceedings of the First Conference on Scientific Research in the National Parks. 1979;1:253-260
  183. 183. Merritt DM, Scott ML, Poff NL, Auble GT, Lytle DA. Theory, methods and tools for determining environmental flows for riparian vegetation: Riparian vegetation-flow response guilds. Freshwater Biology. 2009;55:206-225
  184. 184. Sousa WP. The role of disturbance in natural communities. Annual Review of Ecology & Systematics. 1984;15:353-391
  185. 185. Pringle CM, Naiman RJ, Bretschko G, et al. Patch dynamics in lotic systems: The stream as a mosaic. Journal of the North American Benthological Society. 1988;7:503-524
  186. 186. Townsend CR. The patch dynamics concept of stream community ecology. Journal of the North American Benthological Society. 1989;8:36-50
  187. 187. Townsend CR. The intermediate disturbance hypothesis, refugia, and biodiversity in streams. Limnology and Oceanography. 2003;42:938-949
  188. 188. Crow R, Karlstrom KE, McIntosh W, Peters L, Dunbar N. History of Quaternary volcanism and lava dams in western Grand Canyon based on LIDAR analysis, (40)Ar/(39)Ar dating, and field studies: Implications for flow stratigraphy, timing of volcanic events, and lava dams. Geosphere. 2008;4:183-206
  189. 189. Dalrymple GB, Hamblin KW. K-Ar ages of Pleistocene lava dams in the Grand Canyon in Arizona. Proceedings of the National Academy of Sciences of the United States of America. 1998;95:9744-9749
  190. 190. Fenton CR, Poreda RJ, Nash BP, Webb RH, Cerling TE. Geochemical discrimination of five Pleistocene lava-dam outburst flood deposits, western Grand Canyon, Arizona. Journal of Geology. 2004;112:91-110
  191. 191. Costa JE, Schuster RL. The Formation and Failure of Natural Dams. Washington, DC: U.S. Geological Survey; 1987
  192. 192. Rogers JD, Pyles MR. Evidence of catastrophic erosional events in the Grand Canyon. In: Proceedings of the 2nd Conference on Scientific Research in National Parks. Washington, DC: National Park Service; 1980. pp. 392-454
  193. 193. Castleton JJ, Moore JR, Aaron J, Christ M, Ivy-Ochs S. Dynamics and legacy of 4.8 ka rock avalanche that dammed Zion Canyon, Utah, USA. GSA Today. 2016;26:4-9
  194. 194. Hereford R, Jacoby GC, McCord VAS. Geomorphic History of the Virgin River in the Zion National Park Area, southwest Utah. Washington, DC: U.S. Geological Survey; 1995
  195. 195. Benito G, O’Connor JE. Number and size of last-glacial Missoula floods in the Columbia River valley between the Pasco Basin, Washington, and Portland. Oregon. Geological Society of America Bulletin. 2003;115:624-638
  196. 196. Reeburgh WS, Nebert DL. The Birth and Death of Russell Lake. Fairbanks: University of Alaska; 1987
  197. 197. Webb RH, Betancourt JL, Johnson RR, Turner RM. Requiem for the Santa Cruz River: An Environmental History of an Arizona River. Tucson, AZ: University of Arizona Press; 2014
  198. 198. Smith JA, Auerbach DA, Flessa KW, Flecker AS, Dietl GP. Fossil clam shells reveal unintended carbon cycling consequences of Colorado River management. Royal Society Open Science. 2016;3:160170
  199. 199. Chaghtai SM, Khattak H-U-R. Ecology of a dry stream bed in Peshawar, Pakistan. Pakistan Journal of Botany. 1983;15:93-98
  200. 200. Fukami T, Morin PJ. Productivity-biodiversity relationships depend on the history of community assembly. Nature. 2003;424:423-426
  201. 201. Hooper DU, Chapin FS, Ewel JJ, Hector A, Inchausti P, Lavorel S. Effects of biodiversity on ecosystem functioning: A consensus of current knowledge. Ecological Monographs. 2005;75:3-35
  202. 202. Marquard E, Weigelt A, Roscher C, Gubsch M, Lipowsky A, Schmid B. Positive biodiversity-productivity relationship due to increased plant density. Journal of Ecology. 2009;97:696-704
  203. 203. Tilman D, Reibch PB. Diversity and productivity in a long-term grassland experiment. Science. 2001;294:843-845
  204. 204. Stevens LE. Mechanisms of Riparian Plant Succession. Flagstaff: Northern Arizona University; 1989
  205. 205. Reice SR. Experimental disturbance and the maintenance of species diversity in a stream community. 1985;67:90-97
  206. 206. Wilkinson DM. The disturbing history of intermediate disturbance. Oikos. 1999;84:145-147
  207. 207. MacArthur RH, Wilson EO. The Theory of Island Biogeography. Princeton: Princeton University Press; 1967
  208. 208. Lytle DA. Use of rainfall cues by Abedus herberti (Hemiptera: Belostomatidae): A mechanism for avoiding flash floods. Journal of Insect Behavior. 1999;12:1-12
  209. 209. Bartels P, Cucherousset J, Steger K, Eklov P, Tranvik LJ, Hillebrand H. Reciprocal subsidies between freshwater and terrestrial ecosystems structure consumer resource dynamics. Ecology. 2012;93:1173-1182
  210. 210. Larsen S, Muehlbauer JD, Marti E. Resource subsidies between stream and terrestrial ecosystems under global change. Global Change Biology. 2015;22:2489-2504
  211. 211. U.S. Department of Agriculture Forest Service. Watershed and Air Management, Section 2526.05, in Agriculture Forest Service: 2000. Washington, DC: U.S. Department of Agriculture Forest Service; 2000
  212. 212. Riis T, Kelly-Quinn M, Aguiar FC, Manolaki P, Bruno D, Bejarano MD, et al. Global overview of ecosystem services provided by riparian vegetation. Bioscience. 2020;70(6):501-514
  213. 213. Dufour S, Rodríguez-González PM, Laslier M. Tracing the scientific trajectory of riparian vegetation studies: Main topics, approaches and needs in a globally changing world. Science Total Environment. 2019;25(653):1168-1185. DOI: 10.1016/j.scitotenv.2018.10.383
  214. 214. Tockner K, Stanford JA. Riverine flood plains: Present state and future trends. Environmental Conservation. 2002;29:308-330
  215. 215. Rood SB, Pan J, Gill KM, Franks CG, Samuelson A, Shepherd GM. Declining summer flows of Rocky Mountain rivers: Changing seasonal hydrology and probable impacts on floodplain forests. Journal of Hydrology. 2008;349:397-410
  216. 216. Perry LG, Andersen DC, Reynolds LV, Mark NS, Shafroth PB. Vulnerability of riparian ecosystems to elevated CO2 and climate change in arid and semiarid western North America. Global Change Biology. 2012;18:821-842
  217. 217. Vandenberghe J. Climate forcing of fluvial system development: An evolution of ideas. Quaterinary Science Review. 2003;22(20):2053-2060. DOI: 10.1016/S0277-3791(03)00213-0
  218. 218. Liu HH. Impact of climate change on groundwater recharge in dry areas: An ecohydrology approach. Journal of Hydrology. 2011;407:175-183
  219. 219. Okkonen J, Kløve B. A conceptual and statistical approach for the analysis of climate impact on ground water table fluctuation patterns in cold conditions. Journal of Hydrology. 2010;388(1-2):1-12
  220. 220. Bogaart PW, Van Balen RT, Kasse C, Vandenberghe J. Process-based modeling of fluvial system response to rapid climate change II. Application to the River Maas (the Netherlands) during the Last Glacial–Interglacial Transition. Quaterly Science Review. 2003;22(20):2097-2110. DOI: 10.1016/S0277-3791(03)00144-6
  221. 221. Cecil BC, Dulong FT. Precipitation models for sediment supply in warm climates. In: May C, Cecil B, Edgar NT, editors. Climate Controls on Stratigraphy. Tulsa,OK: SEPM Society for Sedimentary Geology; 2003. pp. 21-27. DOI: 10.2110/pec.03.77.0021
  222. 222. Groeneveld DP, Griepentrog TE. Interdependence of groundwater, riparian vegetation, and streambank stability: A Case Study. In: Riparian ecosystems and their management: Reconciling conflicting uses. First North American riparian conference:1985 April 16-18. Tuscon; 1985. pp. 44-48
  223. 223. Gribovszki Z, Kalicz P, Szilágyi J, Kucsara M. Riparian zone evapotranspiration estimation from diurnal groundwater level fluctuations. Journal of Hydrology. 2008;349(1-2):6-17. DOI: 10.1016/j.jhydrol.2007.10.049
  224. 224. Hai-liang XU, Mao YE, Ji-mei LI. Changes in groundwater levels and the response of natural vegetation to- second paper.pdf. Journal of Environmental Science. 2007;19:1199-1207
  225. 225. Butturini A, Bernal S, Sabater S, Sabater F. The influence of riparian-hyporheic zone on the hydrological responses in an intermittent stream. Hydrological Earth System. 2002;6(3):515-525. DOI: 10.5194/hess-6-515-2002
  226. 226. Keller EA, Kondolf GM, Hagerty DJ. Groundwater and fluvial processes; Selected observations. In: Higgins CG, Coates DR, editors. Groundwater Geomorphology; The Role of Subsurface Water in Earth-Surface Processes and Landforms. Boulder, CO: Geological Society of America; 1990
  227. 227. Kløve B, Ala-Aho P, Bertrand G, Gurdak JJ, Kupfersberger H, Kværner J, et al. Climate change impacts on groundwater and dependent ecosystems. Journal of Hydrology. 2014;518:250-266
  228. 228. Hiscock K, Sparkes R, Hodgens A. Evaluation of future climate change impacts on european groundwater resources. In: Treidel H, Martin-Bordes JJ, Gurdak JJ, editors. Climate Change Effects on Groundwater Resources: A Global Synthesis of Findings and Recommendations. Abingdon-on-Thames: Taylor & Francis Publishing; 2012. pp. 351-366
  229. 229. Hu W, Wang YQ , Li HJ, Huang MB, Hou MT, Li Z, et al. Dominant role of climate in determining spatio-temporal distribution of potential groundwater recharge at a regional scale. Journal of Hydrology. 2019;578:1016
  230. 230. Changnon SA, Huff FA, Hsu C-F. Relations between Precipitation and Shallow Groundwater in Illinois. Journal of Climate. 1988;9:1239-1250
  231. 231. Fu G, Crosbie RS, Barron O, Charles SP, Dawes W, Shi X, et al. Attributing variations of temporal and spatial groundwater recharge: A statistical analysis of climatic and non-climatic factors. Journal of Hydrology. 2019;568:816-834
  232. 232. Fritz P, Drimmie RJ, Frape SK, O’Shea K. The isotopic composition of precipitation and groundwater in Canada. In: Isotope Techniques in Water Resources Development. Proc. IAEA Symposium. Vienna; 1987. pp. 539-550
  233. 233. Lee KS, Wenner DB, Lee I. Using H- and O-isotopic data for estimating the relative contributions of rainy and dry season precipitation to groundwater: Example from Cheju Island, Korea. Journal of Hydrology. 1999;222(1-4):65-74
  234. 234. Yin L, Hou G, Su XS, Wang D, Dong J, Hao Y, et al. Isotopes (δD and δ18O) in precipitation, groundwater and surface water in the Ordos Plateau, China: Implications with respect to groundwater recharge and circulation. Hydrogeological Journal. 2011;19(2):429-443. DOI: 10.1007/s10040-010-0671-4
  235. 235. Jasechko S, Birks SJ, Gleeson T, Wada Y, Fawcett PJ, Sharp ZD, et al. The pronounced seasonality of global groundwater recharge. Water Resources Research. 2014;50(11):8845-8867. DOI: 10.1002/2014WR015809
  236. 236. Eastoe CJ, Wright WE. Hydrology of mountain blocks in Arizona and New Mexico as revealed by isotopes in groundwater and precipitation. Geoscience. 2019;9(11):461; DOI: 10.3390/geosciences9110461
  237. 237. Bloomfield JP, Marchant BP. Analysis of groundwater drought building on the standardized precipitation index approach. Hydrological Earth System Science. 2013;17(12):4769-4787
  238. 238. Kumar R, Musuuza JL, Van Loon AF, Teuling AJ, Barthel R, Ten Broek J, et al. Multiscale evaluation of the Standardized Precipitation Index as a groundwater drought indicator. Hydrological Earth System Science. 2016;20(3):1117-1131. DOI: 10.5194/hess-20-1117-2016
  239. 239. Yeh H-F, Hsu H-L. Using the Markov Chain to analyze precipitation and groundwater drought characteristics and linkage with atmospheric circulation. Sustainability. 2019;11(6):1817
  240. 240. Thomas BF, Behrangi A, Famiglietti JS. Precipitation intensity effects on groundwater recharge in the southwestern United States. Water (Switzerland). 2016;8(3):12-17. DOI: 10.3390/w8030090
  241. 241. Asoka A, Wada Y, Fishman R, Mishra V. Strong linkage between precipitation intensity and monsoon season groundwater recharge in India. Geophysical Research Letters. 2018;45(11):5536-5544. DOI: 10.1029/2018GL078466
  242. 242. Liu J, Ma X, Duan Z, Jiang J, Reichstein M, Jung M. Impact of temporal precipitation variability on ecosystem productivity. Wiley Interdisciplinary Review Water. 2020;7(6):1-22. DOI: 10.1002/wat2.1481
  243. 243. Lotsch A, Friedl MA, Anderson BT, Tucker CJ. Coupled vegetation-precipitation variability observed from satellite and climate records. Geophysical Research Letters. 2003;30(14):8-11. DOI: 10.1029/2003GL017506
  244. 244. Rasmusson EM, Arkin PA. A global view of large-scale precipitation variability. Journal of Climatology. 1993;6(8):1495-1522
  245. 245. Panagoulia D. Assessment of daily catchment precipitation in mountainous regions for climate change interpretation. Hydrological Science. 1995;40(3):331-350
  246. 246. Panagoulia D, Bárdossy A, Lourmas G. Diagnostic statistics of daily rainfall variability in an evolving climate. Advance Geoscience. 2006;7:349-354
  247. 247. Panagoulia D, Bardossy A, Lourmas G. Multivariate stochastic downscaling models for generating precipitation and temperature scenarios of climate change based on atmospheric circulation. Global Nest Journal. 2006;10:263-272
  248. 248. Gherardi LA, Sala OE. Enhanced precipitation variability decreases grass- and increases shrub-productivity. Proceedings of the National Academy Science USA. 2015;112(41):12735-12740. DOI: 10.1073/pnas.1506433112
  249. 249. Dong L, Guo Y, Tang W, Xu W, Fan Z. Statistical evaluation of the influences of precipitation and river level fluctuations on groundwater in Yoshino River Basin, Japan. Water. 2022;14(4):12735
  250. 250. Pendergrass AG, Knutti R, Lehner F, Deser C, Sanderson BM. Precipitation variability increases in a warmer climate. Scientific Reports. 2017;7(1):1-9. DOI: 10.1038/s41598-017-17,966-y
  251. 251. Zhang W, Furtado K, Wu P, Zhou T, Chadwick R, Marzin C, et al. Increasing precipitation variability on daily-to-multiyear time scales in a warmer world. Scientific Advances. 2021;7(31):1-12. DOI: 10.1126/sciadv.abf8021
  252. 252. Panagoulia D. Hydrological modeling of a medium-size mountainous catchment from incomplete meteorological data. Journal of Hydrology. 1992;135(1-4):279-310
  253. 253. Panagoulia D, Vlahogianni EI. Nonlinear dynamics and recurrence analysis of extreme precipitation for observed and general circulation model generated climates. Hydrological Process. 2014;28(4):2281-2292. DOI: 10.1002/hyp.9802
  254. 254. Panagoulia D, Economou P, Caroni C. Stationary and non-stationary GEV modeling of extreme precipitation over a mountainous area under climate change. Environmetrics. 2014;25(1):29-43
  255. 255. Zheng W, Wang S, Sprenger M, Liu B, Cao J. Response of soil water movement and groundwater recharge to extreme precipitation in a headwater catchment in the North China Plain. Journal of Hydrology. 2019;576:466-477
  256. 256. Boutt DF, Mabee SB, Yu Q. Multiyear increase in the stable isotopic composition of stream water from groundwater recharge due to extreme precipitation. Geophysical Research Letters. 2019;46(10):5323-5330. DOI: 10.1029/2019GL082828
  257. 257. Liu Y, Jiang X, Zhang G, Xu YJ, Wang X, Qi P. Assessment of shallow groundwater recharge from extreme rainfalls in the Sanjiang Plain, northeast china. Water. 2016;8(10):5323-5330
  258. 258. Qi P, Zhang G, Xu YJ, Wang L, Ding C, Cheng C. Assessing the influence of precipitation on shallow groundwater table response using a combination of singular value decomposition and cross-wavelet approaches. Water. 2018;10(5):23
  259. 259. Shao J, Si B, Jin J. Rooting depth and extreme precipitation regulate groundwater recharge in the thick unsaturated zone: A case study. Water. 2019;11(6):3390
  260. 260. Bailey JK, Whitham TG. Interactions between cottonwood and beavers positively affect sawfly abundance. Ecological Entomology. 2006;31:294-297
  261. 261. Sacchi CF, Price PW. Pollination of the arroyo willow, Salix lasiolepis: Role of insects and wind. American Journal of Botany. 1988;75:1387-1393
  262. 262. Simieon G, Stevens LE. Tamarix (Tamaricaceae), Opsius stactogalus (Cicidellidae), and litter fungi interactions limit riparian plant establishment. Advances in Entomology. 2015;3:65-81
  263. 263. Carlisle JD, Skagen SK, Kus BE, Van Riper CIII, Paxton KL, Kelly JF. Landbird migration in the American West: Recent progress and future research directions. The Condor. 2009;111:211-225
  264. 264. Skagen SK, Kelly JF, Van Riper CIII, Hutto RL, Finch DM, Krueper DJ, et al. Geography of spring landbird migration through riparian habitats in southwestern North America. The Condor. 2005;107:212-227
  265. 265. Kelly JF, Hutto RL. An East–West comparison of migration in North American wood warblers. The Condor. 2005;107:197-211
  266. 266. Stevens LE, Buck KA, Brown BT, Kline N. Dam and geomorphic influences on Colorado River waterbird distribution, Grand Canyon, Arizona. Regulated Rivers: Research & Management. 1997;13:151-169
  267. 267. Geerling GW, Ragas AMJ, Leuven RSEW, van den Berg JH, Breedveld M, Liefhebber D, et al. Succession and rejuvenation in floodplains along the River Allier (France). Hydrobiologia. 2006;565:71-86
  268. 268. Piotrowski JS, Lekberg Y, Harner MJ, Ramsey PW, Rillig MC. Dynamics of mycorrhizae during development of riparian forests along an unregulated river. Ecography. 2008;31:245-253
  269. 269. Woodbury AM, editor. Ecological Studies of Flora and Fauna in GlenCanyon. Salt Lake City: University of Utah; 1959
  270. 270. Holdridge LR. Determination of world plant formations from simple climatic data. Science. 1947;105:367-368
  271. 271. Gomi T, Sidle RC, Richardson JS. Understanding process and downstream linkages of headwater systems. BioScience. 2002;52:905-916
  272. 272. Cantonati M, Stevens LE, Segadelli S, Springer AE, Goldscheider N, Celico F, et al. Ecohydrogeology: The interdisciplinary convergence needed to improve the study and stewardship of springs and other groundwater-dependent habitats, biota, and ecosystems. Ecological Indicators. 2020;110. DOI: 10.1016/j.ecolind.2019.105803
  273. 273. Lubinski KS. A Conceptual Model of the Upper Mississippi River System. Onalaska: U.S. Fish and Wildlife Service; 1993
  274. 274. Stromberg J, Briggs M, Gourley C, Scott M, Shafroth P, Stevens L. Human alterations of riparian ecosystems. In: Baker MB Jr, Ffolliott PF, DeBano L, Neary DG, editors. Riparian Areas of the Southwestern United States: Hydrology, Ecology, and Management. Boca Raton: Lewis Publishers; 2004. pp. 99-126
  275. 275. Steiger J, Tabacchi E, Dufour S, Corenblit D, Peiry J-L. Hydrogeomorphic processes affecting riparian habitat within alluvial channel-floodplain river systems: A review for the temperate zone. River Research and Applications. 2005;21:719-737
  276. 276. Harvey J, Gooseff M. River corridor science: Hydrologic exchange and ecological consequences from bedforms to basins. Water Resources Research. 2015;51:6893-6922
  277. 277. Ward JV. Ecology of alpine streams. Freshwater Biology. 1994;32:277-294
  278. 278. Scrimgeour GJ, Prowse TD, Culp JM, Chambers PA. Ecological effects of river ice breakup: A review and perspective. Freshwater Biology. 1994;32:261-275

Written By

Lawrence E. Stevens, Raymond R. Johnson and Christopher Estes

Reviewed: 19 August 2022 Published: 13 October 2022