Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Received: 21 December 2017 Revised: 18 July 2018 Accepted: 19 July 2018

DOI: 10.1002/eco.2018

RESEARCH ARTICLE

Linking geomorphic change due to floods to spatial hydraulic


habitat dynamics

Aaron Tamminga Brett Eaton

University of British Columbia, ,Vancouver,


British Columbia, Canada Abstract
Large flood events have the capacity to induce geomorphic restructuring that can impact riverine
Correspondence
Aaron Tamminga, University of British ecosystems. However, the detailed morphodynamics associated with infrequent, high-magnitude
Columbia, Vancouver, British Columbia, floods are variable and difficult to capture, and more research is needed into potential rela-
Canada.
tionships between geomorphic change, flow organization, and aquatic habitat dynamics. In this
Email: tamminga@geog.ubc.ca
study, we focus on the reach-scale response of a gravel bed river to a large flood, employing
a combined remote sensing, field measurement, and numerical modelling approach to measure
Handling Editor: Smettem Keith
and interpret conditions bracketing the flood. Documented geomorphic turnover was extensive,
reworking low-flow channel patterns and causing widespread bank erosion and sediment depo-
sition. This resulted in a shift to wide, shallow flow conditions in the post-flood morphology and
a loss of hydraulic diversity, particularly in ecologically important pool and riffle units identified
using a fuzzy statistical classification method. These impacts are most evident at low flows;
higher discharges display relatively similar hydraulic conditions despite geomorphic change.
Smaller-scale adjustments in the year following the flood appear to be driving the reintroduction
of hydraulic diversity, which is interpreted as beneficial for in-stream brown trout. Results from
this study highlight the utility of applying flexible and objective remote sensing and modelling
methods to measure fluvial change and provide a real-world example that can inform broader
theoretical understanding of large flood ecohydrology.

KEYWORDS
aquatic habitat, fluvial remote sensing, flood impacts, geomorphology

1 INTRODUCTION understanding of the physical processes driving fluvial adjustment


(Spencer & Lane, 2017; Vaughan et al. 2009). In particular, informa-
The dynamic interplay between physical and hydrological processes tion on geomorphic process rates and recovery trajectories is needed
that shapes fluvial environments creates complex ecosystems that to inform flood risk management and maintain ecosystem services in
are uniquely sensitive to external pressures. In particular, the effects an age of climate extremes (Gaston and Blackburn 1999; Naylor et al.,
of flood events driven by broad scale climatic trends are integrated 2017; Phillips & Van Dyke, 2016).
in river channels, instigating geomorphic feedbacks that can variably Because geomorphic processes result in nonrandom assemblages of
mediate or increase impacts to stream ecosystems (Fryirs, 2017; Gart- flow properties and sediment sizes throughout a river reach (Wheaton
ner, Dade, Renshaw, Magilligan, & Buraas, 2015; Harrison, Pike, & et al., 2015; Wyrick & Pasternack, 2014), ecologically significant
Boughton, 2017; Magilligan, Buraas, & Renshaw, 2015; Thompson & patches also arise at this scale, with differing utility in terms of habitat
Croke, 2013). Due to their inherent rarity and unpredictability, the for various organisms. Hydraulic attributes determine habitat avail-
effects of large infrequent disturbances such as floods are difficult to ability and quality for aquatic organisms ranging from macrophytes
study reliably and remain poorly understood (Church, 1980; Turner & and benthic invertebrates to fish communities (Crowder & Diplas,
Dale, 1998). However, increases in storm frequency and intensity are 2006; Hughes & Dill, 1990; Madsen, Enevoldsen, & Jørgensen, 1993;
predicted as a consequence of climate change in many regions (Chris- Silva, Ligeiro, Hughes, & Callisto, 2014), with important implications
tensen, Kanikicharla, Marshall, & Turner, 2013), necessitating a deeper for linked aquatic and terrestrial ecosystems (Gende, Edwards, Will-

Ecohydrology. 2018;e2018. wileyonlinelibrary.com/journal/eco © 2018 John Wiley & Sons, Ltd. 1 of 17


https://doi.org/10.1002/eco.2018
2 of 17 TAMMINGA AND EATON

son, & Wipfli, 2002; Nakano & Murakami, 2001). A rich history of and interpret the findings in terms of reachscape hydraulic composi-
research linking flow conditions with species preference and use exists tion and configuration with the objective of answering the question:
(Chapman, 1966; Clifford, Acreman, & Booker, 2008; Leclerc, 2005; What happens to physical habitat during a large flood event and how
Statzner, Gore, & Resh, 1988), based largely on studies of lotic fish quickly does it recover?
habitat that have shown that specific physical conditions are required
for the spawning, rearing, and feeding of different species and life
stages (Armstrong, Kemp, Kennedy, Ladle, & Milner, 2003; Heggenes
& Saltveit, 1990; Morantz, Sweeny, Shirvell, & Longard, 1987). As a
2 METHODS
result, many examples of habitat analyses for target species are based
on mesoscale (1–10 channel widths) discretization of stream environ-
2.1 Study site
ments (Dunbar, Alfredsen, & Harby, 2011; Newson & Newson, 2000; Elbow River is a gravel bed river draining the eastern slopes of
Parasiewicz, 2001; Silva et al., 2014), with similar approaches form- the Canadian Rocky Mountains. Part of the South Saskatchewan
ing a core physical-biological link in management activities (Crowder River basin, it originates in subalpine areas of the mountains, flow-
& Diplas, 2002; Hauer, Mandlburger, Schober, & Habersack, 2012). ing through foothills into prairie and farmland before joining with
However, methods that rely on empirical abundance-environment cor- Bow River in Calgary, Alberta. Although fish assemblages vary geo-
relations have been criticized for ignoring biological interactions such graphically throughout the watershed, the river supports native bull
as density-dependent competition and for unfounded interpretations trout (Salvelinus confluentus), westslope cutthroat trout (Oncorhynchus
of cause-and-effect when other causal mechanisms such as spatial clarkii lewisi), and mountain whitefish (Prosopium williamsoni) pop-
variability may be at play (Angermeier & Schlosser, 1989; Lancaster & ulations, as well as non-native brook trout (Salvelinus fontinalis),
Downes, 2010; Van Horne, 1983). rainbow trout (Oncorhynchus mykiss), and brown trout (Salmo trutta;
Ensuring robust interpretation of the relationships between flood Nelson & Paetz, 1992). Bull trout and cutthroat trout populations
geomorphology and aquatic ecosystem structure and function requires have declined over the past century due to interspecies competi-
detailed field measurement supported by numerical and physical tion, exploitation, and habitat degradation (Post & Johnston, 2002).
modelling to build and test conceptual models of ecosystem evo- All of the local salmonids are valued as sport fish, and under-
lution. Recent theoretical and methodological movements have led standing their contrasting and/or competing ecological requirements
to an increased view of rivers as spatially continuous mosaics and geographical distributions is important to sustainable fisheries
with rich three-dimensional heterogeneity resulting from linked management.
physical–biological processes acting at a range of scales (Carbon- Our study reach was a 1-km section of Elbow River near Bragg
neau, Fonstad, Marcus, & Dugdale, 2012; Fausch, Torgersen, Baxter, Creek, Alberta. At this point, the river has exited the mountains
& Li, 2002; Schlosser, 1991; Ward, 1998; Wiens, 2002). To reliably and exhibits an anabranching/braided morphology flowing through a
characterize such diversity in dynamic fluvial systems, objective, accu- wide (200–300 m) sediment-rich gravel/cobble active channel. The
rate, and efficient methods are needed (Woodget, Visser, Maddock, reach was chosen initially as a test site for fluvial remote sensing
& Carbonneau, 2016). Fortunately, advances in remote sensing and methods using UAVs and was surveyed in detail in September 2012
associated computational methods now allow for accurate determi- (Tamminga, Hugenholtz, Eaton, & Lapointe, 2015b). The following
nation of river morphology to quantify morphological adjustments summer (June 19–22, 2013) the region experienced extreme flooding
over multiple scales, to visualize vertical in addition to planform caused by a sustained high-intensity rainfall event and rapid melting of
channel adjustments, and as input to numerical models of chan- a lingering snowpack in the mountains; water levels were the highest
nel processes (Lane, 2000). In particular, the combined approach of in 60 years and estimated damages exceeded CAD$6 billion (Milrad,
digital photogrammetry techniques and small unmanned aerial vehi- Gyakum, & Atallah, 2015; Pomeroy, 2015). At the Elbow study reach,
cle (UAV)-based image capture has emerged as a valuable way to peak discharge was estimated at greater than 800 m3 s−1 based
measure river channel topography in three dimensions (Fonstad, Diet- on a locally calibrated stage–discharge rating curve. Although the
rich, Courville, Jensen, & Carbonneau, 2013; Javernick, Brasington, & short record of stream gauge data and difficulty in applying statistical
Caruso, 2014; Javernick, Hicks, Measures, Caruso, & Brasington, 2015; analysis to events generated by different mechanisms (snowmelt
Vericat, Brasington, Wheaton, & Cowie, 2008; Woodget, Carbonneau, vs. rainfall vs. rain-on-snow) make flood frequency analysis for the
Visser, & Maddock, 2014). event highly uncertain, annual peak discharges for the Bragg Creek
Combining digital representations of channel form with appropriate Water Survey of Canada gauging station (05BJ004) for 1935–2014
methods of interpreting ecological consequences presents a power- are shown for context with a Log Pearson Type III distribution in
ful tool for fluvial ecohydrologic analysis, yet multitemporal studies Figure 1. In this analysis, the previous largest flow was 231 m3 s−1 in
applying such methods to capture the dynamics associated with flood 2005, which corresponds to a return period of approximately 40 years.
impacts are generally lacking. In this study, we focus on the relation- The 2014 freshet had a return period of about 5 years. The 2013
ship between reach-scale geomorphic change and spatio-temporal flood is an outlier from the data; the return period as shown is just
hydraulic variation by taking advantage of a set of UAV surveys brack- under 100 years, but extending the fitted distribution would predict
eting an extreme flood event. We address changes to channel form a return period of over 500 years for this magnitude of event. A
and apply a fuzzy statistical habitat clustering method (Legleiter & longer time period including other large events would help reduce this
Goodchild, 2005) to modelled flow patterns at a range of discharges uncertainty.
TAMMINGA AND EATON 3 of 17

tured pre-flood conditions, the September 2013 survey shows the


reach structure following the June 2013 flood event (no geomor-
phically effective floods occurred between the flood and the data
collection), and the September 2014 survey shows how the reach
adjusted following a more common (yet still significant) spring freshet.

2.3 Hydrodynamic modelling


The photogrammetry-generated DEMs were then used as topographic
input to Nays2DH, a depth-averaged hydrodynamic model imple-
mented within the International River Interface Cooperative software
framework (Nelson et al. 2016). Nays2DH solves the shallow water
equations using a general curvilinear grid for numerical discretization;
grid spacing for our models was set to 1 m by 1 m based on estimates
of the Peclet number (Papanicolaou, Elhakeem, & Wardman, 2010)
to minimize numerical diffusion while maintaining computational effi-
ciency and the ability to resolve the small-scale geomorphic forms
such unit bars and pool/riffle-associated breaks in slope that con-
trol flow patterns over a sufficiently long study reach. Other model
inputs included downstream water surface elevation, discharge, and
roughness in the form of Manning's n. For each year, we initially
calibrated steady flow models at the discharge captured during the
UAV surveys (6 m3 s−1 in 2013 and 10 m3 s−1 in 2013 and 2014).
This allowed for comparison of modelled and measured water surface
elevations (extracted directly from the DEMs along the water's edge
in the orthomosaic) and fine tuning of the flow resistance param-
eter. Flow resistance was treated as spatially uniform, and starting
values were estimated from D84 values using the Manning–Strickler
approach, and lateral eddy viscosity was estimated from reach aver-
age depth and reach average velocity (Nelson & McDonald, 1996). We
also compared field-measured depths with modelled depths to fur-
FIGURE 1 (a) Flood frequency analysis for Elbow River at Bragg ther check these low-flow model results (Table 1). Relative to the size
Creek, 1935–2014. Grey dashed boundaries indicate 90% confidence of sediment throughout the reach (D84 = 0.068 m), which introduces
intervals, the two flood years covered in this study are annotated by
inherent uncertainty in point-measured elevations and fluctuations in
red points. (b) 2012–2015 hydrograph showing flow conditions
relative to timing of unmanned aerial vehicle surveys (denoted by water surface, model performance following calibration was good and
vertical blue arrows) adequately captured the scale of hydraulic variation in which we were
interested.
We then ran steady flow simulations based on the calibrated model
2.2 UAV surveys parameters for three specified discharges for each year: 10, 30, and
In this study, we make use of three datasets collected through 50 m3 s−1 . These values range from the base flow typical of the late
UAV-based remote sensing. For this scale of analysis, UAVs pro- summer or fall (10 m3 s−1 ) to an approximately 2-year flood (50 m3 s−1 )
vide an ideal platform to provide high-resolution optical imagery and therefore reflect the hydraulic conditions aquatic organisms in
covering the study reach. Each year, 200–300 individual georefer- Elbow River would be subject to throughout a normal year. Although
enced images taken from approximately 100-m flying height with extending low-flow resistance values to higher discharges can be
60–80% overlap were collected. These photos were processed with an uncertain process (Ferguson, 2010), the high-resolution model
photogrammetry software (Pix4D) to generate orthorectified image topography meant that form roughness was largely captured in our
mosaics (4–5 cm/pixel resolution) and digital elevation models (DEMs). calibration procedure, enabling more reliable extension to higher
After areas with submerged topography were corrected using an opti- flows (Williams et al., 2013). Similar scale modelling studies have
cal/empirical depth estimation procedure, DEM vertical root mean also reported few meaningful differences in indicators of boundary
square errors were 5–10 cm. Further details on DEM correction and roughness at low and high flows (Hauer & Habersack, 2009); Krapesch,
assessment can be found in Tamminga et al. (2015b). Timings of the Hauer, & Habersack, 2011; Wyrick & Pasternack, 2014). These nine
surveys relative to the hydrograph for the period 2012–2015 are simulations were then used to examine flow structure in terms of
shown in Figure 1; each survey was conducted in late September at reach average parameters and depth–velocity frequency distribu-
low-flow conditions (6–10 m3 s−1 ). The September 2012 survey cap- tions in a stochastic habitat hydraulics context. As an initial measure
4 of 17 TAMMINGA AND EATON

TABLE 1 Calibration summary of modelled versus measured depth and water surface
elevations for Nays2DH hydrodynamic simulations
Year Discharge (m3 s−1 ) WSE R2 WSE RMSE (m) Depth R2 Depth RMSE (m)

2012 6.1 0.99 0.12 0.72 0.11


2013 10.3 0.97 0.18 0.80 0.09
2014 10.1 0.98 0.17 0.75 0.08

Note. RMSE: root mean square error; WSE: water surface elevation.

of reach-scale heterogeneity, the hydromorphological index of ing a baseline to compare with different geomorphic configurations
diversity (HMID, Gostner, Alp, Schleiss, & Robinson 2013) was and discharges.
calculated as For the clustering, we chose three variables that strongly reflect
( )2 ( )2 hydraulic variability: flow depth, depth-averaged velocity, and Froude
𝜎 𝜎
HMID = 1 + U ∗ 1+ D , number. Including Froude number provides an explicit weighting to
𝜇U 𝜇D
the relationship between depth and velocity; stream habitat structure
where 𝜎 U and 𝜇U are velocity standard deviation and mean velocity, and use are often associated with distinct combinations of depth and
respectively, and 𝜎 D and 𝜇D are depth standard deviation and mean velocity as opposed to either variable independently (Kemp, Harper,
depth, respectively. By this measure, simple morphologies where & Crosa, 1999; Statzner et al., 1988). We normalized each of these
flow depth and velocity distributions display little variability (𝜎 → 0) factors by the reach average value for consistency and then applied
have small HMID values, whereas morphological complexity causes the clustering algorithm, varying the number of clusters and degree
heterogeneous hydraulic variable distributions (greater HMID values) of fuzziness (m) until we settled on four distinct clusters and m = 2
in more diverse reaches. as a way of balancing detail and interpretability while minimizing
the within-group sum of squares to ensure separate and internally
compact clusters. We then extended the analysis to the 2012 higher
flows (30 and 50 m3 s−1 ) and all flows for 2013 and 2014 by calculating
2.4 Fuzzy cluster analysis
partial memberships for each flow model point to each of the four
To further interpret flow patterns in terms of ecologically relevant cluster centres determined through the 2012 10-m3 s−1 classification.
units, we used a fuzzy clustering method to classify modelled flow To facilitate interpretation of spatial patterns, we then also assigned
distributions into cohesive groups. Classification is fundamental to the each point to the cluster to which it belonged most strongly, creating
understanding of the geospatial organization of fluvial environments, rigid hydraulic units with discrete boundaries while still being able
but the subjectivity and reliance on expert knowledge associated with to access the original partial membership values that describe the
many conventional habitat classification methods compromises their uncertainty in the classification. Cluster patterns were then mapped as
transferability and reliability (Poole, Frissell, & Ralph, 1997). Similarly, hydraulic patches displaying particular flow conditions and landscape
rigid Boolean classification schemes tend to obscure uncertainty, tran- metrics describing geospatial organization were calculated using the
sitions between classes, and within-class variability (Fisher, 1998). For ‘‘SDMTools’’ R package. The chosen metrics calculated for each cluster
this reason, we applied the fuzzy c-means algorithm (Bezdek, Ehrlich, type were proportion of wetted area, patch density (a measure of the
& Full, 1984), as proposed by Legleiter and Goodchild (2005). This number of discrete patches present for each cluster), edge density
approach allows individual data points to exhibit partial membership (a measure of the length of the total length of patch edge for each
in multiple classes, retaining a measure of uncertainty in the unsuper- cluster), and mean patch area.
vised creation of fuzzy sets. Data points (in this case, 1-m2 pixels of
Nays2DH model outputs) are grouped into a number of sets based on
Euclidian distance from the cluster centre and have membership val-
ues ranging from 0 (completely unlike the set) to 1 (perfect example
2.5 Brown trout habitat
of the set) in each set. If desired, rigid classes can then be formal- The hydraulic units developed with the clustering approach were cre-
ized by choosing a threshold membership value (Cheng, Molenaar, ated with the objective of dividing reach-scale flow patterns into
& Lin, 2001). ecologically relevant patches yet were created without a specific target
We applied the clustering algorithm (implemented in the R package species in mind. This was done intentionally to focus on the relation-
‘‘e1071’’) to the modelled 10 m3 s−1 discharge in the pre-flood (2012) ship between morphodynamics and spatial flow structure. However, it
morphology. We chose to use low-flow data to define the clusters is also useful to interpret the cluster results and general fluvial adjust-
because it reflects the common conditions that fish and other aquatic ments through a fish habitat lens. For this reason, we also used the
organisms experience throughout most of the year while performing modelled depths and velocities to calculate more conventional mea-
their daily actions such as feeding, resting, and rearing. Periods of low sures of habitat suitability for brown trout. We chose brown trout
flow are also when geomorphic structure exerts the strongest con- because it is a common salmonid in the region with well-documented
trol on hydraulic patterns. Classifying based on low-flow data in the habitat preferences. Using preference curves generated for the nearby
pre-flood morphology therefore stratifies the river reach into hydraulic Kananaskis River (Courtney, Walder, & Vadas, 1998), we calculated
units that have strong associations with underlying landforms, provid- habitat suitability indices reflecting preferences for depth (DHS ) and
TAMMINGA AND EATON 5 of 17

velocity (VHS ) for juvenile and adult brown trout. The values for each but here, fluvial reworking through the 2014 spring freshet appears to
factor range from 0 (completely unsuitable) to 1 (ideal habitat). Over- have entrenched flows even more into a single thread conveying the
all adult and juvenile composite suitability values (AdultHS , JuvHS ) were bulk of the discharge with less access to secondary channels. These
then calculated as HS = (DHS )1/2 (VHS )1/2 (Leclerc, Boudreault, Bechara, patterns are accentuated as modelled discharges increase to 30 and
& Corfa, 1995; Pasternack, Wang, & Merz, 2004). 50 m3 s−1 ; 2012 flows display a wider range of hydraulic conditions
with deep scour and bend apex pools having locally high specific dis-
charges, but a significant portion of the flow is spread through gradually
3 RESULTS sloping bar surfaces, secondary anabranches, and areas of slackwater.
In 2013, these higher discharges also activate several anabranches
The general geomorphic changes for the study period are shown in and fill shallow areas, whereas the 2014 high discharges more strongly
Figure 2. Overall, pre-flood reach morphology was characterized by follow a main thalweg with high specific discharge values relative to
a sinuous primary low-flow thread with several major anabranches the previous 2 years.
associated with bar dissection and chute cut-off. The 2013 flood These changes in flow pattern can be further examined through
event resulted in substantial bank erosion (tens of metres of bank reach average hydraulic parameters and joint depth–velocity fre-
retreat) on both sides of the 2012 channel as width adjusted to quency distributions. Overall, flows in 2012 are deeper, and reach
the large discharges, along with a complete reworking of sediment average widths are less than in 2013 for a given discharge (Table 2).
and morphology within the active channel. The bulk of the low-flow In 2014, widths are less than in 2012 or in 2013, corresponding to
discharge in 2013 appears to be mainly confined into one or two increased average velocities. This is reflected in a large change of
wide, shallow anabranches with a long meander wavelength reflect- width/depth ratio, which is much higher in the 2013 post-flood con-
ing the path cut by the waning flood flows. The 2014 morphology figuration than in 2012. By 2014, width/depth ratio is once again
reflects more subtle changes that occurred during the 2014 freshet reduced to values similar to before the flood.
including bank erosion at bend apices and associated bar deposi- Bivariate depth–velocity frequency distributions confirm these gen-
tion, along with infilling of smaller high-elevation secondary channels eral changes. The associations between depth and velocity (Figure 4)
produced by the 2013 flood but likely not accessed during the for the modelled flow data vary with discharge and by geomor-
2014 freshet. phic configuration (year). In 2012, the positive association between
depths and velocities is weak at 10 m3 s−1 , reflecting a diversity
of hydraulic conditions and leading to a high HMID value of 8.2.
3.1 Flow pattern changes In 2013 and 2014, the HMID values are lower as flows are more
Results from the Nays2DH model runs give insights into the relation- channelized and the diversity offered by low-velocity, high-depth
ships between morphodynamics and flow patterns. Maps of specific areas and high-velocity, low-depth areas is largely lost. All 3 years
discharge (Figure 3) reflect the rearrangement of low-flow topogra- show evidence of a drowning out of mesoscale topographic influ-
phy; 10 m s3 −1
flows in 2012 follow two to three main anabranches ence and roughness elements as discharge increases (ie, more similar
splitting around large medial bars with associated flow convergence flow distributions and HMID values despite different morphologies),
and scour pools where threads meet. In 2013, flow is more restricted but the 2012 morphology remains distinct in its wide range of
to a wide primary channel produced by the flood but occasionally hydraulic conditions even at 50 m3 s−1 ; 2013 and 2014, on the
diverges as flows access other high-elevation secondary channels. The other hand, show a comparatively low diversity of conditions by
2014 10-m s 3 −1
flow is confined the to same general path as in 2013, 30 and 50 m3 s−1 .

FIGURE 2 Unmanned aerial vehicle orthomosaics for each survey year. Flow direction is from bottom to top
6 of 17 TAMMINGA AND EATON

FIGURE 3 Steady flow hydrodynamic model results shown as maps of specific discharge (q, m2 s−1 ) for each year and discharge combination

3.2 Cluster results Cluster 1 (low energy) represents low-slope, low-velocity environ-
ments with shallow depths, typically found along stream margins,
To further interpret the general changes in flow patterns in terms in shallow secondary channels, or in backwater areas of scarcely
of reachscape composition and spatial structure, we classified the perceptible flow. Cluster 2 (pool) is deeper flow with a range of veloc-
2012 10-m3 s−1 flow data. Results of the fuzzy c-means clustering ities below about 1 m s−1 and low water surface slope. Cluster 3
are shown in Figure 5. The four clusters stratify the depth–velocity (glide) is flow with moderate depths, velocities, and water surface
flow field into groups of characteristic flow conditions (Table 3), and slopes, generally found in areas of diverging flow or along lateral
a detailed example view of the relationship between geomorphic set- gradients flanking threads of higher intensity flow. Cluster 4 (riffle)
ting, flow patterns, and clusters is shown in Figure 6. These clusters is high-velocity, high-energy flow that often exhibits broken water
reflect specific hydraulic conditions associated with controls exerted surface texture. In general, riffle and pool clusters are thalweg flow
by mesoscale bedforms and therefore largely correspond with the conveying similar specific discharges, and the four clusters can be
common descriptive geomorphic unit terms pool, riffle, and glide. differentiated by Froude number, increasing from an average of 0.21
TAMMINGA AND EATON 7 of 17

TABLE 2 Summary of reach average hydraulic parameters for each simulation


Year Discharge (m3 s−1 ) Depth (m) Width (m) Velocity (m s−1 ) Froude (—) 𝜏 (Pa) Width/depth (—)

2012 10 0.42 31.0 0.67 0.46 16.8 73.1


30 0.51 45.1 1.03 0.49 32.6 87.5
50 0.56 64.0 1.06 0.49 33.5 115.1
2013 10 0.30 38.4 0.69 0.42 16.6 128.4
30 0.45 54.6 0.94 0.47 26.3 122.6
50 0.52 68.8 1.06 0.49 31.5 132.4
2014 10 0.37 28.3 0.82 0.44 20.6 77.3
30 0.51 42.5 1.07 0.49 31.9 82.5
50 0.58 55.1 1.19 0.51 37.2 94.2

FIGURE 4 Joint depth–velocity frequency distributions. Hydromorphological index of diversity values for each year-discharge combination are
annotated on each panel

in low-energy units, 0.33 in pools, 0.51 in glides, to 0.67 in riffles hydraulic units. Reachscape cluster composition is summed up as per-
(Table 3). The paired blue/green colour scheme also reflects the gen- centage of wetted area for each cluster in Figure 8. In general, flow
eral conditions and locations of the clusters: blue for the low Froude in 2012 morphology appears the most fluvially organized, with deep
number low-energy and pool clusters, green for higher Froude number pools found at low-gradient outer bends of the main anabranch tran-
glide and riffle clusters. The deeper mainstem pool and riffle clusters sitioning into higher velocity riffle units in higher gradient sections.
are then differentiated from their more marginal shallow low-energy Areas where flow diverges into two or three anabranches are typi-
and glide counterparts as dark blue or dark green versus light blue cally wider, shallower environments associated with glides. However,
or light green. low-energy patches are the most prevalent in this morphology (30%
Figure 7 shows the mapped patterns of these clusters for all three of wetted area), due to several prominent low-gradient secondary
years at 10 m3 s−1 . At this discharge, geomorphic control on the flow channels and lateral transitions within the main channel driven by
patterns is most evident, and the arrangement of clusters has the asymmetric cross sectional shape. In 2013, the geomorphic disrup-
most relevance to aquatic organisms that depend on specific low-flow tion from the large flood event and transition to high width/depth
hydraulic conditions throughout most of the year. For all three geo- ratios manifests itself in reductions in pool and riffle clusters; more
morphic configurations, the clustering created spatially well-defined flow is carried through shallow, high-energy slope stream areas cor-
8 of 17 TAMMINGA AND EATON

TABLE 3 Summary of clusters and representative average hydraulic characteristics


Cluster Description Depth (m) Velocity (m) Froude (—) Friction slope (—) Specific discharge (m2 s−1 )

1 Low energy 0.20 0.23 0.21 0.0017 0.051


2 Pool 0.73 0.83 0.33 0.0021 0.58
3 Glide 0.29 0.76 0.51 0.0066 0.24
4 Riffle 0.45 1.39 0.67 0.012 0.64

post-flood morphology relative to the well-defined, elongate features


2.0
Fr = 0.1

.3

in the pre-flood conditions.


0.5

.7
Fr = 0

=0
Fr =
Flow patterns change at higher discharges, resulting in different pro-

Fr
Cluster
1.5 low energy portions and arrangements of patches throughout the reach. Figure 9
pool
shows changes in landscape metrics with discharge. Patterns of patch
glide
Depth (m)

riffle change with discharge are different between the 3 years, shown
1.0
most simply by the proportion of landscape occupied by each class.
Membership In 2012, all four classes are relatively equal in proportion (20–30%
0.4 of area) at all three discharges. In 2013, the low-energy and glide
0.5 0.6
0.8 units that dominate at 10 m3 s−1 become less common as discharge
increases and more pool and riffle-like conditions develop, with a
0.0 more equal distribution of the four classes by 50 m3 s−1 . In 2014, the
0 1 2 3 glide conditions that are most common at 10 m3 s−1 quickly give way
Velocity (m/s) to more pool and riffle areas, with riffle-like hydraulic conditions dom-

FIGURE 5 Depth–velocity scatterplot for classified 2012 m3 s−1 inating by 50 m3 s−1 . Secondary anabranches that at 10 m3 s−1 are
data, with points stratified by colour for each cluster. Point primarily low-energy or glide units tend to transition to pool or rif-
transparency is scaled by the membership of each point to its class. fle conditions with higher discharges as well, and the more marginal
Lines of equal Froude number are overlain for reference. Black low-energy or glide clusters are only found along shallow sloping bar
asterisks indicate centres for each cluster
edges, in backwater areas, or in smaller high-elevation channels that
had no flow at 10 m3 s−1 . Patch density and edge density display
similar trends with discharge for the three years: Pool and riffle units
responding to glide units (43% of area in 2013 vs. 24% in 2012).
remain less frequent in number with low-edge densities, whereas
Pool and riffle units still exist, but they are more sporadic and
low-energy and glide classes decline in patch and edge density as
less easily explained by planform configuration or bar–pool–riffle
discharge increases. Pools and riffles also show increases in patch
sequences. Low-energy patches remain common given the shallow
area with discharge as flow conditions coalesce into a quickly flow-
overall depths and several side channels. By 2014, hydraulic diversity
ing deep mainstem that is less influenced by topographic highs or
increases slightly relative to 2013 as pool and riffle conditions reform
lows. These overall changes are reflected in the Shannon diversity
in the main anabranch thalweg, but the reach is still dominated by
index (Figure 10), which is high for all discharges in 2012 and low-
glide-type flow conditions and low-energy environments are reduced
est in 2013 at 10 m3 s−1 but climbs with discharge as pool and riffle
in area.
units become more common. 2014 has intermediate Shannon Diver-
Figure 8 also includes several metrics of landscape pattern chosen
sity values for 10 m3 s−1 but reaches similar values to 2012 or 2013
to interpret ecologically relevant changes in spatial composition
by 30 m3 s−1 .
and shapes of the cluster patches. Patch density, which reflects the
number of patches present in a class, shows that at 10 m3 s−1 ,
low-energy and glide units are much more frequent compared with
pools or riffles in all 3 years. This generally corresponds to smaller
3.3 Brown trout habitat suitability
mean patch areas for low-energy and glide units, compared with the As a measure of total reach-scale suitability, habitat suitability changes
larger, more spatially contiguous pool and riffle units. However, the can be analysed in terms of average HS for each scenario. This also
flood impacts appear to have resulted in a muting of this pattern; patch facilitates easy examination of overall changes in habitat suitability
densities for low-energy and glide classes are reduced in 2013 and with discharge for the 3 years (Figure 11). Summing up habitat suitabil-
even more in 2014 as glide units in particular become larger and more ity values this way confirms an overall reduction in habitat availability
cohesive. These changes are also reflected in edge density, a measure and quality as a result of the 2013 flood, especially at low flow; adult
of edge length for each class. For all 3 years, the elongate, linear and juvenile HS are highest in 2012 at 10 m3 s−1 and then low in 2013
shape of the more marginal glide and low-energy units is quantified and 2014. In 2012, HS declines as discharge increases, especially for
as higher edge density; pools and riffles are generally more rounded juveniles as velocity tolerances are quickly exceeded. Because 2013
and simple-shaped with lower edge densities. Between 2012 and and 2014 have less suitable 10 m3 s−1 conditions to begin with, this
2013, reductions in mean patch area, patch density, and edge density drop-off with discharge is less strong; adult HS is largely constant with
for pool and riffle clusters likely reflect the disorganized nature of discharge and juvenile HS declines slightly between 30 and 50 m3 s−1 .
TAMMINGA AND EATON 9 of 17

FIGURE 6 Detailed example view of (a) 2013 orthomosaic, (b) specific discharge overlain on digital elevation model, and (c–f) low-energy, pool,
glide, riffle fuzzy unit patterns with transparency scaled by membership to each class

Figure 12 shows distributions of JuvHS and AdultHS at 10 m3 s−1 suitable for adult trout but slightly less due to a lower preference for
associated with each fuzzy c-means cluster to evaluate whether the very shallow depths. Pool habitats are the second most suitable for
hydraulically defined clusters display distinct habitat suitability ranges. juveniles who can inhabit deep waters, but the range of velocities
For juvenile trout, the low-energy cluster provides the most suitable within this cluster means some pool areas are largely inhospitable due
conditions with low velocities; small bodied juveniles can also occupy to low tolerance to velocities greater than about 0.5 m s−1 . Adult
the shallow depths in this cluster. These conditions are also highly trout can use higher velocity conditions, so the pool cluster provides
10 of 17 TAMMINGA AND EATON

FIGURE 7 Spatial patterns of clusters at 10 m3 s−1 . Clusters are overlain on detrended digital elevation models within the active channel

FIGURE 8 Measures of reach composition at 10 m3 s−1

consistently high-quality habitat. The higher Froude number glide and


riffle clusters are generally less suitable for both juveniles and adults; FIGURE 9 Measures of reach composition at all discharges
juveniles in particular cannot tolerate high-energy riffle units at all.
TAMMINGA AND EATON 11 of 17

of large floods reflects a channel form adjustment to a shallower,


wider flow that reduces mean boundary shear stress until the flood
pulse recedes or bank stability is reattained (Tamminga, Eaton, &
Hugenholtz, 2015a). However, the more detailed geomorphic impacts
and system responses of such widening are less understood and
depend on valley and floodplain configuration, geomorphic history, and
initial active channel geometry (Fuller, 2007; Lea & Legleiter, 2016).
At the Elbow study reach, bank erosion supplied large volumes of
alluvium to the channel bed, resulting in substantial local aggradation
despite the event being net degradational. This sediment and the
associated high flood transport rates erased the existing pre-flood
morphology completely as large bars were constructed and dissected
during the flood flows. These bar features have a long streamwise
length scale that likely reflects a much longer sediment path length (the
downstream displacement of particles over time) than the pre-flood
FIGURE 10 Changes in Shannon diversity index with discharge channel forming flows would have exhibited (Pyrce & Ashmore, 2003).
These depositional features are also relatively tall due to the deep
flows in which they formed (Church & Jones, 1982; Nicholas et al.,
2016). Similar effects have been reported by Wheaton et al. (2013),
who documented a shift in braiding mechanisms to favour channel
widening and transverse bar building as the magnitude and duration
of flows above bankfull increased. They postulated that the large
sediment loads and reduced topographic steering during large floods
result in a reduction of braidplain heterogeneity, consistent with
findings from Madej (1999) and Legleiter (2014).
The 2013 flood therefore set the geomorphic constraints for the
processes of readjustment during the following year (2013–2014).
Because the channel configuration was set by such a high-magnitude
FIGURE 11 Changes in average HS with discharge
event that built high-elevation barforms, the 2014 spring freshet
(which peaked at approximately 100 m3 s−1 ) was likely constricted to
zones within the oversteepened low-sinuosity path cut by the waning
4 DISCUSSION 2013 flood flows. Consequently, the 2014 geomorphic adjustment
reflects a lower flow magnitude and shorter duration above sediment
This study aimed to investigate the relationships between river mor- transport thresholds combined with topographic confinement, result-
phodynamics, flow organization, and aquatic habitat. Through repeat ing in less areally extensive geomorphic turnover. As opposed to the
surveys bracketing a large flood event, we characterized geomorphic significant lateral instability evident in the 2012–2013 changes, the
change resulting from the flood and subsequent readjustment in the 2013–2014 morphodynamics appear to be progressing towards the
following year. Overall, we found that restructuring of morphology re-establishment of lateral dynamic equilibrium through sedimenta-
strongly impacted flow patterns in both a reach average sense and tion in secondary channels, bank erosion, bar edge trimming, and unit
in terms of the spatio-temporal arrangement of zones with distinct bar accumulation that results in increases in sinuosity and a more
hydraulic characteristics. These changes resulted in a reduction in single thread channel pattern.
adult and juvenile brown trout habitat quality as measured through These large, channel forming events (2013 flood and 2014 freshet)
habitat suitability indices. configure the geomorphic template within which the lower magni-
The geomorphic adjustments due to the 2013 flood were drastic. tude flows that are more ecohydraulically relevant operate. These
The detailed pre- and post-flood topographic datasets document smaller discharges are what we focused on through numerical flow
spatial patterns of bank erosion, scour, and fill at a resolution that modelling in this study, allowing for investigation of how flow pat-
is rare for such high-magnitude disturbance events. In general, the terns are conditioned by geomorphic change. Overall, we found that
active channel widened by approximately 50 m between 2012 and the 2013 flows were very wide and shallow compared with the same
2013. This is consistent with many other examples of large floods discharges in the 2012 geomorphic configuration (Figure 3, Table 2).
in the geomorphic literature (Hooke, 2008; Kale, 2007; Magilligan, This shows that the reach-scale widening and local deposition associ-
1992; Righini et al., 2017) and reflects an adjustment to forces far ated with the 2013 flood are reflected in smaller discharges as well.
in excess of thresholds for entrainment of large grain sizes and bank This wider shallower flow is also less hydraulically diverse (Figure 4),
erosion (Eaton & Church, 2004; Olsen, Whitaker, & Potts, 1997). with a near complete loss of deep, slow areas at all three modelled
Similar patterns of bank erosion were also observed along the rest of discharges. Overall, it appears the 2013 flow patterns are less fluvially
Elbow River and on rivers throughout the region. This primary effect organized in the conventional sense; insufficient time of competent
12 of 17 TAMMINGA AND EATON

FIGURE 12 Distributions of JuvHS and AdultHS for each 2012 10-m3 s−1 classified cluster

flows following the flood means there is a lack of sediment transport by the geomorphic configuration in the pre-flood reach: alternating
and sorting processes that lead to shear stress concentration and pre- pool–riffle units in the main anabranch along with several substantial
dictable patterns of pools scour and bar deposition or avulsion and low-gradient side channels create a wide range of flow conditions. In
new anabranch creation (Burge, 2006; Harrison, Legleiter, Wydzga, & 2013, the lower hydraulic diversity and wide shallow flow conditions
Dunne, 2011; Whiting & Dietrich, 1993). As a result, the topographic result in an unequal proportional assignment of cluster memberships;
highs and lows or planform bifurcations that result in local differences the reach is dominated by low-energy and glide units at 10 m3 s−1 . By
in gradient and flow resistance are largely absent in the post-flood 2014, geomorphic adjustments result in a slight recovery of pool and
morphology; 2013 flows instead reflect the bar features and patterns riffle units relative to 2013, but diversity is still low. These differences
created by a completely different scale of flow event than those that in flow distributions and reachscape composition are most evident at
normally create hydraulic diversity in this system. 10 m3 s−1 ; conditions tend to converge to more equal proportions of
Although bed mobilization likely occurs somewhere between the classes by 30 and 50 m3 s−1 as small-scale geomorphic forms become
30 and 50 m3 s−1 discharges, patterns of elevated specific discharge at less important.
these discharges can hint at where geomorphic change will occur and A benefit of classifying flow data is that it allows for explicit
explain observed changes. In 2012, at 50 m3 s−1 , we see high specific investigation of spatial flow patterns. With the fuzzy c-means approach
discharges associated with outer bends and anabranch confluences used in this study and our analysis of the resulting clusters, we were
(Figure 3). It is likely that these zones allow for compensating erosion able to address measures of patch shapes and sizes that may reflect
and deposition whereas the channel's average dimensions remain underlying geomorphic features and have direct ecological relevance.
generally constant and that this mode of adjustment was dominant We found some patterns that are relatively universal in our flow
during the previous years before the 2013 flood. In the 2013 50-m3 s−1 models; low-energy and glide patches are generally smaller and more
modelled discharge, there is mainly one zone of elevated specific numerous, with more complex shapes and higher edge densities than
discharge in the upstream half of the reach; by 2014, this zone has pool or riffle patches (Figure 9). This effect is likely a combination
shifted to the right outer bank as lateral migration progresses. This is of the actual small scale of these hydromorphic units along with
associated with the re-establishment of shorter length-scale sediment a classification artefact associated with fitting these slender, linear
transport processes and the reorganization of coherent bar–pool–riffle stream's edge features into 1-m2 pixels (Wyrick & Pasternack, 2014).
sequences. However, the process is still ongoing, as evidenced by the Pool and riffle units, on the other hand, are generally more spatially
high shear stress in this configuration (Table 2) and the location of contiguous and less fragmented due to the strong flow organization
elevated forces at upstream end of outer bends. These more minor they are associated with, resulting in lower edge densities. However,
geomorphic adjustments between 2013 and 2014 result in slight the geomorphic change throughout the study period appears to modify
changes in flow distributions and diversity. these patterns to some degree; pre-flood pool and riffle units are more
Joint depth–velocity distributions provide a framework for elongate and linear at low flow than in 2013 or 2014. This is quantified
interpreting the hydraulic conditions that control classified patch com- by a decrease in edge density and mean patch area for these units
position and distributions. In 2012, the diverse range of depth–velocity as a result of the flood and reflects the disorganized nature of flow
3 −1
pairs (Figure 4) at 10 m s is reflected in the relatively even appor- conditions in the post-flood morphologies.
tioning of flow points into the four clusters (Figure 8). The diverse Because each of the classified clusters has distinct hydraulic char-
class distribution remains relatively constant as discharge increases acteristics, the resulting hydromorphic unit maps can be linked with
(Figure 9); 2012 flow conditions are therefore the most complex, species' preferences for spatial investigation of habitat suitability. In
as measured both on a continuous basis (high HMID, Figure 4) and the brown trout habitat context examined herein, applying prefer-
when interpreted based on patterns of discrete classified units (high ence curves to continuous depth and velocity data gives a traditional
Shannon diversity index, Figure 10). This diversity can be explained appraisal of habitat suitability throughout the reach. By this metric,
TAMMINGA AND EATON 13 of 17

adult and juvenile habitat suitability declined as a result of the 2013 ries. Although unraveling these linkages and the contrasting ecological
flood, especially at low flows. However, many fish species have been requirements for different age classes and species requires more
shown to have strong affinities for discrete habitat types (Nickelson, detailed direct information on spatiotemporal patterns of fish space
Rodgers, Johnson, & Solazzi, 1992; Rosenfeld & Boss, 2001; Schlosser, use on Elbow River specifically, the classification procedure used in
1991) that may relate to more than just the hydraulic conditions in this study provides a robust generic framework to address issues of
those habitat units. For example, availability of prey for drift-feeding the relationship between habitat dynamics for any lotic species and
salmonids is a primary driver of growth differences between slow and hydrogeomorphology.
fast water habitat units (Nielsen, 1992; Reid & Thoms, 2008; Rosen- The results of this study provide one way of quantifying the effects
feld & Raeburn, 2009). Spatial arrangements of habitats that generate of a large flood disturbance event. Although the short time period
benthic invertebrate drift that fish can feed on (riffles) versus energet- (3 years) covered in our UAV surveys makes prediction of future
ically favourable feeding locations (pools), along with factors such as morphodynamic trajectories difficult, the primary immediate effect of
pool length that affect rates of depletion of prey through deposition the flood was a homogenization of sub-bankfull flow conditions and
or consumption by individual fish occupying that unit, are therefore a decrease in hydraulic and habitat diversity. The readjustment dur-
important. Similarly, lateral velocity gradients such as fast currents ing the 1 year following the flood event appears to be progressing
close to velocity shelters are profitable in terms of energy gain (Fausch towards a meandering laterally stable single thread channel, with the
& White, 1981; Hughes & Dill, 1990), and the distribution of hydrauli- high elevation bar surfaces built during the 2013 flood acting as flood-
cally sheltered sites and suitable sediment size patches can control plain features largely disconnected from the low-flow habitats. This
spatial patterns of spawning suitability (Cienciala & Hassan, 2013). reduction in geomorphic complexity is in contrast with many exam-
For these reasons, predicting fish density based on discrete habitat ples of ecological responses to large infrequent disturbance events
classes is often more accurate than based on continuous variables in other ecosystem types, which are often cited as creating or main-
alone (Rosenfeld, 2003), especially when univariate habitat suitability taining heterogeneity (Foster, Knight, & Franklin, 1998; Romme &
criteria that ignore interactions between hydraulic variables are the Knight, 1982). However, other studies of large flood events have doc-
basis for assessment of depth and velocity suitability (Ayllón, Almodó- umented similar reductions in active channel geomorphic complexity
var, Nicola, & Elvira, 2009). In this context, the benefit of a bottom-up associated with large waves of sediment deposition (Legleiter, 2014;
statistical classification approach is that the clustering based on depth, Madej, 2001). In the case of Elbow River, the disturbance appears to
velocity, and Froude number integrates these factors and their interac- be a clock-resetting event that returns the system to an initial state
tions explicitly without a potentially myopic focus on species-specific with complete geomorphic and habitat turnover (Phillips & Van Dyke,
suitability criteria (Poole et al., 1997; Schweizer, Borsuk, Jowett, & 2016). This is different than examples of smaller magnitude flood
Reichert, 2007). This allows for an objective stratification of habi- disturbance dynamics, where shifting mosaic steady-state conditions
tat units that relates to both underlying geomorphic processes and may be maintained (Arscott, Tockner, Nat, & Ward, 2002). For Elbow
resulting habitat dynamics. River, smaller magnitude floods will be necessary to rework the geo-
At the Elbow River study site, the overall picture of a decrease morphic structure set by the 2013 flood, allowing for the recreation
in habitat suitability as the result of the 2013 flood can be refined of heterogeneity through vertical and lateral erosion, mesoscale bed-
by accounting for the spatial characteristics of the classified habitat form construction, and avulsion mechanisms to create new secondary
patches. For example, pool units declined in total areal proportion channels (Wheaton et al., 2013). Judging by the rate of bank erosion
at 10 m3 s−1 between 2012 and 2013 (Figure 8), largely driving the and bar building observed between 2013 and 2014 and the historical
observed decrease in HS. However, riffle units also decreased in pro- record of large flows in the region, the timescale for such readjustment
portion. Although riffle units themselves are not highly suitable habitat is likely on the scale of 50–100 years (Beechie, Liermann, Pollock,
when measured by hydraulic preference curves (Figure 12), their func- Baker, & Davies, 2006). However, potential changes in broader driv-
tion as a zone of benthic invertebrate prey production makes them ing conditions such as shifts in flood frequency or rates of vegetation
ecologically important to drift-feeding salmonids. The observed reduc- recolonization (Hervouet, Dunford, Piégay, Belletti, & Trémélo, 2011)
tions in mean patch area and edge density for riffle units at 10 m3 s−1 with climate change may affect recovery trajectories. Continued care-
could therefore be interpreted as further negative geomorphic con- ful monitoring of fluvial ecosystem conditions in an environmental
sequences of the flood that result in less prey availability and fewer change context will therefore be necessary to fully understand linked
suitable riffle-adjacent holding and feeding zones for brown trout. fluvial and ecological processes.
However, territorial requirements and competition between individ-
uals and different age classes are also mediated by spatial patterns
of habitat zones; Ayllón, Almodóvar, Nicola, and Elvira (2010) docu- 5 CONCLUSION
mented a positive scaling of territory size with brown trout body size,
but salmonid territory area typically decreases as food abundance Stream morphodynamics, hydraulic organization, and aquatic habitat
or habitat quality and complexity increase (Imre, Grant, & Keeley, are closely linked and sensitive to changes in governing conditions.
2002, 2004; Keeley, 2000; Venter, Grant, Noel, & Kim, 2008). The In this study, we took advantage of detailed datasets characteriz-
smaller, less numerous pool units and overall decrease in hydraulic ing river structural changes associated with a large flood event to
diversity in the post-flood morphologies could therefore further stress examine these interrelationships in a disturbance context. By apply-
salmonid populations due to increased competition for prime territo- ing an objective stream habitat unit classification method to modelled
14 of 17 TAMMINGA AND EATON

distributions of hydraulic variables, we mapped patterns of hydrogeo- Chapman, D. (1966). Food and space as regulators of salmonid populations
morphic features at a range of discharges for three unique geomorphic in streams. American Naturalist, 100(913), 345–357.

conditions bracketing the flood. We documented a general homoge- Cheng, T., Molenaar, M., & Lin, H. (2001). Formalizing fuzzy objects
from uncertain classification results. International Journal of Geo-
nization of flow conditions as a result of the flood associated with a graphical Information Science, 15(1), 27–42. https://doi.org/10.1080/
loss of pool and riffle units. The effects of geomorphic change on flow 13658810010004689
patterns were most evident at low discharges. These hydrogeomor- Christensen, J. H., Kanikicharla, K. K., Marshall, G., & Turner, J. (2013).
phic changes resulted in a decline in habitat suitability for adult and Climate phenomena and their relevance for future regional climate
change. In T. F. Stocker, D. Qin, G.-K. Plattner, M. M. B. Tignor, S.
juvenile brown trout when measured by conventional habitat suit-
K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex, & P. M. Midgley
ability criteria; we also showed that a spatially explicit assessment of (Eds.), Climate change 2013: The physical science basis. Contribution of
hydrogeomorphic conditions can complement and refine these tra- Working Group I to the Fifth Assessment of the Intergovernmental Panel
on Climate Change. (pp. 1217–1308). Cambridge: Cambridge University
ditional estimates of habitat impacts. Overall, this study provides an
Press. https://doi.org/10.1017/CBO9781107415324.028
integrative assessment of the mechanisms driving fluvial ecosystem
Church, M. (1980). Records of recent geomorphological events. In R.
dynamics in an environmental change context. Cullingford, D. Davidson, & J. Lewin (Eds.), Timescales in geomorphology
(pp. 13–28). Chichester, UK: Wiley.
ACKNOWLEDGEMENTS Church, M., & Jones, D. (1982). Channel bars in gravel bed rivers. In R.
D. Hey, J. C. Bathurst, & C. Thorne (Eds.), Gravel bed rivers (pp. 1–26).
This research is part of the Natural Sciences and Engineering Research Chichester, UK: Wiley.
Council's HydroNet Strategic Network Program. We thank Owen Cienciala, P., & Hassan, M. A. (2013). Linking spatial patterns of bed surface
Brown, Steve Myshak, and Jordan Walker for assisting with the texture, bed mobility, and channel hydraulics in a mountain stream to
potential spawning substrate for small resident trout. Geomorphology,
UAV surveys and photogrammetric processing and Dave Hutchin-
197, 96–107. https://doi.org/10.1016/j.geomorph.2013.04.041
son/Headwater Analytics for the flood frequency analysis code. The
Clifford, N., Acreman, M., & Booker, D. (2008). Hydrological and hydraulic
authors declare no conflicts of interest. aspects of river restoration uncertainty for ecological purposes. In S.
Darby, & D. Sear (Eds.), River restoration: Managing the uncertainty in
restoring physical habitat (pp. 105–138). Chichester, UK: John Wiley &
ORCID
Sons Incorporated. https://doi.org/10.1002/9780470867082.ch7

Aaron Tamminga http://orcid.org/0000-0002-8116-8731 Courtney, R. F., Walder, G. L., & Vadas, R. L. (1998). Instream flow
requirements for fish in the Kananaskis River. Calgary: AB.
Crowder, D. W., & Diplas, P. (2002). Assessing changes in watershed flow
REFERENCES regimes with spatially explicit hydraulic models. JAWRA Journal of the
American Water Resources Association, 38(2), 397–408. https://doi.org/
Angermeier, P. L., & Schlosser, I. J. (1989). Species-area relationship for
10.1111/j.1752-1688.2002.tb04325.x
stream fishes. Ecology, 70(5), 1450–1462. https://doi.org/10.2307/
1938204 Crowder, D. W., & Diplas, P. (2006). Applying spatial hydraulic principles to
quantify stream habitat. River Research and Applications, 22(1), 79–89.
Armstrong, J. D., Kemp, P. S., Kennedy, G., Ladle, M., & Milner, N. J. (2003).
https://doi.org/10.1002/rra.893
Habitat requirements of Atlantic salmon and brown trout in rivers and
streams. Fisheries Research, 62(2), 143–170. https://doi.org/10.1016/ Dunbar, M. J., Alfredsen, K., & Harby, A. (2011). Hydraulic-habitat modelling
S0165-7836(02)00160-1 for setting environmental river flow needs for salmonids. Fisheries
Management and Ecology, 19(6), 500–517. https://doi.org/10.1111/j.
Arscott, D. B., Tockner, K., van der Nat, D., & Ward, J. V. (2002). Aquatic
1365-2400.2011.00825.x
habitat dynamics along a braided alpine river ecosystem (Tagliamento
River, northeast Italy). Ecosystems, 5(8), 0802–0814. https://doi.org/ Eaton, B. C., & Church, M. (2004). A graded stream response relation
10.1007/s10021-002-0192-7 for bed load–dominated streams. Journal of Geophysical Research: Earth
Surface, 109, F03011. https://doi.org/10.1029/2003JF000062
Ayllón, D., Almodóvar, A., Nicola, G. G., & Elvira, B. (2009). Interactive
effects of cover and hydraulics on brown trout habitat selection pat- Fausch, K., Torgersen, C., Baxter, C., & Li, H. (2002). Landscapes to
terns. River Research and Applications, 25(8), 1051–1065. https://doi. riverscapes: Bridging the gap between research and conservation
org/10.1002/rra.1215 of stream fishes. BioScience, 52(6), 483. https://doi.org/10.1641/
0006-3568(2002)052[0483:LTRBTG]2.0.CO;2
Ayllón, D., Almodóvar, A., Nicola, G. G., & Elvira, B. (2010). Modelling
brown trout spatial requirements through physical habitat simulations. Fausch, K. D., & White, R. J. (1981). Competition between brook trout
River Research and Applications, 26(9), 1090–1102. https://doi.org/10. (Salvelinus fontinalis) and brown trout (Salmo trutta) for positions in
1002/rra.1315 a Michigan stream. Canadian Journal of Fisheries and Aquatic Sciences,
38(10), 1220–1227. https://doi.org/10.1139/f81-164
Beechie, T. J., Liermann, M., Pollock, M. M., Baker, S., & Davies, J. (2006).
Channel pattern and river-floodplain dynamics in forested mountain Ferguson, R. (2010). Time to abandon the Manning equation? Earth Surface
river systems. Geomorphology, 78(1-2), 124–141. https://doi.org/10. Processes and Landforms, 35(15), 1873–1876. https://doi.org/10.1002/
1016/j.geomorph.2006.01.030 esp.2091
Bezdek, J. C., Ehrlich, R., & Full, W. (1984). FCM: The fuzzy c-means Fisher, P. F. (1998). Is GIS hidebound by the legacy of cartogra-
clustering algorithm. Computers & Geosciences, 10(2), 191–203. https:// phy? The Cartographic Journal, 35(1), 5–9. https://doi.org/10.1179/
doi.org/10.1016/0098-3004(84)90020-7 000870498787074056
Burge, L. M. (2006). Stability, morphology and surface grain size patterns of Fonstad, M., Dietrich, J., Courville, B., Jensen, J., & Carbonneau, P. (2013).
channel bifurcation in gravel–cobble bedded anabranching rivers. Earth Topographic structure from motion: A new development in photogram-
Surface Processes and Landforms, 31(10), 1211–1226. https://doi.org/ metric measurement. Earth Surface Processes and Landforms, 38(4),
10.1002/esp.1325 421–430. https://doi.org/10.1002/esp.3366
Carbonneau, P., Fonstad, M. A., Marcus, W. A., & Dugdale, S. J. (2012). Foster, D. R., Knight, D. H., & Franklin, J. F. (1998). Landscape patterns and
Making riverscapes real. Geomorphology, 137(1), 74–86. https://doi. legacies resulting from large, infrequent forest disturbances. Ecosystems,
org/10.1016/j.geomorph.2010.09.030 1(6), 497–510. https://doi.org/10.1007/s100219900046
TAMMINGA AND EATON 15 of 17

Fryirs, K. A. (2017). River sensitivity: A lost foundation concept in fluvial Javernick, L., Hicks, D., Measures, R., Caruso, B., & Brasington,
geomorphology. Earth Surface Processes and Landforms, 42(1), 55–70. J. (2015). Numerical modelling of braided rivers with structure-
https://doi.org/10.1002/esp.3940 from-motion-derived terrain models. River Research and Applications,
Fuller, I. C. (2007). Geomorphic work during a ‘‘150-year’’ storm: Contrast- 32(5), 1071–1081. https://doi.org/10.1002/rra.2918
ing behaviors of river channels in a New Zealand catchment. Annals of Kale, V. S. (2007). Geomorphic effectiveness of extraordinary floods
the Association of American Geographers, 97(4), 665–676. https://doi. on three large rivers of the Indian Peninsula. Geomorphology, 85(3),
org/10.1111/j.1467-8306.2007.00576.x 306–316. https://doi.org/10.1016/j.geomorph.2006.03.026
Gartner, J. D., Dade, W. B., Renshaw, C. E., Magilligan, F. J., & Buraas, E. Keeley, E. R. (2000). An experimental analysis of territory size in juvenile
M. (2015). Gradients in stream power influence lateral and downstream steelhead trout. Animal Behaviour, 59(3), 477–490. https://doi.org/10.
sediment flux in floods. Geology, 43(11), 983. https://doi.org/10.1130/ 1006/anbe.1999.1288
G36969.1 Kemp, J. L., Harper, D. M., & Crosa, G. A. (1999). Use of ‘‘functional habitats’’
Gaston, K. J., & Blackburn, T. M. (1999). A critique for macroecology. Oikos, to link ecology with morphology and hydrology in river rehabilitation.
84(3), 353–368. https://doi.org/10.2307/3546417 Aquatic Conservation: Marine and Freshwater Ecosystems, 9(1), 159–178.
https://doi.org/10.1002/(SICI)1099-0755(199901/02)9:1<159::AID-
Gende, S. M., Edwards, R. T., Willson, M. F., & Wipfli, M. S. (2002). Pacific
AQC319>3.0.CO;2-M
salmon in aquatic and terrestrial ecosystems: Pacific salmon subsidize
freshwater and terrestrial ecosystems through several pathways, Krapesch, G., Hauer, C., & Habersack, H. (2011). Scale orientated analysis
which generates unique management and conservation issues but also of river width changes due to extreme flood hazards. Natural Hazards
provides valuable research opportunities. BioScience, 52(10), 917–928. and Earth System Science, 11(8), 2137–2147. https://doi.org/10.5194/
https://doi.org/10.1641/0006-3568(2002)052[0917:PSIAAT]2.0.CO;2 nhess-11-2137-2011

Gostner, W., Alp, M., Schleiss, A. J., & Robinson, C. T. (2013). The Lancaster, J., & Downes, B. J. (2010). Linking the hydraulic world of
hydro-morphological index of diversity: A tool for describing habi- individual organisms to ecological processes: Putting ecology into eco-
tat heterogeneity in river engineering projects. Hydrobiologia, 712(1), hydraulics. River Research and Applications, 26(4), 385–403. https://doi.
43–60. https://doi.org/10.1007/s10750-012-1288-5 org/doi.org/10.1002/rra.1274

Harrison, L., Legleiter, C., Wydzga, M., & Dunne, T. (2011). Channel Lane, S. (2000). The measurement of river channel morphology using
dynamics and habitat development in a meandering, gravel bed river. digital photogrammetry. The Photogrammetric Record, 16(96), 937–961.
Water Resources Research, 47(4), W04513. https://doi.org/10.1029/ https://doi.org/10.1111/0031-868X.00159
2009WR008926 Lea, D. M., & Legleiter, C. J. (2016). Mapping spatial patterns of stream
power and channel change along a gravel-bed river in northern
Harrison, L., Pike, A., & Boughton, D. A. (2017). Coupled geomorphic
yellowstone. Geomorphology, 252, 66–79. https://doi.org/10.1016/j.
and habitat response to a flood pulse revealed by remote sensing.
geomorph.2015.05.033
Ecohydrology, 10, e1845. https://doi.org/10.1002/eco.1845
Leclerc, M. (2005). Ecohydraulics: A new interdisciplinary frontier for CFD.
Hauer, C., & Habersack, H. (2009). Morphodynamics of a 1000-year flood
In P. D. Bates, S. N. Lane, & R. I. Ferguson (Eds.), Computational
in the Kamp River, Austria, and impacts on floodplain morphology. Earth
fluid dynamics: applications in environmental hydraulics (pp. 429–460).
Surface Processes and Landforms, 34(5), 654–682. https://doi.org/10.
Chichester, UK: Wiley.
1002/esp.1763
Leclerc, M., Boudreault, A., Bechara, T. A., & Corfa, G. (1995).
Hauer, C., Mandlburger, G., Schober, B., & Habersack, H. (2012). Mor-
Two-dimensional hydrodynamic modeling: A neglected tool in the
phologically related integrative management concept for reconnecting
instream flow incremental methodology. Transactions of the Amer-
abandonded channels based on airborne LiDAR data and habitat mod-
ican Fisheries Society, 124(5), 645–662. https://doi.org/10.1577/
eling. River Research and Applications, 30(5), 537–556. https://doi.org/
1548-8659(1995)124%3C0645:TDHMAN%3E2.3.CO;2
10.1002/rra.2593
Legleiter, C. J. (2014). A geostatistical framework for quantifying the
Heggenes, J., & Saltveit, S. J. (1990). Seasonal and spatial microhabitat selec- reach-scale spatial structure of river morphology: 2. Application to
tion and segregation in young Atlantic salmon, Salmo salar L., and brown restored and natural channels. Geomorphology, 205, 85–101. https://
trout, Salmo trutta L., in a Norwegian River. Journal of Fish Biology, 36(5), doi.org/10.1016/j.geomorph.2012.01.017
707–720. https://doi.org/10.1111/j.1095-8649.1990.tb04325.x
Legleiter, C. J., & Goodchild, M. F. (2005). Alternative represen-
Hervouet, A., Dunford, R., Piégay, H., Belletti, B., & Trémélo, M.-L. (2011). tations of in-stream habitat: Classification using remote sensing,
Analysis of post-flood recruitment patterns in braided-channel rivers at hydraulic modeling, and fuzzy logic. International Journal of Geo-
multiple scales based on an image series collected by unmanned aerial graphical Information Science, 19(1), 29–50. https://doi.org/10.1080/
vehicles, ultra-light aerial vehicles, and satellites. GIScience & Remote 13658810412331280220
Sensing, 48(1), 50–73. https://doi.org/10.2747/1548-1603.48.1.50
Madej, M. A. (1999). Temporal and spatial variability in thalweg profiles
Hooke, J. (2008). Temporal variations in fluvial processes on an active of a gravel-bed river. Earth Surface Processes and Landforms, 24(12),
meandering river over a 20-year period. Geomorphology, 100(1), 3–13. 1153–1169. https://doi.org/10.1002/(SICI)1096-9837(199911)24:12
https://doi.org/10.1016/j.geomorph.2007.04.034 %3C1153::AID-ESP41%3E3.0.CO;2-8
Hughes, N. F., & Dill, L. M. (1990). Position choice by drift-feeding Madej, M. A. (2001). Development of channel organization and rough-
salmonids—Model and test for Arctic grayling (Thymallus arcticus) ness following sediment pulses in single-thread, gravel bed rivers.
in sub-arctic mountain streams, Interior Alaska. Journal of the Fisheries Water Resources Research, 37(8), 2259–2272. https://doi.org/10.1029/
Board of Canada, 47(10), 2039–2048. https://doi.org/10.1139/f90-228 2001WR000229
Imre, I., Grant, J., & Keeley, E. (2002). The effect of visual isolation Madsen, T., Enevoldsen, H., & Jørgensen, T. (1993). Effects of water
on territory size and population density of juvenile rainbow trout velocity on photosynthesis and dark respiration in submerged stream
(Oncorhynchus mykiss). Canadian Journal of Fisheries and Aquatic macrophytes. Plant, Cell & Environment, 16(3), 317–322. https://doi.
Sciences, 59(2), 303–309. https://doi.org/10.1139/f02-010 org/10.1111/j.1365-3040.1993.tb00875.x
Imre, I., Grant, J., & Keeley, E. (2004). The effect of food abundance Magilligan, F. (1992). Thresholds and the spatial variability of flood power
on territory size and population density of juvenile steelhead trout during extreme floods. Geomorphology, 5(3), 373–390. https://doi.org/
(Oncorhynchus mykiss). Oecologia, 138(3), 371–378. 10.1016/0169-555X(92)90014-F
Javernick, L., Brasington, J., & Caruso, B. (2014). Modeling the topography Magilligan, F., Buraas, E., & Renshaw, C. (2015). The efficacy of stream
of shallow braided rivers using structure-from-motion photogrammetry. power and flow duration on geomorphic responses to catastrophic
Geomorphology, 213, 166–182. https://doi.org/10.1016/j.geomorph. flooding. Geomorphology, 228(Supplement C), 175–188. https://doi.
2014.01.006 org/10.1016/j.geomorph.2014.08.016
16 of 17 TAMMINGA AND EATON

Milrad, S. M., Gyakum, J. R., & Atallah, E. H. (2015). A meteorological Poole, G. C., Frissell, C. A., & Ralph, S. C. (1997). In-stream habitat unit
analysis of the 2013 Alberta flood: Antecedent large-scale flow pattern classification: Inadequacies for monitoring and some consequences for
and synoptic–dynamic characteristics. Monthly Weather Review, 143(7), management. JAWRA Journal of the American Water Resources Asso-
2817–2841. https://doi.org/10.1175/MWR-D-14-00236.1 ciation, 33(4), 879–896. https://doi.org/10.1111/j.1752-1688.1997.
Morantz, D. L., Sweeny, R. K., Shirvell, C. S., & Longard, D. A. (1987). tb04112.x
Selection of microhabitat in summer by juvenile Atlantic salmon (Salmo Post, J. R., & Johnston, F. D. (2002). Status of the bull trout (Salvelinus
salar). Journal of the Fisheries Board of Canada, 44(1), 120–129. https:// confluentus) in Alberta. Alberta Sustainable Resource Development, Fish
doi.org/10.1139/f87-015 and Wildlife Division, and Alberta Conservation Association, Wildlife Status
Nakano, S., & Murakami, M. (2001). Reciprocal subsidies: Dynamic inter- Report No. 39. Edmonton, AB. (40).
dependence between terrestrial and aquatic food webs. Proceedings of Pyrce, R., & Ashmore, P. (2003). The relation between particle path
the National Academy of Sciences of the United States of America, 98(1), length distributions and channel morphology in gravel-bed streams: A
166–170. https://doi.org/10.1073/pnas.98.1.166 synthesis. Geomorphology, 56(1-2), 167–187. https://doi.org/10.1016/
Naylor, L. A., Spencer, T., Lane, S. N., Darby, S. E., Magilligan, F. J., S0169-555X(03)00077-1
Macklin, M. G., & Möller, I. (2017). Stormy geomorphology: Geomorphic Reid, M., & Thoms, M. (2008). Surface flow types, near-bed hydraulics and
contributions in an age of climate extremes. Earth Surface Processes and the distribution of stream macroinvertebrates. Biogeosciences Discus-
Landforms, 42(1), 166–190. https://doi.org/10.1002/esp.4062 sions, 5(2), 1175–1204. https://doi.org/10.5194/bg-5-1043-2008
Nelson, J., & McDonald, R. (1996). Mechanics and modeling of flow Righini, M., Surian, N., Wohl, E., Marchi, L., Comiti, F., Amponsah, W.,
and bed evolution in lateral separation eddies, report, Grand Canyon & Borga, M. (2017). Geomorphic response to an extreme flood in
Environmental Studies. Glen Canyon Environmental Studies Report, 69. two Mediterranean rivers (northeastern Sardinia, Italy): Analysis of
Nelson, J. S., & Paetz, M. J. (1992). The fishes of Alberta. Edmonton, AB: controlling factors. Geomorphology, 290, 184–199. https://doi.org/10.
University of Alberta Press. 1016/j.geomorph.2017.04.014

Nelson, J. M., Shimizu, Y., Abe, T., Asahi, K., Gamou, M., Inoue, T., & Romme, W. H., & Knight, D. H. (1982). Landscape diversity: The concept
… Kimura, I. (2016). The international river interface cooperative: Public applied to Yellowstone Park. BioScience, 32(8), 664–670. https://doi.
domain flow and morphodynamics software for education and appli- org/10.2307/1308816
cations. Advances in Water Resources, 93, 62–74. https://doi.org/10. Rosenfeld, J. S. (2003). Assessing the habitat requirements of stream fishes:
1016/j.advwatres.2015.09.017 An overview and evaluation of different approaches. Transactions of the
Newson, M., & Newson, C. (2000). Geomorphology, ecology and river American Fisheries Society, 132(5), 953–968. https://doi.org/10.1577/
channel habitat: Mesoscale approaches to basin-scale challenges. T01-126
Progress in Physical Geography, 24(2), 195–217. https://doi.org/10. Rosenfeld, J. S., & Boss, S. (2001). Fitness consequences of habitat use
1177/030913330002400203 for juvenile cutthroat trout: Energetic costs and benefits in pools and
Nicholas, A. P., Smith, G. H. S., Amsler, M. L., Ashworth, P. J., Best, J. L., riffles. Canadian Journal of Fisheries and Aquatic Sciences, 58(3), 585–593.
Hardy, R. J., & … Reesink, A. J. (2016). The role of discharge variability https://doi.org/10.1139/f01-019
in determining alluvial stratigraphy. Geology, 44(1), 3–6. https://doi. Rosenfeld, J. S., & Raeburn, E. (2009). Effects of habitat and internal
org/10.1130/G37215.1 prey subsidies on juvenile coho salmon growth: Implications for stream
Nickelson, T. E., Rodgers, J. D., Johnson, S. L., & Solazzi, M. F. (1992). productive capacity. Ecology of Freshwater Fish, 18(4), 572–584. https://
Seasonal changes in habitat use by juvenile coho salmon (Oncorhynchus doi.org/10.1111/j.1600-0633.2009.00372.x
kisutch) in Oregon coastal streams. Canadian Journal of Fisheries and Schlosser, I. J. (1991). Stream fish ecology: A landscape perspective.
Aquatic Sciences, 49(4), 783–789. https://doi.org/10.1139/f92-088 BioScience, 41(10), 704–712. https://doi.org/10.2307/1311765
Nielsen, J. L. (1992). Microhabitat-specific foraging behavior, diet, Schweizer, S., Borsuk, M. E., Jowett, I., & Reichert, P. (2007). Predicting
and growth of juvenile coho salmon. Transactions of the American joint frequency distributions of depth and velocity for instream habitat
Fisheries Society, 121(5), 617–634. https://doi.org/10.1577/1548- assessment. River Research and Applications, 23(3), 287–302. https://
8659(1992)121<0617:MFBDAG>2.3.CO;2 doi.org/10.1002/rra.980
Olsen, D., Whitaker, A., & Potts, D. F. (1997). Assessing stream channel Silva, D. R. O., Ligeiro, R., Hughes, R. M., & Callisto, M. (2014).
stability thresholds using flow competence estimates at bankful stage. Visually determined stream mesohabitats influence benthic macroin-
JAWRA Journal of the American Water Resources Association, 33(6), vertebrate assessments in headwater streams. Environmental Moni-
1197–1207. https://doi.org/j.1752-1688.1997.tb03546.x toring and Assessment, 186(9), 5479–5488. https://doi.org/10.1007/
Papanicolaou, A., Elhakeem, M., & Wardman, B. (2010). Calibration and s10661-014-3797-3
verification of a 2D hydrodynamic model for simulating flow around Spencer, T., & Lane, S. N. (2017). Reflections on the IPCC and global change
emergent bendway weir structures. Journal of Hydraulic Engineer- science: Time for a more (physical) geographical tradition. The Canadian
ing, 137(1), 75–89. https://doi.org/10.1061/(ASCE)HY.1943-7900. Geographer / Le Géographe Canadien, 61(1), 124–135. https://doi.org/
0000280 10.1111/cag.12332
Parasiewicz, P. (2001). MesoHABSIM: A concept for application of instream Statzner, B., Gore, J. A., & Resh, V. H. (1988). Hydraulic stream ecol-
flow models in river restoration planning. Fisheries, 26(9), 6–13. https:// ogy - Observed patterns and potential applications. Journal of the
doi.org/10.1577/1548-8446(2001)026%3C0006:M%3E2.0.CO;2 North American Benthological Society, 7(4), 307–360. https://doi.org/
Pasternack, G. B., Wang, C. L., & Merz, J. E. (2004). Application of a 10.2307/1467296
2D hydrodynamic model to design of reach-scale spawning gravel Tamminga, A., Eaton, B., & Hugenholtz, C. (2015a). UAS-based remote
replenishment on the Mokelumne River, California. River Research sensing of fluvial change following an extreme flood event. Earth Surface
and Applications, 20(2), 205–225. https://doi.org/doi.org/10.1002/rra. Processes and Landforms, 40(11), 1464–1476. https://doi.org/10.1002/
748 esp.3728
Phillips, J. D., & Van Dyke, C. (2016). Principles of geomorphic disturbance Tamminga, A., Hugenholtz, C., Eaton, B., & Lapointe, M. (2015b). Hyper-
and recovery in response to storms. Earth Surface Processes and spatial remote sensing of channel reach morphology and hydraulic fish
Landforms, 41(7), 971–979. https://doi.org/10.1002/esp.3912 habitat using an unmanned aerial vehicle (UAV): A first assessment
Pomeroy, J. W. (2015). The 2013 flood event in the South Saskatchewan in the context of river research and management. River Research and
and Elk River basins: Causes, assessment and damages. Canadian Water Applications, 31(3), 379–391. https://doi.org/10.1002/rra.2743
Resources Journal, 41(1-2), 1–13. https://doi.org/10.1080/07011784. Thompson, C., & Croke, J. (2013). Geomorphic effects, flood power, and
2015.1089190 channel competence of a catastrophic flood in confined and unconfined
TAMMINGA AND EATON 17 of 17

reaches of the upper Lockyer valley, southeast Queensland, Australia. Whiting, P. J., & Dietrich, W. E. (1993). Experimental constraints on
Geomorphology, 197, 156–169. https://doi.org/10.1016/j.geomorph. bar migration through bends: Implications for meander wavelength
2013.05.006 selection. Water Resources Research, 29(4), 1091–1102. https://doi.org/
Turner, M., & Dale, V. H. (1998). Comparing large, infrequent disturbances: 10.1029/92WR02356
What have we learned? Ecosystems, 1(6), 493–496. Wiens, J. A. (2002). Riverine landscapes: Taking landscape ecology into the
Van Horne, B. (1983). Density as a misleading indicator of habitat quality. water. Freshwater Biology, 47(4), 501–515. https://doi.org/10.1046/j.
The Journal of Wildlife Management, 47(4), 893–901. https://doi.org/10. 1365-2427.2002.00887.x
2307/3808148 Williams, R., Brasington, J., Hicks, M., Measures, R., Rennie, C., & Veri-
Vaughan, I. P., Diamond, M., Gurnell, A. M., Hall, K. A., Jenkins, A., Milner, N. cat, D. (2013). Hydraulic validation of two-dimensional simulations of
J., & … Ormerod, S. J. (2009). Integrating ecology with hydromorphol- braided river flow with spatially continuous aDcp data. Water Resources
ogy: A priority for river science and management. Aquatic Conservation: Research, 49(9), 5183–5205. https://doi.org/10.1002/wrcr.20391
Marine and Freshwater Ecosystems, 19(1), 113–125. https://doi.org/10. Woodget, A., Carbonneau, P., Visser, F., & Maddock, I. (2014). Quanti-
1002/aqc.895 fying submerged fluvial topography using hyperspatial resolution UAS
Venter, O., Grant, J. W., Noel, M. V., & Kim, J.-W. (2008). Mechanisms imagery and structure from motion photogrammetry. Earth Surface
underlying the increase in young-of-the-year Atlantic salmon (Salmo Processes and Landforms, 40(1), 47–64. https://doi.org/10.1002/esp.
salar) density with habitat complexity. Canadian Journal of Fisheries and 3613
Aquatic Sciences, 65(9), 1956–1964. https://doi.org/10.1139/F08-106 Woodget, A., Visser, F., Maddock, I., & Carbonneau, P. (2016). The accuracy
Vericat, D., Brasington, J., Wheaton, J., & Cowie, M. (2008). Accu- and reliability of traditional surface flow type mapping: Is it time for a
racy assessment of aerial photographs acquired using lighter-than-air new method of characterizing physical river habitat?River Research and
blimps: Low-cost tools for mapping river corridors. River Research and Applications, 32(9), 1902–1914. https://doi.org/10.1002/rra.3047
Applications, 25(8), 985–1000. https://doi.org/10.1002/rra.1198 Wyrick, J. R., & Pasternack, G. (2014). Geospatial organization of fluvial
Ward, J. V. (1998). Riverine landscapes: Biodiversity patterns, distur- landforms in a gravelcobble river: Beyond the rifflepool couplet. Geo-
bance regimes, and aquatic conservation. Biological Conservation, 83(3), morphology, 213, 48–65. https://doi.org/10.1016/j.geomorph.2013.
269–278. https://doi.org/10.1016/S0006-3207(97)00083-9 12.040

Wheaton, J., Brasington, J., Darby, S., Kasprak, A., Sear, D., & Veri-
cat, D. (2013). Morphodynamic signatures of braiding mechanisms as
expressed through change in sediment storage in a gravel-bed river.
How to cite this article: Tamminga A, Eaton B.
Journal of Geophysical Research: Earth Surface, 118(2), 759–779. https://
doi.org/10.1002/jgrf.20060 Linking geomorphic change due to floods to spatial
Wheaton, J. M., Fryirs, K. A., Brierley, G., Bangen, S. G., Bouwes, N., hydraulic habitat dynamics. Ecohydrology. 2018, e2018.
& O'Brien, G. (2015). Geomorphic mapping and taxonomy of fluvial https://doi.org/10.1002/eco.2018
landforms. Geomorphology, 248(C), 273–295. https://doi.org/10.1016/
j.geomorph.2015.07.010

You might also like