downstream variability of fluvial form, process, and character in a

216
Downstream Variability of Fluvial Form, Process, and Character in a Small Deglaciated Watershed, Southern Ontario By James Benjamin Thayer A thesis submitted in conformity with the requirements for the degree of Master of Science Graduate Department of Geography The University of Toronto © Copyright by James Benjamin Thayer, 2012

Upload: lehanh

Post on 03-Feb-2017

219 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Downstream Variability of Fluvial Form, Process, and Character in a

Downstream Variability of Fluvial Form, Process, and Character in a Small

Deglaciated Watershed, Southern Ontario

By

James Benjamin Thayer

A thesis submitted in conformity with the requirements

for the degree of Master of Science

Graduate Department of Geography

The University of Toronto

© Copyright by James Benjamin Thayer, 2012

Page 2: Downstream Variability of Fluvial Form, Process, and Character in a

ii

Downstream Variability of Fluvial Form, Process, and Character in a Small

Deglaciated Watershed, Southern Ontario

James Benjamin Thayer

Master of Science, 2012

Department of Geography, The University of Toronto

Abstract

Many deglaciated watersheds possess complex longitudinal profiles and spatially

variable sediment sources derived from the last glaciation. Accordingly, downstream

patterns in fluvial forms and characteristics of the channel and floodplain may diverge

from conventional ‘graded’ watershed models where channel slope decreases

downstream. Within a small, deglaciated watershed, five distinct fluvial forms were

identified and it was found that the watershed is organized in a spatially variable, but

generally inverted arrangement with stable, low-energy forms dominating the upper

watershed and more dynamic, high-energy forms in the lower watershed. As a

consequence of this inverted and variable downstream succession, and the spatially

erratic organization of glacial sediment sources, downstream trends in channel and

floodplain characteristics are poorly defined, and in many cases, deviate from expected

trends. This is most true for sedimentological variables where multiple co-dominant

controls exist, while morphological and hydrological variables better conform to

expected downstream trends.

Page 3: Downstream Variability of Fluvial Form, Process, and Character in a

iii

Acknowledgements

This thesis wouldn’t have come to fruition if it weren’t for the support that I

received over the past two years from various people. I’d first like to thank my advisor,

Dr. Joe Desloges, for generously funding this project. His critiques and suggestions on

various drafts have improved upon the final version of this thesis, and our discussions,

where he challenged many of the ideas presented here, helped me gain a stronger grasp of

the various facets of fluvial geomorphology. I thank my committee members, Drs. Joe

Desloges, Jing Chen, and Carl Mitchell, for providing valuable comments on areas that I

neglected during my writing. Vicki MacDonald of Rouge Park and Lori Colussi and

Edlyn Wong, both of the Toronto and Region Conservation Authority, provided

assistance in gaining and permitting access to the various study sites.

I extend my deepest gratitude to my various field assistants who helped me collect

way too much (though never enough!!) data that I gleefully got to spend months sifting

through. I’d especially like to thank Joyce Arabian and Betty Opalinska who bore the

brunt of the fieldwork and had to contend with leaky chest waders that were many sizes

too large and all of my wonderful idiosyncrasies over the course of the summer. I’d like

to thank my brother, Thomas, a.k.a., rock-skipper extraordinaire, for coming up to

Toronto for a week to help me with fieldwork for the second time in three summers.

Quite the enjoyable experience as usual…except for the part where we were out in the

middle of a field with a 4-metre tall lightning rod with a thunderstorm quick on our heels,

and the 2 hour (at least!) long downpour where we had to spend it in the car, and then the

intermittent downpours when we were out on a gravel bar and just had to sit through it,

and then the ESP kept sticking. Oh, what joyous times indeed! Lastly, Jen Henshaw

Page 4: Downstream Variability of Fluvial Form, Process, and Character in a

iv

bravely crawled into the leaky chest waders and into barely-above-freezing water to

collect some more data in October. Thank you all.

I’d like to thank my officemates Roger Phillips and Jen Henshaw for giving me

people to bother from time-to-time. I’d especially like to thank Roger for giving me

helpful advice these past couple years and for our various fluvial and non-fluvial related

discussions. He enthusiastically helped me with my initial ArcGIS inquiries, trying to get

Arc onto my computer, and sending me a lot of data for this project.

Lastly, I’d like to thank my parents for financially supporting me this past year,

without which I wouldn’t have been able to complete my second year and you wouldn’t

be able to read this thesis.

Page 5: Downstream Variability of Fluvial Form, Process, and Character in a

v

Table of Contents Abstract ............................................................................................................................. ii Acknowledgements .......................................................................................................... iii Table of Contents ............................................................................................................... v List of Figures ................................................................................................................. vii List of Tables ..................................................................................................................... x List of Appendices ........................................................................................................... xi Chapter One: Introduction .................................................................................. 1

1.1 Definition of the Problem .......................................................................... 1 1.2 Research Questions .................................................................................... 2

Chapter Two: Equilibrium and Downstream River Adjustment .................... 4

2.1 The Concept of Equilibrium in River Systems .......................................... 4 2.1.1 Scales of Equilibrium ..................................................................... 5

2.2 Adjustment of the Equilibrium Channel Form .......................................... 7 2.2.1 Predictive Methods of Channel Adjustment in Gravel-Bed Rivers 8 2.2.2 Controls on Channel Morphology: Direction and Magnitude ..... 10

2.2.2.1 Channel Cross-Sectional Adjustment .............................. 11 2.2.2.2 Controls on the Braiding Threshold ................................. 16 2.2.2.3 Controls on Channel Sinuosity ........................................ 19

2.3 Downstream Variation in Channel and Floodplain Character ................. 20 2.3.1 Downstream Variation in Channel Morphology .......................... 20 2.3.2 Downstream Variation in Bed Material ....................................... 22 2.3.3 Downstream Variation in Floodplain Character .......................... 23 2.3.4 The Case for Non-Alpine Deglaciated Watersheds ......................24

2.4 Research Objectives ................................................................................. 26 Chapter Three: Little Rouge River Watershed .................................................. 27

3.1 Introduction .............................................................................................. 27 3.2 Location ................................................................................................... 28 3.3 Land-use ................................................................................................... 28 3.4 Climate and Hydrology ............................................................................ 28 3.5 Physiography ............................................................................................ 37

3.5.1 Physiographic Regions ................................................................. 37 3.5.2 Watershed Stratigraphy ................................................................ 40

3.6 Study Sites ............................................................................................... 45 Chapter Four: Methods ....................................................................................... 51

4.1 Fieldwork ................................................................................................. 51 4.1.1 Surveying and Bed-Augering ...................................................... 51 4.1.2 Grain Size Analysis ...................................................................... 51 4.1.3 Floodplain and Channel Bank Sampling ..................................... 53

4.2 Lab Analysis of Floodplain Sediments .................................................... 54 4.3 Determination of Morphological Bank-full Discharge ............................ 54

Page 6: Downstream Variability of Fluvial Form, Process, and Character in a

vi

Chapter Five: Watershed-Scale Variability of Fluvial Form and Process .... 57

5.1 Introduction .............................................................................................. 57 5.2 Low Energy Mixed-Load Type ............................................................... 60 5.3 Low Energy Bed-Load Type ................................................................... 66 5.4 Mixed Energy Type ................................................................................. 69 5.5 Medium Energy Type .............................................................................. 74 5.6 High Energy Type .................................................................................... 81 5.7 Equilibrium Morphology Prediction ........................................................ 86

Chapter Six: Downstream Organization of Fluvial Form, Process, and Character .................................................................................... 94

6.1 Spatial Distribution of Channel-Floodplain Types .................................. 94 6.2 Downstream Organization of Watershed Zones ...................................... 95 6.3 Downstream Trends in Channel and Floodplain Character ................... 100

6.3.1 Channel Morphology ................................................................. 101 6.3.2 Channel Hydrology .................................................................... 102 6.3.3 Channel Bed Material Size ........................................................ 102 6.3.4 Overbank Alluvium Thickness .................................................. 105 6.3.5 Summary and Controls .............................................................. 107

Chapter Seven: Discussion and Conclusions .................................................... 111

7.1 Reappraisal of the ‘Graded’ River Model for Non-Alpine Deglaciated Watersheds ............................................................................................. 111

7.2 Summary of Findings ............................................................................. 112 7.3 Future Research Considerations ............................................................ 113

References ..................................................................................................................... 115 Appendices .................................................................................................................... 129 Curriculum Vitae ......................................................................................................... 204

Page 7: Downstream Variability of Fluvial Form, Process, and Character in a

vii

List of Figures Figure Page 2.1 Conceptual diagram illustrating channel equilibrium, disequilibrium, and

nonequilibrium as function of time ........................................................................ 5 2.2 Schematic representation of the stages of a geomorphic system’s response to a

disturbance ............................................................................................................. 6 2.3 Classification of gravel-bed channels based on bank erodibility and bed material

supply rate from upstream relative to the transport capacity of the channel ....... 18 2.4 Schematic diagram of the typical ‘graded’ fluvial system and downstream trends

in major system characteristics partitioned by zones ........................................... 21 2.5 Theoretical qualitative characteristics of the production, transport, and deposition

zones .................................................................................................................... 25 3.1 Variation in channel slope along the Little Rouge River ..................................... 27 3.2 Location of the Little Rouge watershed within southern Ontario ........................ 29 3.3 Land-use classification for the Rouge River watershed ...................................... 30 3.4 Mean monthly discharge values for the gauging stations on the Little Rouge River

............................................................................................................................... 31 3.5 Locations of active hydrometric gauging stations operated by the Water Survey of

Canada and the Environment Canada weather station within the watershed ...... 32 3.6 Flood frequency curves for the Locust Hill gauging station ................................ 34 3.7 A comparison of annual patterns in mean discharge between the Little Rouge

watershed and two neighbouring urbanized basins over 40 years ....................... 35 3.8 Decadal variation in average annual mean discharge in the Little Rouge watershed

............................................................................................................................... 36 3.9 Seasonal proportions in annual discharge between the Little Rouge, Rouge, and

Highland watersheds for two time periods .......................................................... 36 3.10 Physiographic regions of the Rouge River watershed ......................................... 38

Page 8: Downstream Variability of Fluvial Form, Process, and Character in a

viii

3.11 Surficial geology deposits within the Rouge River watershed ............................ 39 3.12 Generalized stratigraphic diagram for the Rouge River watershed ..................... 41 3.13 Whitby Formation shale exposed in an outcrop of the Little Rouge River

approximately 1 km upstream of the Rouge River confluence ............................ 42 3.14 Sunnybrook Drift exposed in a terrace downstream of Meadowvale Road ........ 42 3.15 Halton Till exposed in a low-lying terrace downstream of 14th Avenue ............. 43 3.16 Geologic cross-section through the Rouge River watershed ............................... 45 3.17 Location of study sites within the watershed ....................................................... 46 3.18 Location of study sites in the context of the longitudinal profile ........................ 47 3.19a Imagery of sites 1-3 ............................................................................................. 49 3.19b Imagery of sites 4-6 ............................................................................................. 50 5.1 Field image and idealized schematic cross-section of the low energy, mixed-load

channel-floodplain type ....................................................................................... 61 5.2 Plan-form discrimination of channel-floodplain types based off of potential

specific stream power and median bed material size ........................................... 62 5.3 Characteristic floodplain stratigraphy of the low energy, mixed-load type ......... 64 5.4 Field image and idealized schematic cross-section of the low energy, bed-load

channel-floodplain type ....................................................................................... 66 5.5 Characteristic floodplain stratigraphy of the low energy, bed-load type ............. 68 5.6 Field image and idealized schematic cross-section of the mixed energy channel-

floodplain type ..................................................................................................... 70 5.7 Characteristic floodplain stratigraphy of the mixed energy type ......................... 73 5.8 Development of an avulsion channel around an active debris dam ..................... 74 5.9 Field image and idealized schematic cross-section of the medium energy channel-

floodplain type (Site 4d) ...................................................................................... 76

Page 9: Downstream Variability of Fluvial Form, Process, and Character in a

ix

5.10 Field image and idealized schematic cross-section of the medium energy channel-floodplain type (Site 6) ........................................................................................ 77

5.11 Characteristic floodplain stratigraphy of the medium energy type (Site 4d) ....... 78 5.12 Characteristic floodplain stratigraphy of the medium energy type (Site 6) ......... 79 5.13 Extended floodplain stratigraphy of Site 6 .......................................................... 82 5.14 Field image and idealized schematic cross-section of the high energy channel-

floodplain type ..................................................................................................... 83 5.15 Characteristic floodplain stratigraphy of the high energy type ............................ 85 5.16 Comparison of field-derived channel widths and depths against predicted regime

dimensions ........................................................................................................... 89 5.17 Relationship between relative bank strength and dimensionless shear stress ...... 91 6.1 Downstream variation in specific stream power in relation to glacial landforms 96 6.2 Dominant glacial and post-glacial controls on the longitudinal profile ............... 97 6.3 Spatial distribution of inferred watershed zones for each study site ................... 99 6.4 Downstream variation in bank-full width, depth, and cross-sectional area ....... 101 6.5 Downstream variation in bank-full discharge .................................................... 103 6.6 Downstream variation in bank-full velocity ...................................................... 103 6.7 Downstream variation in grain size sorting coefficient and various size fractions .. ..............................................................................................................................104 6.8 Downstream variation in floodplain overbank alluvium thickness ................... 107

Page 10: Downstream Variability of Fluvial Form, Process, and Character in a

x

List of Tables Table Page 2.1 A compilation of exponents predicting channel dimensions from various studies

on gravel-bed rivers ............................................................................................. 13 2.2 A compilation of exponents from various studies predicting how the threshold

channel slope for braiding responds to various controls ...................................... 17 2.3 Classification of genetic floodplain types based on specific stream power ......... 23 3.1 Description of the active hydrometric gauging stations within the Little Rouge

watershed ............................................................................................................. 31 3.2 Return periods for the mean annual flood discharges at the Locust Hill gauging

station ................................................................................................................... 33 3.3 Study site characterization prior to fieldwork ...................................................... 48 5.1 Bank-full channel and floodplain characteristics of the study sites ..................... 58 5.2 Description of the channel-floodplain types identified within the Little Rouge

River watershed ................................................................................................... 59 5.3 Facies classification of the study sites ................................................................. 60 5.4 Radiocarbon dates from the study sites ............................................................... 65 5.5 Relative bank strength values and associated vegetation conditions of the study

sites ...................................................................................................................... 90 6.1 Observed downstream trends in various channel and floodplain characteristics

compared against their expected trends ............................................................. 108

Page 11: Downstream Variability of Fluvial Form, Process, and Character in a

xi

List of Appendices

Appendix Page A Channel Morphology and Sedimentology ............................................. 129 B Floodplain Sedimentology ..................................................................... 135 C Bank-full Discharge Formulae Analysis ................................................ 139 D Flood Frequency Analysis ..................................................................... 143 E Downstream Trends in Channel and Floodplain Characteristics ........... 156 F Glacial Control on Downstream Trends ................................................ 162 G Statistical Tests ...................................................................................... 200

Page 12: Downstream Variability of Fluvial Form, Process, and Character in a

1

Chapter One: Introduction

1.1 Definition of the Problem

It is well understood that rivers adjust their equilibrium morphology to a given

discharge, slope, sediment load, and sediment size (Leopold and Bull, 1979), and the

downstream variation of these control variables dictates the downstream progression of

channel morphology, plan-form, and floodplain characteristics. In conventional alluvial

systems, discharge is expected to increase downstream while channel slope and bed

material size decline (Knighton, 1987). How sediment supply changes downstream is less

certain. These trends often produce a predictable downstream succession of fluvial forms

and characteristics related to high-energy (ie. specific stream power) conditions in the

headwaters and low-energy conditions at the river outlet.

In watersheds composed of mixed alluvial and non-alluvial boundaries, like

southern Ontario, however, such downstream trends do not necessarily apply, as channel

slope, sediment supply, and bed material size are highly variable. The pervasive effects of

the last glaciation in Canada and southern Ontario are obvious. Longitudinal profiles in

these contexts are often composed of both convexities and concavities as a result of

topographic variations imposed by remnant glacial landforms, bedrock outcroppings, and

fluctuations in base-level. Campo and Desloges (1994) studied the sediment budget of a

medium sized southern Ontario watershed and found that up to 68% of the incoming

sediment is derived from fine-grained glacial sources that are transported through the

system as wash load, while gravel inputs, which dominate much of the bed material in

southern Ontario streams, are comparatively small. Furthermore, contrary to typical

models for alluvial rivers (ie. Jiongxin and Yunxia, 2005, their Figure 1), it was found

Page 13: Downstream Variability of Fluvial Form, Process, and Character in a

2

that due to progressive downstream incision into glacial deposits and an increase in

channel-hillslope coupling, specific sediment yield tends to increase with drainage area

(Campo and Desloges, 1994). Finally, bed material does not conform to the expected

downstream fining described in the literature (Knighton, 1987), but instead, coarse gravel

and cobble deposits derived from nearby glacial sources typically dominate the bed

material along the entire channel length with variable amounts of glacial clay exposed in

the channel perimeter. Such morphology has typically been referred to as semi-alluvial

for its superficial resemblance to alluvial channels without being fully alluvial in nature

(Foster, 1998).

Based on the observed departure in controlling variables from traditional models,

it is not clear whether southern Ontario streams exhibit similar downstream trends in

fluvial form, and channel and floodplain characteristics, as seen in their conventional

counterparts. Nor is it clear how glacial inheritance influences downstream channel

adjustment. Moreover, Ferguson (1981) suggests that many semi-alluvial channels may

be in a state of persistent passive disequilibrium due to the imbalance between sediment

supply, sediment calibre, and transport capacity, such that these streams are incapable of

modifying their channel morphology and slope to the contemporary hydrologic and

sediment regimes. However, such a hypothesis has yet to be fully tested.

1.2 Research Questions

The research presented herein aims to address three primary questions:

1. What are the downstream trends in fluvial form, processes, and characteristics in

watersheds that have been conditioned by past glaciation?

2. To what extent does glacial inheritance influence these trends?

Page 14: Downstream Variability of Fluvial Form, Process, and Character in a

3

3. To what extent are these rivers in equilibrium with respect to the prevailing

hydrological and sediment regimes?

Page 15: Downstream Variability of Fluvial Form, Process, and Character in a

4

Chapter Two: Equilibrium and Downstream River Adjustment 2.1 The Concept of Equilibrium in River Systems

While the concept of channel equilibrium was studied prior to the twentieth

century, significant focus on the topic was not undertaken until the mid-twentieth century

with the influential papers of Mackin (1948) and Langbein and Leopold (1964).

However, despite this and subsequent research, there remains discrepancies on what

equilibrium is in the geomorphic context as different workers adopt different definitions

(Thorn and Welford, 1994). Generally though, the consensus is that river channels are

adjusted to a ‘dynamic equilibrium’ or ‘quasi-equilibrium’ state whereby the channel

characteristics fluctuate around a static or dynamic mean value (Langbein and Leopold,

1964; Renwick, 1992; Thorn and Welford, 1994).

Renwick (1992) provides a very useful review of equilibrium, disequilibrium, and

nonequilibrium fluvial landforms in geomorphology with particular emphasis on river

systems. He defines disequilibrium as a state trending towards equilibrium but has not

had sufficient enough time to reach it, while a nonequilibrium state is characterized by a

state not migrating towards any identifiable mean form over any length of time (Fig. 2.1).

Ferguson (1981) introduced a subset of disequilibrium, which he termed persistent

passive disequilibrium. Here, the channel appears to be stable under the present hydro-

climate but would not have formed under it. This is to imply that the present channel has

inherited many of its characteristics (ie. slope and grain size distribution) from a previous

channel formed under different hydro-climatic conditions, but is unable to adjust these

characteristics. In essence, a channel under these conditions may be considered ‘frozen’.

However, persistent passive disequilibrium may be a misnomer, as the channel may not

Page 16: Downstream Variability of Fluvial Form, Process, and Character in a

5

be trending towards an equilibrium state since it is unable to adjust itself fully under

present conditions. This idea put forth by Ferguson (1981) may be more akin to a

threshold-dominated nonequilibrium state (Fig. 2.1c), whereby the channel adjusts during

high magnitude but low frequency discharge events.

Figure 2.1 - Conceptual diagram illustrating equilibrium (a), disequilibrium (b), and various states of nonequilibrium (c-e) as functions of time (after Renwick, 1992).

2.1.1 Scales of Equilibrium

In his review, Renwick (1992) emphasizes the importance of both the spatial and

temporal scales to equilibrium landforms. Implicit in the definition of dynamic

equilibrium is the time-scale at which the ‘mean condition’ is defined. Hickin (1983)

supplies three time-scales applicable to geomorphic processes: the geological time-scale

defined as a time period greater than 106 years, the geomorphic time-scale, which falls

between 102 to 106 years, and the engineering time-scale, which is defined as anything

less than 102 years. Whether or not a geomorphic system reaches equilibrium is

Page 17: Downstream Variability of Fluvial Form, Process, and Character in a

6

dependent on the time-scale of the response and the time passed since the initial

disturbance to the system. The response time is the combination of the reaction time (the

period of time where the system absorbs the disturbance) and the relaxation time (where

the system adjusts to the disturbance). If the response time is greater than the time passed

since the initial disturbance, then equilibrium may not be fully attained and the landform

can be considered to be in a state of disequilibrium (Fig. 2.2) (Renwick, 1992). This

relationship holds particular significance in light of global environmental change that has

occurred during and since the last glaciation to which many river systems may still be

adjusting.

Figure 2.2 - Schematic representation of the stages of a geomorphic system's response to a disturbance. The reaction time occurs where the system absorbs the disturbance and changes are imperceptible. The relaxation time is defined by visible changes to the system as it adapts to the disturbance and reaches a new equilibrium. Equilibrium and disequilibrium can be defined in the context of the diagram where disequilibrium occurs before the relaxation period is complete (after Charlton, 2008).

Page 18: Downstream Variability of Fluvial Form, Process, and Character in a

7

Moreover, the time-scale of relaxation is dependent on the size of the landform,

such that larger landforms may have longer relaxation periods then smaller landforms.

This is illustrated through the study by Church et al. (1989), who show that for

watersheds in British Columbia between 10 km2 and 30 000 km2, specific sediment yield

tends to increase with drainage area – a trend does not follow conventional alluvial

models – and has been attributed to the downstream remobilization of glacial valley fill

sediments. It has been concluded that the sediment yield within these basins is still out of

phase with the present-day climate and that the time-scale necessary to remove the glacial

sediments from the river system is many tens of thousands of years (Church and

Slaymaker, 1989). This pattern in specific sediment yield is not restricted to alpine

environments either, as Ashmore (1993), Campo and Desloges (1994), and Church et al.

(1999) have all described similar increasing or stable trends in non-alpine regions of

Canada. Conversely, reach-scale morphology, in some cases, may have adjusted to

present-day conditions as Dury et al. (1972) show that the bed-morphology of the River

Severn is scaled to the current channel discharge while Phillips and Robert (2007)

demonstrate that wave-lengths on a small meandering stream in southern Ontario are

strongly controlled by the present-day discharge regime.

2.2 Adjustment of the Equilibrium Channel Form

The Lane Balance (Charlton, 2008) states that river channels adjust themselves to

a balance between discharge (Q), median grain size of the bed material (D50), channel

slope (SC), and sediment supply (Qs) such that

C 50 s.Q·S D ·Q 2.1

Page 19: Downstream Variability of Fluvial Form, Process, and Character in a

8

In theory, should one of these variables change then the others would need to

adjust accordingly to maintain balance. Mackin (1948) introduced the idea that an

equilibrium channel is one that is ‘graded’, whereby the slope is adjusted in order to

convey the imposed sediment load with the minimum velocity for a given discharge. Any

changes in the above-mentioned variables will elicit a change in channel slope. While

Mackin’s (1948) model of channel adjustment is attractive, it has since been shown to be

only partially correct as Leopold and Bull (1979) re-state a graded channel as one where

the channel dimensions (width and depth), slope, grain size, bed roughness, and plan-

form mutually adjust to each other. In responding to changes in the driving variables,

river channels have at least eight variables to which channel morphology can adjust

(Hickin, 1983), which underlines the complexity in understanding and predicting channel

response, as one must account for all of these potential directions of change – a task

which has proven quite difficult.

2.2.1 Predictive Methods of Channel Adjustment in Gravel-Bed Rivers

While Mackin’s (1948) graded river concept is purely qualitative in nature,

subsequent researchers have aimed to quantify channel adjustment through empirical

relationships based on field and flume data, and theoretical models based on physical

relations. Empirical relationships were pioneered by Leopold and Maddock (1953) in

their classical study on hydraulic geometry. Channel dimensions were individually

plotted against bank-full discharge and a power trend-line was fitted to the data.

Hydraulic geometry relationships have since evolved to incorporate other variables such

as grain size and bank characteristics. A primary criticism aimed at hydraulic geometry is

that it is only a statistical relation describing what is there and does not have much

Page 20: Downstream Variability of Fluvial Form, Process, and Character in a

9

explanatory power in regards to the underlying fluvial processes (Hickin, 1983).

However, Parker et al. (2007) show that the exponents and coefficients of hydraulic

geometry relations can be physically explained. Empirical relationships have also been

useful in the establishment of plan-form thresholds by plotting channels of different

patterns on a plot based on two governing characteristics (typically discharge or

dimensionless discharge and channel slope) and manually fitting a line that separates the

plan-forms. Again, while this method gives inferential evidence on why different plan-

forms exist, little physical rationale can be extended. Moreover, empirical relationships

suffer from the fact that they can only be reliably used within the region that they were

developed in and for the range of data that they were derived from.

Theoretical, ‘rational’, relationships, on the other hand, use physical relations to

establish equilibrium morphology. These relationships come in two forms; the first of

which determines the channel width by the simultaneous solution of the flow continuity

relation, a flow resistance relation, and dimensionless shear stress (Griffiths, 1981).

These relations are designed to be used on threshold channels, which are defined as

gravel-bed channels that marginally exceed the threshold for sediment transport at bank-

full flow and which transport minor amounts of bed-load (Ferguson, 1986).

The second type of rational relation uses physical relations combined with an

extremal hypothesis constraint, which purport to explain the underlying behaviour of

river channels, to model the morphology. Examples of these hypotheses include

Minimum Stream Power, Minimum Unit Stream Power, Maximum Friction Factor, and

Maximum Sediment Transport Capacity (Davies and Sutherland, 1983; Bettess and

White, 1987). Despite their differing names, these hypotheses tend to suggest that rivers

Page 21: Downstream Variability of Fluvial Form, Process, and Character in a

10

minimize their channel slopes to convey the imposed sediment load. Extremal

hypotheses, however, have been tepidly received, as they are not regarded as physically

sound. Griffiths (1984) notes that extremal hypotheses predict a constant Shield’s value,

which is a trend observed in neither natural channels nor flume studies. Early models

neglected to include some measure of bank resistance, which often resulted in inaccurate

channel dimension predictions (Ferguson, 1986). Eaton and Millar (2004) addressed this

by applying a bank stability constraint to a model employing the Maximum Transport

Efficiency extremal hypothesis and found that when this constraint was applied, the

agreement with field derived data was considerably higher than that predicted by the

unconstrained model. The success experienced by Eaton and Millar (2004) has been

echoed in other studies where their model was employed (Eaton et al., 2004; Eaton and

Church, 2007). However, despite being shown as a promising avenue of research going

forward, extremal hypotheses continue to be relatively inaccessible due to their

computational demands.

2.2.2 Controls on Channel Morphology: Direction and Magnitude

As mentioned, there are four dominant controls on channel morphology

represented in the Lane balance with bank strength being subsequently added (Millar,

2005). These variables, however, disproportionately control channel morphology in both

the negative and positive directions. Thus, a wide variety of morphologies may exist

depending on the combination of the control variables. Generally speaking though, the

morphology of the channel is dictated by the relationship between the erosive force of the

flow and the resisting forces of the bed and channel banks.

Page 22: Downstream Variability of Fluvial Form, Process, and Character in a

11

Comprehensive reviews on early accounts of the controls on channel morphology

and plan-form have been given elsewhere (Carson, 1984b; Ferguson, 1987; Knighton,

1987). Thus, the intention of the following discussion is to provide a general review of

the present knowledge on the direction and relative magnitude of the dominant

morphologic controls with reference to various predictive relations for predominantly

gravel-bed rivers. Furthermore, while several types of channel patterns have been

identified, the transition between meandering and braided channels will be emphasized

here, as they represent the two most common plan-forms.

2.2.2.1 Channel Cross-Sectional Adjustment

The earliest hydraulic geometry relations related changes in channel dimensions

to the downstream change in discharge. Leopold and Maddock (1953) originally

proposed that for a specific set of alluvial rivers, downstream adjustments in channel

form take on the relations of

0.50w = aQ 2.2

0.40d = cQ 2.3

for the width (m), and depth (m), respectively where Q is a representative discharge

value, typically bank-full, in m3s-1, and a and c are constants. Their analysis suggests that

much of the adjustment to changes in discharge is taken up almost evenly by width and

depth. Subsequent research has revealed a high degree of consistency of the exponents in

Equations 2.2 and 2.3 from other physiographic regions (Table 2.1; Knighton, 1987).

However, Park (1977) has demonstrated that despite the tendency to adopt these

exponents, there exists some scatter about these values. The width exponent was found to

vary between 0.03 and 0.89, and depth between 0.09 and 0.70. These discrepancies are

Page 23: Downstream Variability of Fluvial Form, Process, and Character in a

12

thought to be a result of climatic differences between the study regions in addition to, and

perhaps secondarily to, other local controls on channel morphology.

Grain size has since been acknowledged to exert control over the channel

morphology, though the direction of change and relative magnitude are not fully

resolved. Parker (1979), Andrews (1984), and Parker et al. (2007), for example,

developed dimensionless hydraulic geometry relations, which, when dimensionalized,

show that channel width inversely varies with the D50 between the powers of -0.17 and

-0.25 while depth is less sensitive and directly varies between 0 and 0.08 (Table 2.1).

These values are confirmed by Millar (2005) when sediment concentration is used. While

Parker (1979) and Griffiths (1981) found much larger exponents in their relations, the

same inverse relation with grain size is observed. The implication here is that as grain

size becomes larger for gravel-bed rivers, channels become narrower and slightly deeper,

which is in direct contrast to what Davidson and Hey (2011) found in their study where,

after removing sediment load from their relations, the width and depth vary with D50 to

the powers of 0.17 and –0.16 respectively, such that the channel becomes wider and

shallower. The inclusion of sediment load in their relations reduces the D50 exponents to

–0.05 for width and –0.20 for depth. Hey and Thorne (1986) observed similar values for

depth. While the observations of Davidson and Hey (2011) are typical of what is

observed for gravel-bed rivers, the inverse trend reported by Parker (1979), Griffiths

(1981), Andrews (1984), Millar (2005), and Parker et al., (2007) does have a physical

rationale. Since the ability to transport sediment of a given size depends on the channel

slope and depth of the channel, an increase in grain size necessitates an increase in one of

these two variables. However, for a fixed slope and discharge, the channel can only

Page 24: Downstream Variability of Fluvial Form, Process, and Character in a

Regime Equation Variables Method Study Dimension Qbf D50 Qs SC Bank Strength

Empirical Leopold & Maddock (1953) Width Depth

0.50 0.40

- -

- -

- -

- -

Ferguson (1971)† Width 0.58* - - - -0.66 Parker (1979)‡ Width 0.50 -0.25 - - - Andrews (1984)‡ Width

Depth

0.48 0.48 0.38 0.37

-0.20 -0.20 0.06 0.08

- - - -

- - - -

2.54 (Thin Veg) 2.00 (Thick Veg) 0.288 (Thin Veg) 0.29 (Thick Veg)

Hey & Thorne (1986) Width

Depth

0.52

0.39

-

-0.15

-0.01

-0.02

- -

3.98 (Type I) 3.08 (Type II) 2.52 (Type III) 2.17 (Type IV) 0.16 (Type I)

0.19 (Types II/III)0.20 (Type IV)

Parker et al. (2007)‡ Width Depth

0.47 0.40

-0.17 -0.00

- -

- -

- -

Davidson & Hey (2011) Width

Depth

0.52

0.38

-0.05

-0.20

-0.05

-0.01

-0.26***

0.08***

1.48 (Type I) 1.13 (Type II) 0.82 (Type III) 0.53 (Type IV) 0.27 (Type I) 0.28 (Type II) 0.32 (Type III) 0.39 (Type IV)

Semi-Theoretical Parker (1979) Depth 0.42 -0.04 - - - Theoretical Parker (1979) Width 1.00 -1.50 - 1.22 -

Griffiths (1981) Width 1.00 -1.50 - 1.26 - Millar (2005)‡ Width

Depth Width Depth

0.50 0.37 0.70 0.16

-0.25 0.08 -0.75 0.60

1.12** -1.16**

- -

- -

0.60 -0.62

-1.66 1.22 -1.10 0.64

Page 25: Downstream Variability of Fluvial Form, Process, and Character in a

14

Table 2.1 (previous page) - A compilation of exponents for predicting channel dimensions from various studies on gravel-bed rivers where Qbf is the bank-full discharge, D50 is the median bed material size, Qs is the sediment discharge, and SC is the channel slope. † - data from sand-bed rivers; ‡ - data from dimensionless discharge analysis; * - discharge as the mean annual flood; ** - exponents presented as negative in the original paper for dimensionless sediment concentration defined as the –log(sediment concentration); *** - valley slope. Values italicized represent coefficients instead of exponents.

change its dimensions and thus, the channel must narrow in order to accommodate the

requisite increase in depth (Chew and Ashmore, 2001). In terms of magnitude though,

grain size seems relatively ineffective at forcing much channel change, especially in

regard to channel depth, and is over-shadowed by the control of discharge on channel

morphology. Moreover, the discrepancy between Hey and Thorne (1986), Davidson and

Hey (2011), and the other studies in Table 2.1 perhaps alludes to other, stronger controls

on channel morphology.

Bed-sediment load is not often included in regime relations due to its difficulty in

accurate quantification in both space and time. As a substitute, channel slope is often

employed, as it has been shown to scale with bed-sediment load (Schumm and Khan,

1972; Eaton and Church, 2004). However, slope has only been included in theoretical

relations and just recently in empirical relations. In terms of bed-sediment load, there

exists a marked disparity in its importance to the channel dimensions. Both Hey and

Thorne (1986) and Davidson and Hey (2011) found that sediment load has no effect on

width or depth, a fact attributed to co-linearity with discharge. Millar (2005), however,

shows that sediment load has a profound morphological impact, where an increasingly

wide and shallow channel exists with an increasing sediment load. If we believe that

channel slope is an apt surrogate for sediment load, then the regime relations of Parker

(1979) and Griffiths (1981) add further credence to sediment load’s positive effect on

channel width. Moreover, it is suggested that the positive relationship between channel

slope and width is due to the greater erosive power of the channel (Chew and Ashmore,

Page 26: Downstream Variability of Fluvial Form, Process, and Character in a

15

2001). The regime relations of Davidson and Hey (2011), on the other hand, suggest that

as valley slope increases, depth is unaffected and channels become narrower, which may

be attributed to higher flow velocities that necessitate a decrease in width, in light of

constant depth, to maintain continuity.

It is also important to note the morphological implications on the relative amount

of suspended-load to bed-load. Schumm and Khan (1972) found that channel width

decreased and depth increased in response to a concomitant increase in suspended

sediment transport and reduction in bed-load sediment transport as a proportion of the

total sediment load.

Finally, bank cohesion, which consists of the combined effects of bank vegetation

density and the silt/clay percentage, has been shown to exert varying degrees of control

over channel morphology. In general though, increases in bank strength tend to result in

narrower and deeper channels as shown in Table 2.1 due to the cohesive nature of the

sediment buffered by the roots of trees, which resist the erosive forces of the flow. Millar

(2005) has demonstrated in his theoretical relations that this variable may possess the

greatest amount of control over the channel dimensions. However, since Andrews (1984),

Hey and Thorne (1986), and Davidson and Hey (2011) used coefficients in their analyses,

the relative importance of bank strength cannot be assessed outside of Millar (2005) and

Ferguson’s (1971) studies.

The foregoing review suggests that the two strongest controls on channel cross-

sectional adjustment are the sediment supply and bank strength. However, these findings

are not universal, as Hession et al. (2003) and Allmendinger et al. (2005) have observed

increasing width with an increase in forest cover, opposite of what is suggested here.

Page 27: Downstream Variability of Fluvial Form, Process, and Character in a

16

Moreover, inaccurate representation of the relative magnitude and direction each variable

has on the channel dimensions could occur if important variables were neglected in

empirical formulae or if multi-collinearity exists. This effect is illustrated in the relations

of Millar (2005) where the exponents radically adjust when channel slope is substituted

for sediment supply (Table 2.1).

2.2.2.2 Controls on the Braiding Threshold

Discrimination between meandering and braided plan-forms has historically been

accomplished through the use of a slope threshold. This idea is predicated on the idea that

once a meandering channel crosses a critical slope, it will begin to braid. To this end,

research has been focused on determining what variables influence the critical slope

(Table 2.2).

Originally, discharge was thought to be the dominant controlling variable such

that an increase in discharge decreases the critical channel slope for braiding (Leopold

and Wolman, 1957). Although a range of values has been achieved, channel slope

typically increases to the –0.25 or –0.50 power of discharge. However, median grain size

of the channel bed has been shown to possess 2.5 times the power of discharge over the

critical slope in the opposite direction suggesting that as the channel bed becomes coarser

for a given discharge, the slope, and thus energy necessary for sediment transport, must

increase as well. Bank strength is the final variable included in braiding thresholds and

appears to exert the most control over the threshold being 2 times more impactful than

grain size and almost 5 times more so than discharge (Table 2.2).

Not explicitly stated in Table 2.2 but nevertheless paramount to the initiation of

braiding is the sediment supply of bed-calibre sediment to the river channel. This is

Page 28: Downstream Variability of Fluvial Form, Process, and Character in a

Braiding Threshold Variables Method Study Channel

Slope Qbf D50 Bank Strength

Empirical Leopold & Wolman (1957) -0.44 - - Kleinhans & van den Berg (2011) -0.50 0.42 -

Semi-Theoretical Henderson (1963) -0.46 1.15 - Carson (1984b) -0.33 1.00 - Ferguson (1987) -0.21 0.52 - Ferguson (1987)† -0.24* - 1.00

Theoretical Millar (2000) -0.25 0.61 1.75 Millar (2005)‡ -0.25 0.63 1.00 Eaton et al. (2010)‡

SC

-0.43 1.08 1.41

Table 2.2 - A compilation of exponents from various studies predicting how the threshold channel slope for braiding responds to various controls where Qbf is the bank-full discharge and D50 is the median bed material size. † - data from sand-bed rivers; ‡ - data from dimensionless discharge analysis; * - discharge as the mean annual flood.

Page 29: Downstream Variability of Fluvial Form, Process, and Character in a

18

implicit in the strong control exhibited over the threshold slope by bank strength. Carson

(1984a) provides a qualitative diagram relating the importance of bank stability and

sediment supply from upstream to channel plan-form in gravel-bed rivers (Fig. 2.3).

Braided rivers occur where highly erodible banks deliver bed-calibre sediment to the

channel whereas meandering rivers exist due to low bank erodibility and sediment

supply. Furthermore, the link between sediment supply and braiding is emphasized in the

study by Smith and Smith (1984) where braiding was induced after a 400% increase in

bed-load sediment along the William River.

Figure 2.3 - Classification of gravel-bed channels based on bank erodibility and bed material supply rate from upstream relative to the transport capacity of the channel (after Carson, 1984a).

Page 30: Downstream Variability of Fluvial Form, Process, and Character in a

19

2.2.2.3 Controls on Channel Sinuosity

Much like the preceding discussion on channel dimensions and the braiding

threshold, the controls on channel sinuosity are seemingly diverse. An oft-cited control is

that of Schumm’s (1963) account that sinuosity is positively related to the weighted

percent silt/clay in the channel perimeter. However, this conclusion is marred by the fact

that his weighting is based on the width-depth ratio such that the coefficient of

determination between sinuosity and the width-depth ratio is 0.79 while the correlation

with the weighted silt/clay percentage is 0.83. Clearly, the addition of the silt/clay content

only marginally improves the correlation. Nevertheless, Schumm (1963) inferred that

higher sinuosity is a consequence of channels transporting a greater proportion of

suspended/wash load sediment versus bed-load sediment. This inference has been

reinforced by the flume studies of Schumm and Khan (1972) where, when kaolinite was

introduced into the sediment load with a concomitant decrease in bed-load supply,

cohesive sediment deposition began to stabilize submerged bars and the channel sinuosity

consequently increased. Furthermore, the width-depth ratio of the channel decreased,

which accounts for the strong correlation between sinuosity and the width-depth ratio

found by Schumm (1963) earlier. Eaton and Church (2004; 2009) parallel Schumm and

Khan (1972) in showing that for a fixed valley slope and discharge, sinuosity of the

thalweg can be increased by decreasing the bed-load supply, however, bank-cohesion

may retard this.

Lastly, valley slope has often been thought to be a dominant control on sinuosity

(Schumm, 1993). The rationale here is that as valley slope increases, the channel has

more potential energy for erosion, and in order to reduce this energy, the channel erodes

Page 31: Downstream Variability of Fluvial Form, Process, and Character in a

20

the outer banks of meander bends, thereby decreasing the channel slope and increasing

sinuosity. This idea was, in part, borne out of flume studies where the thalweg sinuosity

increases with increasing valley gradient before the channel begins to braid at a critical

slope (Schumm and Khan, 1972). However, empirical analysis demonstrates no

correlation between sinuosity and valley slope (Kleinhans and van den Berg, 2011) and

the most convincing explanation for sinuosity remains the proportion of suspended/wash

load to bed-load sediment.

2.3 Downstream Variation in Channel and Floodplain Character

The downstream variation of channel and floodplain characteristics is dictated by

the downstream variation in grain size, sediment supply, valley slope, and discharge as

well as bank vegetation. In conventional ‘graded’ fluvial systems, the watershed can be

divided into three contiguous zones related to the production, downstream movement,

and deposition of sediment (Fig. 2.4). These zones are very much a qualitative abstraction

of watersheds with no quantitative definition or distinct boundaries. However, each zone

is typically characterized by specific channel and floodplain morphologies reflective of

the control variables within the zone.

2.3.1 Downstream Variation in Channel Morphology

The production zone is often characterized by steep valley slopes and high

sediment supply from extensive hillslope-channel coupling (Fig. 2.4). Where sediment

supply is particularly high, such as in recently deglaciated regions, braided patterns can

typically develop, while in non-glaciated mountainous and piedmont regions, step-pool

and cascade morphologies typically dominate with debris and log jams being of major

geomorphic significance (Montgomery and Buffington, 1997). Local sections of bedrock

Page 32: Downstream Variability of Fluvial Form, Process, and Character in a

21

Figure 2.4 - Schematic diagram of the typical ‘graded’ fluvial system and downstream trends in major system characteristics partitioned by zones. Erosion typically happens in production (Zone 1) and sediment is transferred down-valley (Zone 2) to the depositional zone (Zone 3) (after Charlton, 2008).

may outcrop in the riverbed where channels are supply limited. With an increase in

downstream distance, the degree of hillslope-channel coupling diminishes and with it a

decrease in sediment supply. As a consequence of this, the dominant supply of sediment

is from upstream and bank erosion. Channels in the transportation and depositional zones

Page 33: Downstream Variability of Fluvial Form, Process, and Character in a

22

are characterized by meandering riffle-pool channels. The channel bed in these zones is

entirely alluvial in nature as sediment supply typically is equal to or exceeds transport

capacity. The width-depth ratio of the channel has historically been thought to increase

marginally downstream but due to the complicating factors of other controls, this may not

entirely be the case as is explored further in the next section.

2.3.2 Downstream Variation in Bed Material

The downstream variation in bed material is typically complex due to tributary

inputs (Knighton, 1980), but the general trend is that bed material fines downstream (Fig.

2.4) (Sambrook Smith and Ferguson, 1995). Downstream fining has been shown to be a

result of dominantly size-selective transport, while abrasion is of less importance

(Brierley and Hickin, 1985; Ferguson et al., 1996). Since valley slope often decreases

downstream faster than channel depth can increase, shear stresses tend to decrease

downstream. As a result, the competence of the channel wanes downstream and only

smaller particles can be transported. This mechanism accounts for the observed

downstream increase in sediment sorting of some rivers (Knighton, 1980; Brierley and

Hickin, 1985). Once the median grain size reaches about 10 mm, there exists an abrupt

gravel-sand transition, which reflects a threshold decrease in valley slope where the fall

velocity of the suspended sediment becomes greater than the shear velocity of the flow

(Sambrook Smith and Ferguson, 1995). As a result, much of the suspended load falls out

onto the bed creating a sand-bed channel. The downstream change in bed material size,

particularly at the gravel-sand transition, has implications for channel morphology.

Reflecting on Schumm and Khan’s (1972) experiment with kaolinite, increases in the

percent load transported in suspension results in a sinuous, narrow, and deep channel.

Page 34: Downstream Variability of Fluvial Form, Process, and Character in a

23

Thus, channels in the production zone are expected to be relatively wide, straight, and

shallow due to their predominance of bed-load transport. Conversely, channels occupying

the depositional zone tend to have a greater suspended load component and, as a result,

should be more sinuous, narrower, and deeper.

2.3.3 Downstream Variation in Floodplain Character

Reinfelds et al. (2004) show that the peak of specific stream power corresponds to

the peak in channel gradient. In conventional fluvial systems, this often occurs in the

headwaters with specific stream power decreasing downstream. As such, it is possible to

classify landforms based on specific stream power and arrange them in their expected

downstream trend. Nanson and Croke (1992) devised a classification scheme of

floodplain types based on specific stream power and bed-sediment type that permits this

(Table 2.3).

Order

Type

Specific Stream Power (Wm-2)

Class A: High Energy Non-Cohesive Floodplains A1 Confined coarse-textured floodplains >1000 A2 Confined vertical accretion floodplains 300-1000 A3 Unconfined vertical accretion sandy floodplains 300-600 A4 Cut and fill floodplains ~300

Class B: Medium Energy Non-Cohesive Floodplains B1 Braided river floodplains 50-300 B2 Wandering gravel-bed river floodplains 30-200 B3 Meandering river, lateral migration floodplains 10-60

Class C: Low Energy Cohesive Floodplains C1 Laterally stable, single thread channel floodplains <10 C2 Anastomosing river floodplains <10

Table 2.3 - Classification of genetic floodplain types based on specific stream power (modified after Nanson and Croke, 1992).

Class A floodplains are often formed through catastrophic stripping of the

floodplain by floods, which leave behind a cobble-boulder lag and is followed by

reconstruction of the floodplain through vertical accretion during flood recession. Such

Page 35: Downstream Variability of Fluvial Form, Process, and Character in a

24

floodplains occur in steep headwater reaches with confined floodplains that prevent

floodwaters from extensive lateral flow. Floodplains are typically poorly developed and

may be turned over frequently. Where channels and floodplains are less confined, Class

B type floodplains dominate and occupy the upper part of the production zone through to

the deposition zone. This class of floodplain is the most prevalent in most watersheds and

forms through lateral accretion of gravels and sands with vertically accreted sands, silts,

and clays. Due to wider floodplains, sediment storage is typically larger than upstream,

which is reflected in floodplain thickness and age.

Lastly, Class C type floodplains dominate the deposition zone. These floodplains

form primarily through vertical accretion of silts and clays as lateral migration is

restricted by low erosive energy relative to cohesive riverbanks. Low channel energy

leads to slow turnover rates and high sediment storage on wide floodplains

Reduction of stream energy is also correlated with an increase in floodplain

thickness, such that Class A floodplains have the thinnest floodplains and Class C

floodplains are the thickest (see stored sediment in Fig. 2.4). This was partially shown on

the Squamish River where the downstream succession of river plan-form resulted in a

downstream increase in maximum floodplain thickness of 3.5 m, 4.5 m, and 6 m for a

type B1 braided channel floodplain, B2 wandering channel floodplain, and B3

meandering channel floodplain respectively (Brierley and Hickin, 1991). These

relationships are summarized in Figure 2.5.

2.3.4 The Case for Non-Alpine Deglaciated Watersheds

The preceding discussion presented a stereotypical ‘graded’ model for the

downstream variability in channel and floodplain characteristics, and is predicated on the

Page 36: Downstream Variability of Fluvial Form, Process, and Character in a

25

Figure 2.5 - Theoretical qualitative characteristics of the production, transport, and deposition zones. Arrow directions give downstream direction.

notion that channel slope decreases downstream. In recently deglaciated non-alpine

watersheds however, channel slope is often spatially inconsistent suggesting that fluvial

forms, and channel and floodplain characteristics may be just as spatially variable.

Studies by Brown (1990) and Croke (1994) demonstrate that the erratic downstream

variability in floodplain properties and formative processes in two recently deglaciated

basins are related to the shape of the longitudinal profiles, which were inherited from the

Pleistocene infilling of bedrock basins and bedrock valley confinement. Brewer et al.

(2009) note that the patterns of floodplain sedimentation in Wales are related to glacial

landforms, particularly relict alluvial fans and cross-valley moraines, which act to restrict

channel migration. Upstream of these constrictions the channel is sinuous and inactive

while the channel is more energetic within the valley constrictions. Additionally, the

spatial distribution of glacial sediments is such that sediment inputs are highly variable,

particularly when the degree of hillslope-channel coupling fluctuates downstream. As a

result, lateral inputs from glacial sediments tend to disrupt downstream fining (Foster,

1998). Since channel geometries and plan-forms depend on the interaction between

Page 37: Downstream Variability of Fluvial Form, Process, and Character in a

26

sediment supply, grain size, and channel slope, it seems logical to surmise that river

patterns within deglaciated basins are highly variable in the downstream direction and do

not conform to traditional notions of watershed characteristics as alluded to by previous

studies (Croke, 1994).

2.4 Research Objectives

The specific objectives of the research are as follows:

1. To describe various fluvial forms and the physical processes that shape them in a

deglaciated watershed and to describe downstream trends of various morphologic,

hydrologic, and sedimentologic characteristics of the river channel and floodplain.

2. To relate how glacially inherited slopes and sediment inputs dictate the spatial

distribution of fluvial forms and the downstream progression of the observed

characteristics.

3. To infer whether the channel is in equilibrium based on rational or empirical

relations.

Page 38: Downstream Variability of Fluvial Form, Process, and Character in a

27

Chapter Three: Little Rouge River Watershed

3.1 Introduction

Various southern Ontario watersheds were initially screened based on channel

longitudinal profiles, constructed from topographic maps with a contour interval of 10

metres. Ideal profiles contained at least one or two pronounced convexities and

concavities, such that a complex longitudinal profile was attained. The Little Rouge River

was selected as it meets this criterion, is located fairly close-by, and does not suffer from

heavy urbanization. Figure 3.1 shows the longitudinal profile of the Little Rouge River,

derived from a 10-metre digital elevation model, and the downstream variation in channel

slope, which varies downstream by approximately two orders of magnitude. Specifically,

there exist four peaks in channel slope separated by four areas of low slope. Presumably,

these peaks and valleys contain contrasting fluvial forms and were a primary

consideration for site selection discussed further on.

0.0000

0.0020

0.0040

0.0060

0.0080

0.0100

0.0120

0.0140

0 5 10 15 20 25 30 35 40 45

Distance Downstream (km)

Ch

ann

el S

lop

e (-

)

0

50

100

150

200

250

300

350

Ele

vati

on

(m

ab

ove

sea

leve

l)

Channel Slope

Longitudinal Profile

Figure 3.1 - Variation in channel slope along the Little Rouge River (data provided by R. Phillips).

Page 39: Downstream Variability of Fluvial Form, Process, and Character in a

28

3.2 Location

The Little Rouge River watershed is located in south-central Ontario and

encompasses portions of the cities of Toronto and Pickering, and the towns of

Whitchurch-Stouffville and Markham. The drainage basin is approximately 115 km2 in

area and occupies the eastern third of the Rouge River watershed, which drains south into

Lake Ontario (Fig. 3.2).

3.3 Land-use

While land-use percentages for the Little Rouge River sub-watershed are not

available, they can be inferred from data for the greater Rouge River watershed and

Figure 3.3. Presently, 40% of the Rouge River basin is composed of agriculture and

another 24% is forests, wetlands, and meadows. 35% of the basin is urban land-use with

the remaining 1% as water bodies (TRCA, 2007a). Based on Figures 3.2 and 3.3, it is

apparent that the vast majority of the Little Rouge basin is characterized by non-urban

land-uses and is predominantly agricultural with the downstream portion of the basin

dominated by Rouge Park. The town of Stouffville in the upper northeast corner of the

watershed constitutes the greatest urban land-surface.

3.4 Climate and Hydrology

Based on data from the Environment Canada weather station in Stouffville (ID#

6158084), the average annual temperature between 1971 and 2000 is 6.9ºC. The daily

maximum and minimum are 26.2ºC and –11.6ºC respectively. The average annual

precipitation for the 30-year period is 869 mm. Of this, 85% falls as rain and the

remainder as snow. However, snowfall storage over the winter months means that much

of the runoff is generated by snowfall melt in the months of March and April (Fig. 3.4). A

Page 40: Downstream Variability of Fluvial Form, Process, and Character in a

Figure 3.2 - Location of the Little Rouge watershed within southern Ontario.

Page 41: Downstream Variability of Fluvial Form, Process, and Character in a

30

Figure 3.3 - Land-use classification for the Rouge and Little Rouge (outlined in red) watersheds (modified after TRCA, 2007a).

Page 42: Downstream Variability of Fluvial Form, Process, and Character in a

31

second flood peak is experienced in the late autumn and early winter months, which is

associated with rainstorms and early snowmelt events. Rain-on-snow events typically

produce the highest discharge events and occur in spring. While snowmelt is responsible

for producing the annual flood discharge, the relatively low amount of snowfall

compared to other regions of southern Ontario means that flood peaks are comparatively

low, which is reflected in the analysis by Javelle et al. (2003) who show that watersheds

in south-central Ontario are characterized by low to medium intensity floods that occur

over short time periods.

Station Name Station ID Record Period Record Length (years)

Drainage Area (km2) (% Total Area)

Dicksons Hill 02HC053 2002-2010 9 59.3 (51.7%) Locust Hill 02HC028 1963-2010 48 83.8 (73.2%)

Table 3.1 - Description of the active hydrometric gauging stations within the Little Rouge watershed. Drainage areas are derived from a 10m-resolution digital elevation model.

0.00

0.50

1.00

1.50

2.00

2.50

Jan Feb Mar Apr May Jun Jul Aug Sept Oct Nov Dec

Month

Mea

n M

on

thly

Dis

char

ge

(m3 s-1

)

Locust Hill (1964-2010)

Dicksons Hill (2003-2010)

Figure 3.4 - Mean monthly discharge values for the gauging stations on the Little Rouge River. Periodsof record are indicated beside each station name.

Page 43: Downstream Variability of Fluvial Form, Process, and Character in a

32

Figure 3.5 - Locations of active hydrometric gauging stations operated by the Water Survey of Canada (WSC) and the Environment Canada (EC) weather station within the watershed.

The Water Survey of Canada presently operates two gauging stations on the Little

Rouge River (Fig. 3.5 and Table. 3.1). Since the gauging station near Dicksons Hill is

relatively recent, spatial analyses on the hydrology of the watershed is restricted to the

Page 44: Downstream Variability of Fluvial Form, Process, and Character in a

33

last decade. The mean annual runoff at each station between the years 2003-2010 is 370

mm and 363 mm for the Dicksons Hill and Locust Hill stations respectively, which

suggests a decrease, albeit small (6.2%), in run-off contribution downstream of the

Dicksons Hill station. Figure 3.4 shows that the annual hydrologic regime at both stations

behaves similarly suggesting that major downstream changes in the timing of the annual

flood hydrograph are not significant.

A flood frequency analysis on the Locust Hill station was performed using the

General Extreme Value distribution on the maximum instantaneous and maximum daily

discharges for the annual maximum series (AMS), and the General Pareto distribution on

the maximum daily discharges for the partial duration series (PDS) (Kidson and

Richards, 2005). The cut-off discharge for the PDS was 5.38 m3s-1, which represents the

lowest peak discharge on the AMS (Langbein, 1949). The results (Fig. 3.6) indicate that

the watershed has had a modest history of flooding where the maximum discharge on

record is 36 m3s-1 using maximum instantaneous values and has a recurrence interval of

18.5 years. The mean annual flood is approximately 22 m3s-1 using the maximum

instantaneous values and has a recurrence interval of 2.10 years (Table 3.2).

Max Daily (AMS) Max Instantaneous Max Daily (PDS) Mean Annual Flood (m3s-1) 15.36 21.75 15.36 Return Period (years) 2.16 2.10 2.44

Table 3.2 – Return periods for the mean annual flood discharges at the Locust Hill gauging station. Return period values were derived from the flood frequency curves in Figure 3.5.

While the neighbouring watersheds to the west have experienced increased

urbanization over the last several decades, the Little Rouge watershed has been left

dominantly rural, which is reflected in a comparison of the annual mean discharges of the

Rouge and Little Rouge rivers with Highland Creek since the 1960s (Fig. 3.7). While the

Page 45: Downstream Variability of Fluvial Form, Process, and Character in a

34

0

5

10

15

20

25

30

35

40

0.1 1 10 100

Return Period (years)

Dis

char

ge

(m3s-1

)

Max Daily - PDS (1964-2010)

Max Daily - AMS (1964-2010)

Max Inst. - AMS (1966-2007)

Figure 3.6 - Flood frequency curves for the Locust Hill gauging station. Periods of record are indicated beside each curve label. PDS = partial duration series, AMS = annual maximum series. Little Rouge River does show a modest increase in discharge, the slope of the regression

is not significant at the 90% confidence level, which is in contrast to the other, more

urbanized basins. An analysis of the decadal variation in discharge shows that much of

the observed increase in the Little Rouge basin occurred between the 1960s and 1980s,

remaining relatively constant since (Fig 3.8). In terms of seasonal variation, there has

been no change in the proportion of the annual flow occurring in winter/spring and

summer/fall since the 1960s within the Little Rouge basin (Fig. 3.9). However, this

cannot be said for the Rouge and Highland watersheds, which show decreases in

winter/spring discharge percentages and increases in summer/fall percentages. Further

inspection reveals that much of the summer/fall increases in these two basins occur

Page 46: Downstream Variability of Fluvial Form, Process, and Character in a

35

Figure 3.7 - A comparison of annual patterns in mean discharge between the Little Rouge watershed and two neighbouring urbanized basins over 40 years (after TRCA, 2007b).

Page 47: Downstream Variability of Fluvial Form, Process, and Character in a

36

0.50

0.60

0.70

0.80

0.90

1.00

1960s 1970s 1980s 1990s 2000s

Decade

Ave

rag

e A

nn

ual

Mea

n D

isch

arg

e

(m3s-1

)

Figure 3.8 - Decadal variation in average annual mean discharge in the Little Rouge watershed (data from the Locust Hill station between 1964 and 2009). Error bars represent +/- 1 standard error.

Figure 3.9 - Seasonal proportions in annual discharge between the Little Rouge, Rouge, and Highland watersheds for two time periods. Winter is defined as January-March, Spring as April-June, Summer as July-September, and Fall as October-December (after TRCA, 2007b).

Page 48: Downstream Variability of Fluvial Form, Process, and Character in a

37

almost exclusively in the summer, which is rationalized as an effect of the impervious

nature of urban structures contributing greater run-off during summer convective storms

(TRCA, 2007b). While the summer and fall flow percentages in the Little Rouge basin

are unchanged between the two time periods, the winter discharges appear to have

increased at the expense of the spring discharges (Fig. 3.9). This, however, may be

related to regional climatic changes causing an earlier melt period compared to past

decades.

3.5 Physiography

The highest point within the Little Rouge watershed is 370 meters above sea level

along the Oak Ridges Moraine from where the headwater channel descends nearly 300

meters over a distance of 45 kilometres to the confluence with the Rouge River. Along

the way it passes over an array of glaciogenic sediments within several different

physiographic regions.

3.5.1 Physiographic Regions

The watershed can be divided into four primary physiographic regions that

represent past depositional environments. From north to south, these are the Oak Ridges

Moraine, the South Slope, the Peel Plain, and the Lake Iroquois Sand Plain (Fig. 3.10).

Though not directly within the Little Rouge basin, but still within the greater Rouge

River basin, the Lake Iroquois Clay Plain physiographic region occupies the lowest

portion of the Rouge River basin. Each of the four regions is composed of a variety of

surficial sediments ranging from clay to gravel (Fig. 3.11).

At over 300 metres above sea level, the Oak Ridges Moraine acts as the drainage

divide between channels draining north to Lake Simcoe and south towards Lake Ontario.

Page 49: Downstream Variability of Fluvial Form, Process, and Character in a

38

Figure 3.10 - Physiographic regions of the Rouge River watershed. The Little Rouge watershed is outlined in red (modified after TRCA, 2007a).

Page 50: Downstream Variability of Fluvial Form, Process, and Character in a

39

Figure 3.11 - Surficial geology deposits within the Rouge River watershed. The Little Rouge watershed is outlined in red (modified after TRCA, 2007a).

Page 51: Downstream Variability of Fluvial Form, Process, and Character in a

40

Due to a mixture of relatively permeable sediments, the moraine is an important

source for groundwater recharge, which helps maintain the channel discharge in the

summer (TRCA, 2007a). The South Slope is a relatively flat till plain (Halton Till) that

slopes from the southern flanks of the Oak Ridges Moraine southward towards Lake

Ontario. Overlying portions of the South Slope is the Peel Plain, which is generally more

horizontally angled than the South Slope and consists of glaciolacustrine sands, silts, and

clays deposited in remnant glacial ponds.

Aside from the Oak Ridges Moraine, the Lake Iroquois shoreline is perhaps the

most well known physiographic feature within the watershed and separates the South

Slope to the north from the Sand and Clay plains to the south. This shoreline is composed

of beach gravels and sands deposited by proglacial Lake Iroquois (Fig. 3.11) (Eyles and

Boyce, 1991; TRCA, 2007a).

3.5.2 Watershed Stratigraphy

While the spatial distribution of the surface deposits is complex in nature, the

underlying glacial stratigraphy is similarly complex. Underlying the watershed is a

succession of glacial sediments of varying thickness and granular composition, which

records the Illinoian and Wisconsin glaciations over the last ~100 thousand years (Fig.

3.12). These deposits, however, are not all exposed within the Little Rouge watershed

and only those that are will be discussed here.

The basement bedrock underlying the glacial deposits is the Whitby Formation,

which is a highly fissile bluish-grey shale deposited during the Ordovician Period (Eyles

and Boyce, 1991). The spatial extent of bedrock exposure is quite limited and restricted

Page 52: Downstream Variability of Fluvial Form, Process, and Character in a

41

to brief outcroppings downstream of Twyn Rivers Drive (Fig. 3.13) (Karrow, 1967; Eyles

and Boyce, 1991).

Figure 3.12 - Generalized stratigraphic diagram for the Rouge River watershed (after TRCA, 2007a).

The Scarborough Formation developed during an interstadial period within the

early Wisconsin glaciation and is thought to represent a delta complex prograding into a

proglacial lake (Eyles and Boyce, 1991). The deposits consist of a fining upward

sequence with basal silts and clays capped by fluvial sands and vary in thickness but

typically are several tens of metres thick. Within the Little Rouge watershed, these

deposits outcrop in the bluffs between the confluence with the Rouge River and several

hundred metres upstream of Twyn Rivers Drive (Karrow, 1967). Overlying the

Page 53: Downstream Variability of Fluvial Form, Process, and Character in a

42

Scarborough Formation is the Sunnybrook Drift, which is a clast-poor clayey-silt till

formed during the early Wisconsin stadial with an ascribed thickness of 5 to 10 metres

(Eyles and Boyce, 1991). Sunnybrook Drift can be seen to outcrop in the valley sides,

terraces, and the channel bed between Meadowvale Road and its downstream contact

with the Scarborough Formation, downstream of which it seems to be absent (Fig. 3.14)

(Karrow, 1967).

Figure 3.13 - Whitby Formation shale exposed in an outcrop of the Little Rouge River approximately 1 km upstream of the Rouge River confluence.

Figure 3.14 - Sunnybrook Drift exposed in a terrace downstream of Meadowvale Road.

Page 54: Downstream Variability of Fluvial Form, Process, and Character in a

43

The Halton Till, and its correlatives the Northern and Newmarket Till, represent

the most spatially extensive glacial unit within the watershed having been deposited

during the late Wisconsin period and forms the vast majority of surficial sediments within

the South Slope and Peel Plain physiographic regions. The till is generally regarded to be

stone and sand-rich compared to the Sunnybrook Till, and incorporates sand and gravel

inclusions derived from the underlying glacial deposits (Fig. 3.15) (Eyles and Boyce,

1991). Much of the river channel between Meadowvale Road and 14th Avenue is cut into

this till as documented in bluff analysis by Karrow (1967) and Eyles and Boyce (1991)

and it seems likely that this is the case for the channel farther upstream. Drift thickness is

variable but likely ranges between a couple meters to at least 40 meters (Eyles and

Boyce, 1991).

Figure 3.15 - Halton Till exposed in a low-lying terrace downstream of 14th Avenue.

Page 55: Downstream Variability of Fluvial Form, Process, and Character in a

44

Overlying the Halton Till are superficial lacustrine sands, silts, and gravels. The

most prominent of which are those in the south deposited by Lake Iroquois. Sediment

thickness of the near-shore gravel deposits typically does not exceed 5 meters but reduces

to less than a meter for the offshore silts and clays, which only sporadically drape Halton

Till (Eyles and Boyce, 1991). Though their spatial extent is not known, outwash gravels

upstream of the Lake Iroquois shoreline have been documented overlying Halton Till in a

valley bluff and seem to be several meters thick (Mahaney and Hancock, 1993).

Lacustrine deposits above the Iroquois shoreline are typically patchy and composed of

clays to sands reaching a maximum thickness of several meters (TRCA, 2007a). The

complexity of the Oak Ridges Moraine sediments in the northernmost part of the

watershed is significant, but, in brief, much of the surficial sediments are composed of

silts and sands with pockets of gravel deposited through glaciofluvial and glaciolacustrine

sedimentation (Barnett et al., 1998).

The overall drift thickness within the watershed tends to decline downstream from

a thickness of around 170 metres within the Oak Ridges Moraine complex to 5 metres

near the confluence with the Rouge River (Fig. 3.16). In many cases the glacial units tend

to thin down-valley, which combined with progressive incision means valley bluffs tend

to be composed of several stratigraphic units with the number increasing downstream. In

general, and as alluded to in the previous paragraphs, the channel appears to be down

cutting into older glacial deposits with increasing distance downstream.

Page 56: Downstream Variability of Fluvial Form, Process, and Character in a

45

Figure 3.16 - Geologic cross-section through the Rouge River watershed. Overall drift thickness tends to decrease downstream while channel incision increases exposing multiple glacial units in valley bluffs. Note: valley sections are for the Rouge River and not the Little Rouge, though they should be fairly similar (after Eyles and Boyce, 1991).

3.6 Study Sites

As channel slope appears to be a dominant control on fluvial form (ie. Ferguson

and Ashworth, 1991; Nanson and Croke, 1992) and is easily determined through maps

and GIS software, channel slope was the primary basis for selecting study sites. A

longitudinal profile of the channel was created from a 10-metre digital elevation model

and 1-metre contour interval data of the drainage basin, and sites that represented various

gradients along the longitudinal profile were selected (Figs. 3.17 and 3.18). Site selection

was supplemented by field reconnaissance, land accessibility, and digital imagery

analysis of river plan-form to ensure a variety of fluvial forms were chosen. Due to time

constraints, only a handful of sites could be chosen, which represent only a portion of the

variability within the watershed. However, they are thought to be representative of the

range of slope and boundary conditions. Table 3.3 provides an overview of major site

characteristics prior to data collection and imagery of each site are given in Figures. 3.19a

and b.

Page 57: Downstream Variability of Fluvial Form, Process, and Character in a

46

Figure 3.17 - Location of study sites within the watershed.

Page 58: Downstream Variability of Fluvial Form, Process, and Character in a

Figure 3.18 - Location of study sites in the context of the longitudinal profile. Selected road names are given for reference. The longitudinal profile was derived from combining a 10-metre digital elevation model (first 3.145 kilometres) with 1-metre contour interval data (rest of the profile). Both datasets were created by the Ontario Ministry of Natural Resources. Approximate boundaries of the physiographic regions are given. Vertical exaggeration = 78x.

Page 59: Downstream Variability of Fluvial Form, Process, and Character in a

Site Drainage Area (km2)

Slope Channel Confinement

Plan-form Sinuosity Bed Structure

1 16.99 0.0024 Unconfined Single-thread, no bars

High Plane Bed

2 21.70 0.0071 Unconfined Single-thread with medial bars and poorly developed lateral bars

Low Run/Riffle-Pool

3 73.14 0.0004 Unconfined Single-thread, no bars

Moderate-High

Plane Bed

4 86.60 0.0027 Semi-confined to Unconfined

Single-thread with poorly developed lateral and point bars

Low-Moderate

Riffle-Pool

5 107.42 0.0057 Confined to Semi-confined

Single-thread with variably developed lateral, medial, and

point bars

Low-Moderate

Riffle-Pool

6 114.12 0.0044 Unconfined Single-thread with well developed lateral, medial, and point bars

Low-Moderate

Riffle-Pool

Table 3.3 - Study site characterization prior to fieldwork. Channel slopes were derived from the longitudinal profile in Figure 3.18.

Page 60: Downstream Variability of Fluvial Form, Process, and Character in a

49

Figure 3.19a - Imagery of sites 1-3. Images on the left are field images, while images on the right are Google Earth™ imagery. Dense forest canopy prevented an aerial image of Site 4 from being obtained.

Page 61: Downstream Variability of Fluvial Form, Process, and Character in a

50

Figure 3.19b - Imagery of sites 4-6. Images on the left are field images, while images on the right are Google Earth™ imagery.

Page 62: Downstream Variability of Fluvial Form, Process, and Character in a

51

Chapter Four: Methods 4.1 Fieldwork

Fieldwork was carried out at the study sites during the months of June to August

2011 with supplemental data collected in October. Data collection consisted of levelling

surveys of the channel, channel-bed augering, grain size analysis, and floodplain

sampling.

4.1.1 Surveying and Bed Augering

At each site, 5 to 11 cross-sections of the channel were taken in addition to the

longitudinal profile, using a Sokkia™ engineering level and stadia rod, over a distance

equivalent to approximately 9 to 32 bank-full channel widths with a mean of 20.5

channel widths. Cross-sections were split nearly even between riffles and pools. Intervals

between elevation measurements in each cross-section were typically on the order of 50

cm or less in order to characterize the cross-sectional shapes as best as possible.

Following surveying, up to four points were selected evenly spaced across the alluvial

section of the channel to be manually drilled into using a Dutchfield auger until glacial

clay or bedrock was reached, or refusal due to cobbles. Depth to clay or refusal was then

recorded with the distance along the cross-section. Lateral changes in bed composition

(ie. glacial clay, gravel, sand) were also noted.

4.1.2 Grain Size Analysis

At each riffle, the Wolman (1954) pebble count method was employed to

characterize the distribution of the surface bed material. A clast was selected every 50 cm

along the cross-section and placed through a template with square holes equivalent to b-

axis diameters ranging from 4 mm up to 256 mm spaced in half-phi intervals. Several

Page 63: Downstream Variability of Fluvial Form, Process, and Character in a

52

lines parallel to the cross-section were taken until a sample size of at least 100 clasts was

attained. This sample size was chosen because it is expected that the sample variance will

begin to stabilize at this point (Wolcott and Church, 1990). Boulders (b-axis diameters

>256 mm) that were included within the sample were measured with a measuring tape.

Bulk sampling was employed at sites 1 and 3 as bed material here consisted of a

considerable amount of sand, which is beyond the limits of the template. Since obviously

defined riffles did not exist at these sites, samples were taken from shallow locations

within the channel that were felt to be analogous to riffles. At each site, 3-4 buckets were

filled with 10-15 kg of sediment taken from the channel bed to a depth of approximately

half the blade length of the shovel. The samples were taken back to the lab,

disaggregated, and dry sieved through screens of diameters between 0.045 mm up to 64

mm in half-phi increments for 15 minutes. The a, b, and c axes of the largest sediment

clast within each study site were also measured.

Grain size distributions were pooled together to give site-average distributions, as

advocated by Wolcott and Church (1990). Where pebble counts were used, this resulted

in sample sizes between 200 to 600 particles, but typically at least 300. Rice and Church

(1996) suggested that precision in the distribution at the 95% confidence level does not

get appreciably greater beyond 400 particles. Therefore, the site-averaged distributions

derived are expected to be reliable. Combination of the individual bulk samples from

sites 1 and 3 yielded total weights of 32 and 46 kg respectively, with the largest particle

class representing, respectively, 3.2% and 1.7% of the distribution by weight. Church et

al. (1987) have advocated that the largest class should not constitute more than 1% by

Page 64: Downstream Variability of Fluvial Form, Process, and Character in a

53

weight. Therefore, the precision of the upper end of the distributions may not be as strong

as desired.

4.1.3 Floodplain and Channel Bank Sampling

Floodplain data were collected from bank exposures, manually dug pits, auger

holes, and sediment cores. Banks were initially cleared of slumped material in order to

reveal the entire stratigraphy. Individual strata that were distinct from one another based

on apparent grain size and/or colour were identified and their respective thickness

measured and recorded. Each stratum was then described making note of features such as

colour, weathering, grain size, the presence of organics, sedimentary structures (ie. cross-

beds, imbrication, and grading), and the nature of the contact with the overlying stratum.

Sediment samples of strata composed of sediment in the sand fraction or finer were taken

for laboratory particle size analysis. In instances where the sand stratum was particularly

thick, two samples were taken – one from near the bottom and one from near the top – to

discern differences in the sediment sizes upward through the deposit. Organic samples

were also taken where appropriate.

Pits were selectively dug on the inside of meander bends where the overbank

thickness was less than 50 cm. Where overbank thickness exceeded 50 cm, a hand-driven

Oakfield Auger was used until refusal, typically defined by the gravel surface, and the

sediment laid out onto a tray for examination. The same procedure described for the

cutbanks above was used to characterize the sediments, and samples of organic material

and representative grab samples of the strata were collected. At selected locations at Site

6, a hand-driven JMC Environmentalist’s Soil Probe was used to take narrow-diameter

Page 65: Downstream Variability of Fluvial Form, Process, and Character in a

54

sediment cores of the floodplain for further lab analysis. Refusal was fluvial gravel,

though these deposits were not often reached.

4.2 Lab Analysis of Floodplain Sediments

For fine floodplain analysis, selected floodplain sediments from each site were

dried at 100°C overnight, disaggregated using a mortar and pestle, weighed, and left to sit

in 100 mL of 5% sodium hexametaphosphate and then topped up with 400 mL of

distilled water for 24 hours. The samples were wet sieved through a 63 μm screen. The

sand fraction was retained, dried, and weighed in order to determine the percent silt and

clay in each sample. Loss-on-ignition was also performed on selected floodplain samples

for organic content. Samples were first dried, weighed, and heated to 550°C for an hour

(Dean, 1974). Following this, the samples were cooled to room temperature and re-

weighed, from which percent organics was derived. Sediment cores were analyzed for

major lithostratigraphic breaks, sediment structures, colouration, and grain size based on

hand texturing.

4.3 Determination of Morphological Bank-full Discharge

The determination of morphological bank-full discharge has many important

implications for morphological, sediment transport, and flood frequency relations

(Navratil et al., 2006). Indeed, an accurate representation of morphological bank-full

discharge forms a major component of this study, as it is necessary for equilibrium

relations, stream power computations, plan-form analysis, and bank-full discharge

frequency analysis. Thus, it was imperative to determine the best indicator(s) of

morphological bank-full stage and how best to calculate the corresponding discharge.

Page 66: Downstream Variability of Fluvial Form, Process, and Character in a

55

In any given river reach, there usually is not a single feature that indicates bank-

full stage. More often, multiple indicators exist and it is up to the researcher to determine

which is the correct one at any given cross-section. It is for this reason that Williams

(1978) and Navratil et al. (2006) advise that a reach-average bank-full estimate be

calculated incorporating the measurements from multiple cross-sections. The

morphological bank-full indicators used within the study were the minimization of the

width-depth ratio, bar-top elevation, the limits of vegetation (ie. shrubs or trees), and the

elevation of the active floodplain (see Williams (1978) for a full list). It was often the

case that the minimization of the width-depth ratio (or a close approximation of it)

coincided with the elevation of the active floodplain and this indicator appears to have

been reliable in a regional study conducted in nearby watersheds (Annable et al., 2010).

Each cross-section was viewed graphically to assess bank-full stage in the context of the

four indicators and compared to orthoimagery where possible. A portion of Site 5

experienced localized down cutting of 70 cm above bar-top surfaces and 90-130 cm

above summer low-flow stage. Since many of the aforementioned indicators do not work

for entrenched channels, the stage at these cross-sections needed to be critically assessed

in order to not over-estimate bank-full stage and thus discharge. Once a stage was

determined, bank-full width and average depths were calculated for each cross-section

and compiled to form a reach average cross-section.

Discharge was estimated through the use of four relations. In his analysis,

Williams (1978) formulated a regression relation using cross-sectional area (Abf, m2) and

channel slope (SC, - ) given as

1.28 0.28bf bf CQ = 4A S 4.1

Page 67: Downstream Variability of Fluvial Form, Process, and Character in a

56

The other three relations use an estimate of velocity (Vavg) multiplied by the

cross-sectional area. Bray (1979) reviewed the velocity calculation problem and

determined that the two best performing relations are (1) the Limerinos-Manning relation

0.67 0.5h C

avg

R SV =

n 4.2

where Rh is the bank-full hydraulic radius (m) and n is the Manning’s roughness value

estimated as

0.17h

-1h 84

0.113Rn =

1.16+2log(R D ) 4.3

where D84 is the grain size that represents the coarsest 84th percentile of the grain size

distribution (m), and (2) the Lacey relation

0.67 0.33avg h CV = 10.8R S 4.4

The final velocity relation utilized was developed by Hey (1979) and uses the

Darcy-Weisbach relation defined as

0.5

h Cavg

8gR SV =

ff

4.5

where g is the gravitational acceleration (9.81 ms-2) and ff is the friction factor given by

Hey (1979) as

2

h

84

13.4R = 2.03log

3.5Dff

4.6

Before these relations were used to calculate bank-full discharge at each study

site, they were first tested on independent field data in order to assess the reliability of

each relation. The results of this analysis are presented in Appendix C.

Page 68: Downstream Variability of Fluvial Form, Process, and Character in a

57

Chapter Five: Watershed-Scale Variability of Fluvial Form and Process

5.1 Introduction

The study sites were divided into five channel-floodplain types based primarily on

specific stream power, accretionary style, and channel plan-form. Accretionary styles are

a reflection of the energy conditions and sediment size within each channel-floodplain

type. Using data acquired from cross-sectional surveys, floodplain sampling, and

stratigraphic profiles, these channel-floodplain types are described with reference to

important physical and sedimentary controls. The derivation and rationale for these

controls are too extensive to be explained here but are analyzed in detail in Appendix F.

Bank-full specific stream power, based on data from the study sites, is assigned to each

type for comparison to other channel and floodplain classification schemes.

The primary bank-full morphological, hydrological, and sedimentological

characteristics of each site along with floodplain characteristics are presented in Table

5.1. A more detailed set of characteristics, including the full data-set, and the derivation

of the bank-full discharges and return periods are provided in appendices A, B, C, and D.

Major characteristics of the channel-floodplain types are summarized in Table 5.2.

To characterize floodplain stratigraphy at each site, a selected group of bank

exposures, auger holes, and pits were analyzed based on sediment structures, colour,

texture, and other relevant properties. Grain size variations shown are derived from data

presented in Appendix B and from hand texturing of sediment cores. Facies units were

assigned to each core and bank log based on Table 5.3 and broad depositional

environments are interpreted.

Page 69: Downstream Variability of Fluvial Form, Process, and Character in a

Site Site Characteristics 1 2 3 4u 4d 5 6 Drainage Area (km2) % Total Area

16.99 14.84

21.70 18.96

73.14 63.88

86.53 75.57

86.65 75.67

107.42 93.82

114.12 99.67

Distance Downstream (km) 6.45 8.29 21.77 27.80 28.00 38.42 43.36 Channel Slope 0.0002* 0.0072 0.0002 0.0017 0.0043 0.0069 0.0035 Channel Morphology Width (m) 5.64 0.51 9.42 0.77 9.11 0.80 13.38 0.23 13.64 0.74 17.45 1.67 20.51 1.77 Depth (m) 0.68 0.04 0.65 0.03 1.04 0.09 0.73 0.12 0.73 0.08 0.70 0.05 0.81 0.10 Width:Depth 8.29 14.49 8.76 18.33 18.68 24.93 25.32 Area (m2) 3.86 0.42 6.07 0.62 9.75 1.63 9.78 1.55 10.09 1.37 12.11 1.82 16.14 1.95 Sinuosity 1.71 1.14 1.70 1.15 1.18 1.17 1.26 Channel Hydrology Velocity (ms-1) 0.49 1.47 0.59 1.08 1.41 1.68 1.48 Discharge (m3s-1) 1.87 8.95 5.57 10.59 14.02 20.55 24.64 Return Period (years) 0.50 20.00 0.50 1.10 2.00 3.00 5.00 Channel Sedimentology Bed Thickness (cm) 38 2 45 1 58 3 37 1 26 2 50 1 91 3 Percent Clay (%) 15.03 7.50 4.19 4.19 5.20 5.20 0.76 0.76 25.98 9.53 16.02 4.60 4.26 4.26 D50 (mm) 8 36 3 62 56 54 31 Sorting () 3.09 1.67 2.80 0.94 0.78 1.14 0.82 Floodplain Sedimentology Overbank Thickness (cm) ~ 45† 82 8 145 12 63 11 64 9 20 5 ~ 112† Silt/Clay Content (%) 63.41 5.87 59.91 5.31 76.62 2.96 49.82 5.83 47.11 4.71 38.51 3.49 52.31 6.09 Organic Content (%) 6.70 0.69 4.88 0.74 4.28 1.06 4.46 1.08 3.65 0.73 1.47 0.17 1.58 0.21

Table 5.1 - Bank-full channel and floodplain characteristics of the study sites. Drainage areas derived from 10-m digital elevation model provided by the Ontario Ministry of Natural Resources. Downstream distances derived from Figure 3.18. Error ranges represent one standard error. * - the water surface slope at the time of survey is substituted for channel slope as a better measure of the energy condition at the study site. Channel slope is an order of magnitude higher. † - values are estimates of overbank thickness based on floodplain logs.

Page 70: Downstream Variability of Fluvial Form, Process, and Character in a

Type Plan-form Specific

Stream Power (Wm-2)

Floodplain Accretion

Style

Channel Accretion

Style

Width-Depth Ratio

Dominant Bed

Material

Representative Study Sites

Low Energy Mixed-Load

Sinuous single-thread with no bar-forms

<2 Oblique Lateral and oblique

8 Sandy-gravel

1 and 3

Low Energy Bed-Load

Low sinuosity single-thread with no to weakly developed bar-forms

13 Oblique Lateral 18 Gravel-cobble

4u

Mixed Energy

Low sinuosity single-thread with variously developed bar-forms

12-70 Oblique Lateral and vertical

15 Gravel-cobble

2

Medium Energy

Low sinuosity single-thread with weakly to well-developed bar-forms

40-45 Vertical Lateral and vertical

18-25 Gravel-cobble

4d and 6

High Energy Low sinuosity single-thread with well-developed bar-forms

80+ Vertical Lateral and vertical

25 Gravel-cobble

5

Table 5.2 - Description of the channel-floodplain types identified within the Little Rouge River watershed.

Page 71: Downstream Variability of Fluvial Form, Process, and Character in a

60

Facies Code Facies Description Interpretation Gcm Massive clast-supported gravel/cobble;

not imbricated

Lag deposits

Gh Massive or weakly horizontally-bedded clast-supported gravel; often imbricated

Bed-load deposits

Gms Massive matrix-supported gravel with coarse sandy-matrix; weakly imbricated

Bed-load and suspended-load mixture

Sm Massive sand or silt; no visible structures

Deposited out of suspension or as bed-load

Pi Incipient soil development; roots common; very similar to facies Sm

Minimal sediment deposition

P Well developed soil profile or paleosol; roots prevalent; likely similar to facies Sm

Minimal to no sediment deposition

Table 5.3 - Facies classification of the study sites (adapted after Miall, 1996).

5.2 Low Energy Mixed-Load Type

Low energy, mixed-load channels are characterized by sites 1 and 3. A

representative cross-section illustrating floodplain sediments and their depositional

mechanisms are shown in Figure 5.1. Characteristic specific stream powers are very low

(<2 Wm-2), which is primarily driven by very low channel slopes. As a result of

transport-reduced conditions, sand is often deposited on the channel bed blanketing

gravels inherited from past channel states, resulting in very poorly sorted channel beds.

The increase in the importance of sand to sediment transport relative to the generally

immobile gravel results in an increase in the importance of suspended sediment transport

in addition to bed-load transport. Consequently, channels have low width-depth ratios

(<10) and are generally sinuous in plan-form. Bed-forms were not observed and the

structure of the channel bed seems to take on a lower regime plane-bed morphology.

Plotting sites 1 and 3 on a plan-form discrimination plot developed by Kleinhans and van

Page 72: Downstream Variability of Fluvial Form, Process, and Character in a

61

Figure 5.1 - Field image taken during the summer (a) and idealized schematic cross-section (b) of the low energy, mixed-load channel-floodplain type. The diagram is based on cross-sectional surveys, lateral bed material identification, augering into the channel bed with a Dutchfield auger, and channel bank stratigraphy. The flow direction is given by the white arrow and the direction of channel migration is the same as depicted in the cross-section. Elevation given on the cross-section is in metres above an arbitrary datum.

den Berg (2011) that uses potential specific stream power and median grain size of the

bed material shows that Site 3 plots in the laterally immobile with no bars domain, which

is congruent with the plan-form style observed in the field (Fig. 5.2). Site 1 however,

plots in the meandering domain, which is a consequence of beaver dams along the

channel. While the channel is naturally at a higher energy state, the presence of beaver

Page 73: Downstream Variability of Fluvial Form, Process, and Character in a

62

dams lowers the water surface slope (and thus energy slope) of the channel enough to

change the channel behaviour.

Figure 5.2 - Plan-form discrimination of channel-floodplain types based off of the model provided by Kleinhans and van den Berg (2011) relating potential specific stream power to median bed material size.

Individual sites are labelled. Potential specific stream power (pv) is defined as bf Vpv 0.5

bf

ρgQ Sω =

aQ, where

is the density of water (1000 kgm-3), g is gravitational acceleration (9.81 ms-2), Qbf is the bank-full discharge (m3s-1), SV is the valley slope (-), and ‘a’ is a coefficient that is different for sand-bed (4.7) and gravel-bed (3.0) rivers.

According to the classification scheme presented by Nanson and Croke (1992),

these low energy, mixed-load channels are comparable to the laterally stable, single-

channel types that develop their floodplains through vertical accretion of overbank

sediments forming low-lying levees proximal to the channel with distal backswamps.

However, levees are absent at sites 1 and 3 suggesting that, despite multiple overbank

flooding events per year, vertical accretion is not likely a significant depositional process

Page 74: Downstream Variability of Fluvial Form, Process, and Character in a

63

at these sites, which likely explains why post-settlement alluvium was absent on the

floodplains.

Page et al. (2003) offer a model of floodplain development for low energy

migrating channels with negligible floodplain relief and pervasive overbank flooding,

much like the low energy sites presented here, where the dominant floodplain formative

process is the progressive erosion of the outer channel bank and lateral accretion of

suspended sediments against the inner channel bank (termed oblique accretion). It is

believed that this process is active at sites 1 and 3.

Floodplain formation at sites 1 and 3 is achieved initially through the lateral

accretion of coarse pebbly sand transported as bed-load over an immobile gravel lag,

which form steeply dipping (17), structureless sub-aqueous point bars (Figs. 5.1b and

5.3). These point bars may fine upwards to silty-sand, which represents a lateral gradation

of grain size, and they often contain small shells and abundant wood pieces at, or just

below, the low water surface. A sample of these wood deposits yields a radiocarbon age

of 3050 30 years BP at Site 3 indicating floodplain formation during at least the mid- to

late-Holocene and likely very slow sedimentation and channel migration (Table 5.4).

Suspended sediments are accreted to the inner bank above the point bar forming

steeply dipping (up to 25) banks. In this way, sedimentation is limited to the inner bank

of the channel and slow, progressive migration produces a relatively flat floodplain.

These overbank deposits have a diffuse contact with the underlying channel sands and

consist of massive sandy-silts that show no strong indication of upward fining. In some

cases, distinguishing between channel and overbank deposits is difficult. Soil profiles are

very well developed owing to the lateral stability of the channel.

Page 75: Downstream Variability of Fluvial Form, Process, and Character in a

Figure 5.3 – Characteristic floodplain stratigraphy of the low energy, mixed-load type based on S3-B1.

Page 76: Downstream Variability of Fluvial Form, Process, and Character in a

65

Site Lab Number Material 14C Age Two-Sigma Calibrated Range 3 Beta-315421 Wood 3050 30 BP 3360-3210, 3180-3170 BP 6 Beta-315420 Wood 140 30 BP 280-170, 150-60, 50-0 BP

Table 5.4 - Radiocarbon dates from the study sites. Uncalibrated dates are reported with one standard deviation. Dates calibrated using INTCAL09. Before present (BP) = before 1950 AD.

Based on sediment logs, laterally accreted channel deposits consist of

approximately half of the fine floodplain sediment while obliquely accreted deposits

account for the other half. Where point bars do not form, obliquely accreted deposits

dominate the entire fine sediment sequence. Vertical accretion, if it does occur, typically

makes up less than 5% of the floodplain stratigraphy (Page et al., 2003). The thickness of

the fine floodplain deposits is controlled by the quantity of incoming fine sediment from

upstream. Upward fining trends in the sediment may also occur.

The channel and floodplain morphology and sedimentology described for sites 1

and 3 are very similar to that reported for other low-energy fine-grained channels

(Jackson, 1981; Miall’s (1996) Model 7; Page et al., 2003) suggesting that these

accretionary processes may be pervasive amongst low energy migrating rivers. However,

differences exist between these studies and the floodplain sedimentology described here.

Primarily, the floodplain stratigraphy for sites 1 and 3 are more simplistic than that

reported in the literature. Page et al. (2003) report that obliquely accreted sediments often

consist of couplets of fine sand and silt, which have been similarly described by Jackson

(1981) and Miall (1996). These couplets are absent at sites 1 and 3 giving the sediment a

structureless appearance. The sand thickness of a couplet tends to take on approximately

a 3:1 ratio with the silt (Page et al., 2003). Conversely, the obliquely accreted deposits at

sites 1 and 3 have a far greater silt component by weight than sand (~3:1), suggesting that

the paucity of sand within the deposits precludes couplets from forming. Even still, any

Page 77: Downstream Variability of Fluvial Form, Process, and Character in a

66

couplets that may form would be very thinly laminated and would likely be destroyed due

to bioturbation.

5.3 Low Energy Bed-Load Type

The low energy, bed-load channel-floodplain type is based on Site 4u and shown

schematically in Figure 5.4. The channel plan-form takes on a low-sinuosity course with

a riffle-pool bed structure. Bar-forms are generally absent but are weakly developed

when visible. Specific stream power is low (~13 Wm-2) due to low channel slopes,

Figure 5.4 - Field image taken during the summer (a) and idealized schematic cross-section (b) of the low energy, bed-load channel-floodplain type. The flow direction is given by the white arrow and the direction of channel migration is the same as depicted in the cross-section.

Page 78: Downstream Variability of Fluvial Form, Process, and Character in a

67

however, is high enough to prevent significant sand deposition as in the low energy,

mixed-load type. The channel bed is composed of glacially derived gravel-cobble for

which the channel is competent to transport only a small fraction. Due to the lack of

transport capacity within the channel, glacial clay is not often exposed. However, Figure

5.4 does show that there is a tendency for the channel to plane off the underlying glacial

surface as it progressively migrates, which means that the lag deposits must be mobilized

at some point, perhaps during exceptionally large magnitude flood events.

Plan-form discrimination predicts that this type is a laterally immobile channel

with no bars, which is in agreement with field observations (Fig. 5.2). The well-

developed soil profiles characterized by relatively high organic contents provide

inferential evidence that the channel is very laterally stable, which is gained through low

stream power and resistant, root-stabilized banks.

The Nanson and Croke (1992) floodplain classification scheme suggests that the

floodplain developed by this channel-floodplain type is created through progressive

lateral accretion of point bars capped with vertically accreted deposits, which is broadly

confirmed here. It is suggested, though, that the floodplain is formed through the oblique

accretion of sandy-silt, similar to that described above for low energy, mixed-load

channels. Initial floodplain development involves the lateral accretion of fine-medium

gravel over an immobile coarse gravel-cobble lag producing a vertical fining trend (Fig.

5.5). Deposition appears to be reduced to the inner third of the channel while the

remaining two-thirds are dominated by the immobile lag. The laterally accreted gravel

may be intermixed with small shells, wood pieces, and sand that increases in thickness

towards the inner channel margin. The depositional surface is sub-aqueous at low flow

Page 79: Downstream Variability of Fluvial Form, Process, and Character in a

Figure 5.5 – Characteristic floodplain stratigraphy of the low energy, bed-load type based on S4-B2 (a) and S4-AH5 (b).

Page 80: Downstream Variability of Fluvial Form, Process, and Character in a

69

and tends to dip shallowly (6º) towards the channel centre, which contrasts with the

steeply dipping laterally accreted surface described for the low energy, mixed-load

system.

As mentioned, it is believed that suspended-load sediment is accreted to the side

of the inner bank-face, which dips sharply into the channel while the floodplain is

relatively flat beyond the bank-full channel limits. Furthermore, recent (ie. post-

settlement) deposition on the floodplain surface is absent despite annual to sub-annual

overbank flooding, which argues against vertical accretion and may be indicative of the

capacity of overbank flows to carry coarse sands onto the floodplain. Fining upward

trends within the sandy-silt are commonly observed. The proportion of fine, obliquely

accreted sediments to coarse, vertically accreted sediments is about 1:1, similar to that of

the low energy, mixed-load system. Silt-sand couplets are absent due to a lack of

significant sand in the system, bioturbation, and pedogenic processes.

5.4 Mixed Energy Type

The mixed energy channel-floodplain type is very similar to the low energy, bed-

load type described above in terms of channel and floodplain deposition. Indeed the

schematic illustrations in Figures 5.4 and 5.6 are nearly identical. However, the primary

difference between the two types is the energy level within the channel. In the mixed

energy type, the channel tends to alternate between a low energy state (~12 Wm-2) and a

high energy state (~70 Wm-2) brought on by the presence of debris dams, where low

energy conditions prevail upstream of dams and high energy conditions prevail

downstream. Since the presence of large woody debris is an important characteristic, it is

Page 81: Downstream Variability of Fluvial Form, Process, and Character in a

70

likely that this channel-floodplain type is restricted to the upper portion of the watershed

where fallen trees easily span the width of the channel.

Figure 5.6 - Field image taken during the summer (a) and idealized schematic cross-section (b) of the mixed energy channel-floodplain type. The flow direction is given by the white arrow. Since the image in a is not the same as the cross-section in b, the direction of channel migration is not the same.

Plan-form predictions predict a meandering or weakly braided channel with bars

and chute cut-offs (Fig. 5.2), which is not entirely congruent with field observations.

Overall, the plan-form of the channel is fairly straight with occasional bar-forms of

various degrees of development. Most bar-forms develop downstream of debris dams or

behind fallen debris within the channel. The disparity between predicted and observed

Page 82: Downstream Variability of Fluvial Form, Process, and Character in a

71

plan-forms arises from the occurrence of low energy states along the channel, which are

more comparable to laterally immobile channels, in addition to bank reinforcement by

extensive tree rooting, which acts to prevent erosion and channel migration.

Sediment transport is strongly conditioned by debris dams and the hydrological

regime of the basin. Upstream of dams, only the finest gravels can be entrained. Due to

damming of the flow, bank-full stage is often reached several times throughout the year

meaning the finest gravel and granules are entrained frequently. Downstream of the

dams, much more of the gravel is entrained during bank-full events creating thinly

bedded mid-channel bars and gravel sheets. However, due to enlarged cross-sectional

dimensions, and the low and often quick peak flows, bank-full discharge is rarely

reached, and large-scale sediment transport is consequently not common. The structure of

the channel bed is typically riffle-pool, however, aggradation of fine gravel and granules

under low energy conditions creates a bed structure more akin to run-pool. The mixing of

fine gravels and granules with glacially derived coarse gravels and cobbles leads to a

poorly sorted, bimodal channel bed. Following the return to high energy conditions

through either breaching of the debris dam or diversion of the channel, the accumulated

fine gravel and granules can be quickly mobilized, which may result in a pulse of bed-

load under bank-full conditions. Alternatively, under sub-bank-full conditions, the

deposits may be gradually winnowed. Glacial clay is not readily exposed in the channel

due to aggradation of the channel bed or slow reworking of aggraded deposits. However,

glacial clay was observed downstream of debris dams where the gradient is steeper and

turbulence greater, and downstream of steep riffles.

Page 83: Downstream Variability of Fluvial Form, Process, and Character in a

72

Floodplain development is very similar to that described for the low energy, bed-

load dominated type and will be reiterated here (Figs. 5.4 and 5.6). Under high energy

conditions, the channel bed is composed of coarse-gravels and cobbles that are mobile

during infrequent bank-full conditions but less mobile or immobile during sub-bank-full

conditions. As a result they are considered lag deposits. Proximal to the inner margins of

the channel are laterally accreted fine to medium gravels and sandy-gravels that are

mobile under more frequent, but smaller discharges. Under low energy conditions, the

same general lateral gradation of sediments exists, however the proportion of proximal

sand increases while the proportion of fine gravel and granules increases towards the

outer, eroding channel bank. The lateral accretion of the fine gravels over coarse gravels

creates a fining upward sequence (Fig. 5.7). Wood deposits and small shells may be

found proximal to the inner bank.

Sandy-silt deposits carried in suspension are accreted obliquely to the inner bank

of the channel in much the same fashion as that described for the low energy, bed-load

type, which may result in an upward fining trend from silty-sand to sandy-silt. Channel

migration appears to be very slow, inferred from very well developed soil profiles (Fig.

5.7), which is a consequence of root-stabilized banks and generally low energy conditions

either through debris dams or infrequent bank-full flooding. It is of note that abandoned

channels were observed, which may indicate that, in addition to progressive lateral

migration, channel avulsion is important as well. Such avulsions may have been induced

by flow diversion around debris dams. A large scour pool adjacent to the present debris

dam and perpendicular to the active channel was observed indicating that headward

Page 84: Downstream Variability of Fluvial Form, Process, and Character in a

Figure 5.7 - Characteristic floodplain stratigraphy of the mixed energy type based on S2-B1 (a) and S2-AH1 (b).

Page 85: Downstream Variability of Fluvial Form, Process, and Character in a

74

incision of a new channel is in progress (Fig. 5.8), though the rate of incision is not

known.

Figure 5.8 - Development of an avulsion channel (outlined in black) around an active debris dam (to the bottom left of image). The active channel is shown in the top of the image with the flow direction given. The flow direction through the avulsion channel is given, as is the direction of headward incision. Standing water can be seen in the scour pool.

5.5 Medium Energy Type

The medium energy channel-floodplain type is characterized by sites 4d and 6,

and is associated with a specific stream power of about 40-45 Wm-2. The channel plan-

form between the two sites is strikingly different despite having similar energy levels.

This is due to the grain size of the bed material relative to the channel slope. Bed material

at Site 4d consists of glacially derived gravel and cobble-sized sediment, only half of

which can be transported during bank-full events, while at Site 6, the bed material is

Page 86: Downstream Variability of Fluvial Form, Process, and Character in a

75

dominantly gravel derived from size-selective transport from upstream, such that 90% of

the sediment is mobile at bank-full stage. The effect that sediment mobility has on the

channels’ plan-form is illustrated in Figure 5.2 where Site 6 is predicted to be weakly

braided while Site 4d is not. This prediction is broadly conformable with observed plan-

forms where occasional medial bars and extensive lateral and point bars are seen at Site 6

while Site 4d lacks these features (compare Figs. 5.9a and 5.10a). The overall structure of

the channel bed is riffle-pool. The width-depth ratio of the bank-full channel varies

between 18-25 and scales with the bed-load concentration. Under supply-limited

conditions, glacial clay or bedrock is readily exposed in channel pools and may make up

the majority of the channel bed. Channel stability was not examined in detail for this

study, but based on aerial imagery, Site 6 appears to be much more active than Site 4d.

This is evident in Figures 5.9 and 5.10 where vegetation along the channel grows almost

up to the low-flow level at Site 4d but a large unvegetated buffer exists at Site 6. The

suspected reasoning behind this is that the channel at Site 6 is cut into relatively erodible

alluvium, which allows the channel to migrate more easily than at Site 4d, which is

bound by resistant glacial clay. Consequently, vegetation has sufficient time to colonize

the channel margins.

Floodplains developed by this channel type tend to form through a combination of

lateral and vertical accretion (Figs. 5.9 and 5.10). At Site 4d, large-scale sediment

mobility exposes glacial clay along the outer half of the channel where a patchy coarse

gravel and cobble lag forms. Advancing overtop of the lag is a moderately dipping (13)

sandy, fine-medium gravel point bar that tends to gently fine upwards reflecting a lateral

Page 87: Downstream Variability of Fluvial Form, Process, and Character in a

76

fining trend towards the inner bank of the channel (Fig. 5.11). The surface of the point

bar tends to be exposed at low-flow revealing a dominantly sandy to fine gravely texture.

Figure 5.9 - Field image taken during the summer (a) and idealized schematic cross-section (b) of the medium energy channel-floodplain type (Site 4d). The flow direction is given by the white arrow and the direction of channel migration is the same as the cross-section.

Page 88: Downstream Variability of Fluvial Form, Process, and Character in a

77

Figure 5.10 - Field image taken during the summer (a) and idealized schematic cross-section (b) of the medium energy channel-floodplain type (Site 6). The flow direction is given by the white arrow and the direction of channel migration is the same as the cross-section.

Similarly, point bars and lateral bars at Site 6 possess a comparable accretionary

angle, though typically less than Site 4d, and exhibit a lateral gradation of sediment

consisting of medium gravel near the channel floor that fines towards the inner margin of

the channel. Bar surfaces at Site 6 consist of a mixture of gravel and sand, and are often

much more extensively sub-aerially exposed during low-flow, which is a result of higher

sediment supply and channel mobility than at Site 4d. Medial bars, though observed, are

rare, and form through vertical accretion onto stalled gravel slugs (Miall, 1996).

Overlying bar deposits at both sites are vertically accreted massive sands

deposited during overbank flooding that may weakly fine upwards. Return periods for

overbank flooding varies between 2-5 years. Where return periods are long, such as at

Site 6, individual flood layers may be evident separated by thin incipient soils with

rootlets, which represent periods of little flooding (Fig. 5.12b). Overbank deposits may

Page 89: Downstream Variability of Fluvial Form, Process, and Character in a

Figure 5.11 - Characteristic floodplain stratigraphy of the medium energy type based on S4-B3 (a) and S4-P1 (b).

Page 90: Downstream Variability of Fluvial Form, Process, and Character in a

Figure 5.12 - Characteristic floodplain stratigraphy of the medium energy type based on S6-B1 (a) and S6-C2 (b).

Page 91: Downstream Variability of Fluvial Form, Process, and Character in a

80

contain shells and occasional small pebbles. The thickness of the deposits varies with

sediment supply such that Site 6 has thicker overbank deposits than Site 4d.

The style of floodplain development described for this type is in good agreement

to that suggested by Nanson and Croke (1992), and exemplifies the overall point bar

morphology and sedimentology of other gravel-bed meandering rivers described in the

literature (Wolman and Leopold, 1957; Bluck, 1971; Jackson, 1976). Nevertheless,

differences between the sites described here and observations along other rivers exist.

Most notably, the lack of observed sedimentary structures compared to conventional

point bar models (Bluck, 1971; Jackson, 1976; Miall, 1996). Whether sedimentary

structures exist and were destroyed during the falling stages of floods or were just not

apparent during the field campaign remains uncertain.

Post-settlement alluvium (PSA) was observed at both sites, though is thicker and

more extensive at Site 6 due to a greater fine sediment supply from upstream. PSA was

distinguished from paleosols based on its light tan or grey colour, which is indicative of

lower organic contents, compared to the organic-rich brown colour of paleosols. The

presence of rootlets also indicates paleosol formation. Mahaney et al. (1989) documented

paleosols farther downstream of Site 6 but at approximately the same stratigraphic

position as that shown in Figure 5.12a. Radiocarbon dates for these paleosols are not

precise but are thought to have developed within the last one hundred radiocarbon years

(Mahaney et al., 1989). This is confirmed by an underlying radiocarbon date retrieved

from the lower sand unit (140 30 BP, Table 5.4), which indicates that the paleosol at

Site 6 had formed shortly before European settlement and that the floodplain is quite

young compared to the previously described sites.

Page 92: Downstream Variability of Fluvial Form, Process, and Character in a

81

Of particular interest is the difference in boundary conditions between sites 4d

and 6. As shown in Figures 5.9 and 5.10, Site 4d is underlain by glacial clay while Site 6

is underlain by alluvium. A sediment core taken from the floodplain of Site 6 allows for a

much more extensive interpretation of the underlying floodplain deposits, which may

provide an explanation on why alluvium, instead of glacial clay exists (Fig. 5.13).

Though basal gravels were never reached to a core sampling depth of 3.7 metres, the

bottommost part of the core is believed to represent a relic floodplain surface on the basis

of its brown colour. A small pebble was found within these deposits, which may indicate

high-energy conditions. Overlying this surface is a succession of three, possibly four,

coarse sand and granule deposits representing a channel environment separated by graded

silt-sand deposits indicative of overbank deposition. The basal contact of the inferred

channel deposits with the underlying overbank deposits are abrupt and perhaps erosional

in nature, while the contacts within individual overbank-channel couplets tend to be

either abrupt or diffuse. The uppermost channel-overbank sequence underlies the pre-

settlement paleosol and post-settlement alluvium discussed previously. The present day

floodplain surface is approximately 4 metres above present day Lake Ontario water levels

suggesting that the thick, multi-story floodplain deposits underlying the present-day

channel may be a result of aggradation following lake level rise during the mid- to late-

Holocene (Weninger and McAndrews, 1989; Anderson and Lewis, 2012).

5.6 High Energy Type

The high energy channel-floodplain type is characterized by Site 5 and shown

schematically in Figure 5.14. Characteristic specific stream powers are approximately 80

Wm-2 but higher values (~100 Wm-2) are expected in the steeper sections of the channel

Page 93: Downstream Variability of Fluvial Form, Process, and Character in a

82

Figure 5.13 – Extended floodplain stratigraphy of Site 6 based on S6-C1.

Page 94: Downstream Variability of Fluvial Form, Process, and Character in a

83

downstream of Site 5. Much like Site 6, this channel type is dominated by well-defined

lateral, point, and medial bars along the channel, which is a product of widespread bed

sediment mobility. The spatial extent of the bar-forms is dictated by the supply of the

coarse gravel from valley sides and upstream. Bed material is dominantly gravel-cobble

derived from valley sides. The bed structure is dominantly riffle- pool, however in higher

energy sections not studied here, segments of run-pool morphology were observed.

Figure 5.14 - Field image taken during the summer (a) and idealized schematic cross-section (b) of the high energy channel-floodplain type. The flow direction is given by the white arrow and the direction of channel migration is the same as the cross-section.

The plan-form of the channel takes on a low sinuosity course with variable lateral

and point bars, while medial bars are less prominent. The plan-form is comparable to

gravel meandering and wandering type 1 channels reported in the literature (Lewin, 1976;

Forbes 1983; Carson, 1986; Miall, 1996) and predictions agree well with observations

(Fig. 5.2). While bar-forms within this type are the most spatially extensive and

Page 95: Downstream Variability of Fluvial Form, Process, and Character in a

84

developed within the watershed, they are generally smaller than other gravely

meandering channels owing to the limited sediment supply inherent in semi-alluvial

basins. Due to pervasive bed-load sediment mobility and comparatively large supply,

bank-full width-depth ratios are around 25, which are comparable to the meandering

gravel-bed rivers described by Carson (1986). The channel is often confined within a

narrow (<40-50 metres wide) floodplain bound by terraces 2-6 metres high composed of

glacial clay. As a result, channel migration is limited and likely gradual.

Glacial clay often floors the low-flow channel and is composed of patchy gravels

and cobbles, though the degree to which clay is exposed is dependent upon sediment

supply and the angle of the channel bend. Sandy gravel point bars described previously

for Site 4d (medium energy) were observed at Site 5, though it is not clear how

prominent they are, and likely decrease in importance with increasing stream power.

Large gravel bar-forms are typically horizontally-bedded and consist of a coarse gravel-

cobble framework from which imbricated gravel of various sizes are deposited over,

producing a weakly defined upward fining trend (Fig. 5.15). Gravel point bars and lateral

bars often possess low accretionary angles (5º) dipping towards the channel and in some

places across the entire channel. Clay till rip-up clasts from the channel bed may also be

found within bar deposits. Overbank deposits tend to be much thinner than bar deposits

and consist of massive sand vertically accreted onto the bars during overbank flooding,

which recurs about 3 years. Upward grain size trends are inconsistent as both normal and

inverse grading conditions were documented. The existence of the occasional gravel clast

within the sand stratum indicates periodic energetic overbank flows onto the floodplain.

Page 96: Downstream Variability of Fluvial Form, Process, and Character in a

Figure 5.15 - Characteristic floodplain stratigraphy of the high energy type based on S5-B1 (a) and S5-P2 (b).

Page 97: Downstream Variability of Fluvial Form, Process, and Character in a

86

Based on the bank-full specific stream power, floodplain development here

should be comparable to that of wandering and braided channels as per Nanson and

Croke’s (1992) classification scheme. Braided and wandering rivers develop their

floodplains predominantly through channel avulsion and abandoned channel accretion,

and vertical accretion of fine overbank sediment. Additionally, wandering rivers also

experience lateral accretion of gravely point bars. The inferred floodplain formative

processes at Site 5 diverge from those described for braided and wandering rivers. While

accretion of abandoned slough channels may occur episodically, floodplain development

at Site 5 is thought to occur through the progressive lateral accretion of gravel point bars

behind the laterally migrating channel and vertical accretion of overbank sand. The

incidence of channel avulsion is not known but may be limited due to floodplain

confinement, vegetated floodplains, infrequent bank-full flooding, and observed channel

down-cutting.

Comparisons of the stratigraphy for Site 5 with Miall’s (1996) idealized

stratigraphic logs for gravel meandering, wandering, and braided channels reveal broad

similarities in sediment facies, though his models include structured sand facies, often

seen in such rivers (e.g. Forbes, 1983; Desloges and Church, 1987; Brierley, 1989), that

were not observed here. Floodplain stratigraphy for a braided gravel-bed river described

by Reinfelds and Nanson (1993) conforms well with the floodplain deposits at Site 5 in

terms of sediment facies, lack of structures in the overbank fines, alluvial thickness, and

relative thickness of the overbank fines to the bar deposits suggesting that depositional

processes are similar.

Page 98: Downstream Variability of Fluvial Form, Process, and Character in a

87

5.7 Equilibrium Morphology Prediction

One method in assessing whether the channels have reached an equilibrium form

with the controlling variables is to predict the channel dimensions using regime relations

and analyze why the channels deviate from the predicted values. An important

consideration is the selection of regime relations to be used. As discussed in Chapter

Two, empirical relations are often used but lack the physical justification to explain the

resulting channel forms and are often developed under limiting constraints, such as

specific hydro-climatic regions. Moreover, if the goal is to predict equilibrium

dimensions, then empirical relations introduce the uncertainty that the data the relations

were derived from are, themselves, in equilibrium. Rational regime relations do not suffer

these issues as they are developed from physical relations instead of empirical data.

A second consideration is the appropriateness of using bank-full discharge for the

determination of channel dimensions. Since the bank-full discharge was determined from

the dimensions themselves, predicting the dimensions based off of bank-full discharge

would be circular. This argument has been raised by Ferguson (1986) and it has been

suggested that a frequency-based discharge be used instead since such a discharge would

be independent of the channel dimensions. In spite of this concern, however, many

regime relations developed still utilize bank-full discharge in their formulation.

Millar (2005) developed a set of theoretical rational regime relations based upon

the Maximum Transport Efficiency extremal hypothesis, which states that under the

constraints of discharge, grain size, bank vegetation, and channel slope, the channel will

adopt a preferential pair of stable dimensions in order to transport the given sediment

Page 99: Downstream Variability of Fluvial Form, Process, and Character in a

88

supply without long-term deformation of the channel bed through erosion or deposition.

His relations for bank-full width (wbf, m) and depth (dbf, m) are given as

0.70 -0.75 0.60 -1.10bf bf 50 Cw = 6.23Q D S μ’ 5.1

0.16 0.60 -0.62 0.64bf bf 50 Cd = 0.10 Q D S μ’ 5.2

where Qbf is bank-full discharge (m3s-1), SC is the channel slope (-), D50 is median bed-

material size (m), and μ’ is a dimensionless index value for bank strength relating the

relative resistance of the bank to the channel bed. A value of 1 indicates equal resistance,

while any value greater than 1 represents the effects of increasing vegetation density.

Theoretically, values cannot be below 1. Since bank strength cannot easily be measured

in the field, it was set to 1 and the other variables substituted into the relations to derive

the theoretical equilibrium channel morphology.

The results of this analysis show that there is a fairly even split between under–

prediction and over-prediction for both width and depth for the study sites (Fig. 5.16). In

their analysis on gravel-bed rivers, Parker et al. (2007) attributed deep and narrow

channels to dense bank vegetation and/or a decrease in sediment supply to the channel

from upstream or from eroding banks, though the two are often inter-related. Back-

calculating the bank strength indices using bank-full discharge, channel slope, median

grain size, and the field-measured width-depth ratio of the channel (as per Millar’s (2005)

guidelines) reveals values varying between 0.42 and 1.94 (Table 5.5). The values

calculated for sites 2, 3, 5, and 6 are in agreement with field-observed vegetation patterns

where Site 3 is dominated by grass (low bank strength) and Site 2 consists of root-

dominated banks (high bank-strength). Sites 5 and 6 are in between with more grasses

than trees lining the banks. The remaining sites, however, are in disagreement with their

Page 100: Downstream Variability of Fluvial Form, Process, and Character in a

89

Figure 5.16 - Comparison of field-derived channel widths (a) and depths (b) against predicted rational regime dimensions. Error bars around the field-derived measurements are one standard error. Site numbers given for reference.

calculated indices and sites 1 and 4u are drastically different. While it could be argued

that the poor prediction achieved at Site 1 is a result of the channel bed having a mixed

sand-gravel composition, which the relations were not necessarily designed for, Site 3

possesses a finer median bed grain size and was well predicted. Since, in theory, bank

strength values cannot be below 1, the calculated bank strength values in Table 5.5 may

Page 101: Downstream Variability of Fluvial Form, Process, and Character in a

90

indicate the presence of disequilibrium forms at sites 1 and 4u, and perhaps less so for

Site 4d.

Site Bank Strength Index Vegetation Conditions 1 0.43 Grasses, shrubs, and minimal trees 2 1.93 Thick forest canopy, shrubs, and dense rooting 3 1.07 Grasses

4u 0.42 Thick forest canopy, shrubs, and dense rooting 4d 0.94 Grasses, shrubs, and trees on outside bank 5 1.29 Grasses, shrubs, and some trees 6 1.28 Grasses, shrubs, and some trees on outside bank

Table 5.5 - Relative bank strength values and associated vegetation conditions of the study sites.

The dimensions predicted by Equations 5.1 and 5.2 are based on the discharge,

vegetation, grain size, and slope constraints inputted into them. More fundamentally,

however, these dimensions are based on maintaining the capacity to transport the

sediment supply at bank-full discharge. Since the capacity to transport a given supply of

sediment depends upon the channel slope and discharge, the predicted and observed

dimensions can be viewed as a balance between the erosional resistance of the channel

bed versus the erosional resistance of the channel banks. The erosional resistance of the

channel bed can be represented by the dimensionless shear stress (τ*), defined as

* h C

s 50

ρgR Sτ =

(ρ - ρ)gD 5.3

where ρ is the density of water (1000 kgm-3), ρs is the density of sediment (2650 kgm-3),

g is the gravitational acceleration (9.81 ms-2), Rh is the hydraulic radius of the channel

(m), SC is the channel slope (-), and D50 is the median grain size of the channel bed (m).

When relative bank strength values are plotted against the dimensionless shear stress a

strong positive correlation is observed where μ’ values of 1 or greater are associated with

Page 102: Downstream Variability of Fluvial Form, Process, and Character in a

91

dimensionless shear stresses above 0.03 (Fig. 5.17), which is often regarded as the lower

limit for sediment mobility (Buffington and Montgomery, 1997).

Figure 5.17 - Relationship between relative bank strength and dimensionless shear stress of the study sites.

The rationale behind this trend is that as the banks become increasingly resistant

to erosion due to increasing vegetation density, the channel is able to maintain high

dimensionless shear stresses without widening (Parker et al., 2007). Should the

vegetation be removed, the banks would become more erodible than the bed and the

channel would widen and shallow, lowering the erodibility (ie. dimensionless shear

stress) of the channel bed and reducing the relative bank strength to a value closer to 1.

This control that vegetation exerts on the channel dimensions explains why sites 2, 5, and

6 are deeper and narrower than predicted.

Bank strength values below 1, however, suggest that the channel banks are more

erodible than the channel bed and coincide with a dimensionless shear stress 67% less

than critical. As the dimensionless shear stresses are less than critical under the present

channel depths at sites 1 and 4u, the necessary channel depths to transport the bed

Page 103: Downstream Variability of Fluvial Form, Process, and Character in a

92

material given the present discharges and channel slopes must be much higher, and

consequently, the channels must be narrower. The calculated dimensions, however, have

not been attained, perhaps due to the dimensions not being physically stable enough, and

preferential erosion of the channel banks (despite being root-reinforced at Site 4u)

maintains wider and shallower channels than predicted, and ultimately results in a

mismatch between the transport capacities and the bed material at these two sites.

Based on the preceding analysis, it is suggested here that since the channel

dimensions at sites 2, 3, 5, and 6 can be well predicted, they are in equilibrium with the

prevailing hydrological and sediment conditions, while sites 1 and 4u are not. These two

sites are presently experiencing a mismatch between the sediment size along the channel

bed and the capacity to transport it, which stems from flatter slopes than necessary for

transport. Ferguson (1981) has classified this state as ‘persistent passive disequilibrium’.

Such channels have inherited their bed material and slopes from past channel states but

are incapable of modifying them under present hydrological conditions.

Using a classification scheme presented by Renwick (1992), persistent passive

disequilibrium may be comparable to a threshold-dominated non-equilibrium state,

whereby the channel is inactive until a threshold is crossed, rejuvenating the channel. In

the case of sites 1 and 4u, rejuvenation may be achieved through either a drastic increase

in discharge or an increase in channel slope. Site 4d may have undergone rejuvenation

through channel slope steepening, as the channel slope increases by 150% downstream of

Site 4u. This steepening is reflected in the dimensionless shear stress at Site 4d, which is

higher than upstream. However, the relative bank strength value is below 1, while the

Page 104: Downstream Variability of Fluvial Form, Process, and Character in a

93

vegetation conditions suggest it should be closer to sites 5 and 6. It is thought that Site 4d

is adjusting towards an equilibrium state but has not had enough time to reach it.

Page 105: Downstream Variability of Fluvial Form, Process, and Character in a

94

Chapter Six: Downstream Organization of Fluvial Form, Process, and Character 6.1 Spatial Distribution of Channel-Floodplain Types

Using the specific stream power values associated with each channel-floodplain

type, it is possible to broadly assess their spatial distributions within the watershed. A

map of the downstream variation in specific stream power (ω2) was constructed (R.

Phillips, personal communication) based on the relation

2 C2

bf

ρgQ Sω =

w 6.1

where ρ is the density of water (1000 kgm-3), g is gravitational acceleration (9.81 ms-2),

Q2 is the 2-year discharge (m3s-1), SC is the channel slope (-), and wbf is the bank-full

channel width (m). Bank-full width and discharge were estimated using regional rating

curves with drainage area (Ad, km2) given respectively as

0.50bf dw = 1.20A 6.2

0.912 dQ = 0.25A 6.3

Channel slope was derived by extracting the longitudinal profile from a 10-metre

digital elevation model and applying a moving window to smooth out artefacts within the

data. Drainage area was calculated through the D8 flow accumulation algorithm, which

measures the number of contributing upstream cells at any given point along the flow

path. Each cell represents 100 m2 of drainage area and the lower limit of the flow path

(and thus the start of the longitudinal profile) was 2 km2. While these values are not an

entirely accurate representation of bank-full specific stream power, the downstream

variation within the values are expected to be a reasonable reflection of the variation in

bank-full values.

Page 106: Downstream Variability of Fluvial Form, Process, and Character in a

95

The downstream variation in calculated specific stream power shows that the

highest values are located in the downstream end of the watershed while the upper half

possesses the lowest values (Fig. 6.1). From this, it is apparent that the upper half of the

watershed upstream of Site 4 is dominantly low energy and mixed energy channel-

floodplain types though the medium energy type may exist, while the lower half

downstream of Site 4 is dominated by the medium and high energy types. This pattern is

opposite to what is typically found in watersheds and, since channel slope largely drives

the spatial distribution of specific stream power, this is a reflection of the large increase

in slope along the downstream half of the longitudinal profile (Fig. 6.2).

The role that glacial landform inheritance and geomorphic history plays in

influencing the channel slope, and thus the distribution of fluvial form, is illustrated in

Figures 6.1 and 6.2. The low stream powers characterizing the upper half of the

watershed seem to largely stem from the prevalence of glaciolacustrine plains, which the

channel flows over. Low discharges, however, are largely responsible for maintaining

moderate to low stream power values where slopes are steeper, which are a consequence

of the channel passing over moraines or through ridges (possibly small moraines)

separating successive glaciolacustrine deposits. Conversely, high stream powers

downstream are driven by channel incision due to post-glacial isostatic rebound and

prehistoric lake level fluctuations in Lake Ontario (Anderson and Lewis, 2012).

6.2 Downstream Organization of Watershed Zones

The study sites were broadly classified as production, transport, and deposition

zones according to the qualitative criteria presented in Chapter Two (Fig. 2.5). To

measure channel confinement, a simple ratio of the width of the channel to the width of

Page 107: Downstream Variability of Fluvial Form, Process, and Character in a

96

Figure 6.1 - Downstream variation in specific stream power in relation to glacial landforms (courtesy of R. Phillips). Channel-floodplain types are listed with the study sites.

Page 108: Downstream Variability of Fluvial Form, Process, and Character in a

Figure 6.2 - Dominant glacial and post-glacial controls on the longitudinal profile. Selected well logs are given along the longitudinal profile to illustrate the downstream variability in glacial sediments and the relative degree of incision along the channel. Well logs were selected from the Ontario Ministry of Natural Resources Borehole Database Version 1.2 and supplemented with log descriptions of valley sides described by Karrow (1967) and Mahaney and Hancock (1993). The tops of the logs represent valley side topographic variations. Vertical exaggeration = 78x.

Page 109: Downstream Variability of Fluvial Form, Process, and Character in a

98

floodplain was used. Higher values mean that the channel is wide compared to the

floodplain. As a result, more opportunities for coupling with valley/terrace sides exist.

Overbank alluvium thickness and grain size values were taken from Table 5.1 and

floodplain ages were assigned based on radiocarbon dates presented in Chapter Five and

relative ages based on an analysis presented in Appendix F, section F.3.2. It should be

emphasized that the site classifications are subjective but do tend to conform to the broad

qualitative descriptions of each zone. For instance, low specific stream powers, sandy bed

material, and long term sediment storage on thick and old floodplains suggest that the

deposition zone is an appropriate classification for sites 1 and 3. Conversely, high

specific stream power coupled with high channel confinement, and a thin, poorly

developed floodplain suggest that Site 5 likely represents a production zone. The

remaining sites were classified as transport zone. These sites are more subjective since

they fall within the poorly defined area between the end member classifications.

However, their classification is based on low hillslope-channel coupling and moderate

specific stream powers. Channel beds are gravel to gravel-cobble indicating that sand is

carried through the system. Floodplain thickness is variable, indicative of variations in

upstream sediment sources.

The spatial distribution of the zones tends to suggest that both transport and

deposition zones exist in the upper watershed while the production zone exists in the

lower watershed (Fig. 6.3), which is the opposite of traditional watershed models. Since

high specific stream power and channel confinement are two dominant characteristics of

the production zone, there is no reason to suggest that the upper watershed has a

Page 110: Downstream Variability of Fluvial Form, Process, and Character in a

Figure 6.3 - Spatial distribution of inferred watershed zones for each study site. SSP = specific stream power; Con. Ratio = confinement ratio (bank-full channel width/floodplain width, both in metres); Bed Mat. = channel bed material composition; OB Thickness = overbank alluvium thickness; Fp Age = floodplain age. Vertical exaggeration = 78x.

Page 111: Downstream Variability of Fluvial Form, Process, and Character in a

100

production zone. If one does exist, it is likely very inefficient. Only along one stretch of

the channel length is the traditional production zone-transport zone sequence inferred and

that is the lowermost 8 kilometres of the river. Whether by coincidence or not, both of the

study sites (sites 5 and 6) within this stretch are suggested to be in equilibrium, while

many of the remaining sites are not. The inferred inverted and spatially variable

arrangement of watershed zonation likely explains the increasing or stable trends in

specific sediment yields with drainage area described in some southern Ontario

watersheds (Ashmore, 1993; Campo and Desloges, 1994; Church et al., 1999). The

following is proposed as a potential cause for these patterns. In the headwaters, due to

low stream power and little to no channel-hillslope coupling, sediment recruitment is

generally low. However, as the channel moves downstream and progressively becomes

more energetic and incised, the level of sediment recruitment increases. If the degree of

incision and channel slope steepening is faster than drainage area can increase

downstream, it is probable that specific sediment yield increases. Otherwise, a stable

trend might be observed indicating consistent downstream recruitment of sediment and/or

limited floodplain storage.

6.3 Downstream Trends in Channel and Floodplain Character

The inverted and spatially variable organization of the watershed zones raises the

question as to how channel and floodplain characteristics vary downstream and whether

they conform to expected downstream trends. Using data presented in Table 5.1, the

characteristics were plotted against downstream distance. Only the characteristics that

can be reliably compared against prior knowledge are reported in detail here. The

remaining characteristics are presented in Appendix E.

Page 112: Downstream Variability of Fluvial Form, Process, and Character in a

101

6.3.1 Channel Morphology

Standard hydraulic geometry relations for alluvial rivers suggest that bank-full

channel width, depth, and cross-sectional area increase downstream, with width

increasing marginally faster than depth (Leopold and Maddock, 1953). For the Little

Rouge River, bank-full channel width tends to show a strong increase downstream (Fig.

6.4). However, variation within this trend is evident, particularly between sites 2 and 3

where there is a downstream decrease in width, though this decrease is not significant (p

= 0.855). Depth, on the other hand, shows a lack of a downstream trend. Though a slight

deepening trend is detected, the relation is rather weak and the regression slope is not

significantly different from zero. Site 3, which is significantly deeper (p < 0.05) than the

other study sites, plots above the best-fit line. Channel adjustment at Site 3 appears to be

accommodated by a deepening of the channel rather than widening, which is in contrast

to the other study sites.

Figure 6.4 - Downstream variation in bank-full channel width, depth, and cross-sectional area. Error bars are given as one standard error. Study sites labelled for reference. Site 4 represents two data points (sites 4u and 4d).

Page 113: Downstream Variability of Fluvial Form, Process, and Character in a

102

The product of width and depth yields bank-full cross-sectional area, for which

the downstream trend is strongly positive. Variation exists within this relation with Site 6

showing a significantly larger (p < 0.05) channel area over a relatively small increase in

drainage area compared to Site 5. Sites 2, 3, and 6 all plot above the trend-line, which is

driven by large channel widths for sites 2 and 6, and a large channel depth for Site 3.

6.3.2 Channel Hydrology

It has conventionally been thought that bank-full discharge recurs at a periodicity

of 1-2 years on the annual maximum daily series (Wolman and Leopold, 1957). Since the

mean annual flood, which is the discharge approximated by the 2.33 year return period,

increases downstream in humid and temperate climates (Knighton, 1987), it logically

follows that bank-full discharge should increase downstream as well.

Overall, there is a moderate trend towards increasing morphological bank-full

discharge with distance downstream (Fig. 6.5). However, a 38% decrease in

morphological bank-full discharge exists between sites 2 and 3 despite a 237% increase

in drainage area. Since Figure 6.4 demonstrated that cross-sectional area increases in the

downstream direction, the trend in morphological bank-full discharge is necessarily

controlled by flow velocity, which was calculated by dividing the bank-full discharge by

the average cross-sectional area. Flow velocity shows an overall weak downstream

increase with the downstream decrease in bank-full discharge between sites 2 and 3

explained by a 61% decrease in bank-full velocity (Fig. 6.6).

6.3.3 Channel Bed Material Size

It is useful to view the downstream variation in channel bed material size as a

single measure, which is done using the sorting coefficient. Grain size sorting was

Page 114: Downstream Variability of Fluvial Form, Process, and Character in a

103

Figure 6.5 - Downstream variation in bank-full discharge.

Figure 6.6 - Downstream variation in bank-full velocity. calculated for each site based on the Inclusive Graphic Standard Deviation presented by

Folk and Ward (1957) and plotted on Figure 6.7a. In conventional grain size models,

whether through abrasion or size selective transport, the coarsest grain sizes decrease

Page 115: Downstream Variability of Fluvial Form, Process, and Character in a

104

Figure 6.7 - Downstream variation in grain size sorting coefficient (a) and various size fractions (b-g). Sites labelled for reference.

downstream quicker than the finest, which results in a downstream tightening of the grain

size distribution. Less dispersion within the distribution means that the grain sizes are

Page 116: Downstream Variability of Fluvial Form, Process, and Character in a

105

better sorted and one should expect to see a downstream decrease in the sorting

coefficient. Sorting coefficients along the Little Rouge River tend to exhibit this

downstream pattern where distributions evolve from very poorly sorted towards

moderately sorted (Folk and Ward, 1957), though considerable scatter is observed in the

upstream portion of the watershed and the slope of the regression is not significantly

different from zero at the 95% confidence interval. This trend is realized, however, not

from a decrease in the coarse fractions, which exhibit little downstream trend, but instead

a downstream increase in the fine fractions (Fig. 6.7b-g). Considerable scatter exists

within each fraction though, leading to poorly defined relationships and regression slopes

not significantly different from zero at the 95% confidence interval for any fraction. The

observed downstream trends in sorting and grain size fractions are a result of the presence

of sites 1 and 3, which consist of sandy-gravel bed material. Whether or not these trends

would persist with the inclusion of more sites in the upper portion of the watershed is not

certain.

6.3.4 Overbank Alluvium Thickness

The thickness of floodplain overbank alluvial deposits is typically thought of as

the boundary between gravel and sand. This discrimination may be true for higher energy

sites but is less true for sites where sand forms a component of the channel bed. In these

sites, the boundary between channel and overbank deposits is less obvious. As such,

overbank deposits may be over-estimated in some cases. Based on bank exposure S4-B2,

Site 4 may be over-estimated by up to 20%. At sites 1 and 6, where over-estimation

interferes with the accurate representation of the downstream trend, overbank estimates

are based on the inferred depositional environments shown for their respective sediment

Page 117: Downstream Variability of Fluvial Form, Process, and Character in a

106

logs. At Site 1, the boundary between channel and floodplain deposits could only be

distinguished in one sample. Accordingly, a value of 45 cm was assigned to the site so as

to not over-estimate the overbank thickness, which otherwise may have been over-

estimated by up to 100%.

Due to its complex geomorphic history, perhaps more so than the other sites, the

sub-2mm thickness at Site 6 ranges from 95 cm to over 380 cm. While the latter value is

believed to be more representative of the site based on auger holes, cores, and banks

analyzed, closer inspection of sediment cores taken from the floodplain reveal that much

of this sediment includes several fining upward trends thought to be indicative of

multiple episodes of channel and floodplain aggradation (Fig. 5.13). In order to be

comparable to the other sites for analysis, only the overbank deposits deposited under the

present energy conditions of the channel (ie. specific stream power) were included in the

analysis. The uppermost overbank sequence was chosen for this purpose, which was

estimated to be approximately 112 cm based off of S6-C2 (Fig. 5.12b) and a cut-face not

presented here.

Plotting overbank thickness against distance downstream reveals no overall

downstream trend and a large degree of scatter about the trendline (Fig. 6.8). The

floodplain tends to become thicker downstream, consistent with conventional thought,

peaking at Site 3. However, downstream of Site 3 there is an 86% reduction in thickness,

which is followed by a second peak at Site 6.

Page 118: Downstream Variability of Fluvial Form, Process, and Character in a

107

Figure 6.8 - Downstream variation in floodplain overbank alluvium thickness. Error bars represent one standard error. Error bars for sites 1 and 6 are not reported due to a lack of suitable data.

6.3.5 Summary and Controls

The downstream trends described above, including those in Appendix E, are

summarized in Table 6.1. While there appears to be some agreement with the expected

downstream trends, in many cases the agreement is rather tenuous as indicated by the r2

values of the trendlines and regression slopes are often not significantly different from

zero. This is a consequence of both the variation within the watershed and the emphasis

on sampling sites of contrasting characteristics instead of systemic downstream sampling.

Despite this latter point, these trends are believed to be representative of the watershed. In

more than half of the variables, the observed trend is opposite of what is expected.

Morphological and hydrological variables appear to most generally conform to expected

downstream trends while sedimentological variables tend to exhibit the most departure.

At a very broad scale, the downstream trends reflect the assertion that the watershed is

arranged in a spatially variable but generally inverted manner.

Page 119: Downstream Variability of Fluvial Form, Process, and Character in a

Characteristic Expected Trend (Source) Observed Trend (r2) Controls Channel Morphology Bank-full Width Increasing (Knighton, 1987) Increasing (0.91)** Discharge (+), channel slope (+), large woody debris (+),

boundary resistance (-) Bank-full Depth Increasing (Knighton, 1987) Increasing (0.05) Discharge (+), channel slope (-), large woody debris (+),

boundary resistance (-) Bank-full Cross-sectional Area Increasing (Knighton, 1987) Increasing (0.93)** Discharge (+), boundary resistance (-), large woody debris

(+) Bank-full Width-Depth Ratio Decreasing? (n.a.) Increasing (0.74)* Channel slope (+), discharge (+/-), bed material size (+/-),

bed-load sediment supply (+), boundary resistance (-) Channel Sinuosity Increasing? (n.a.) Decreasing (0.19) Channel slope (-), discharge (+/-), bed material size (-) Channel Hydrology Bank-full Discharge Increasing (Knighton, 1987) Increasing (0.81)** Not analyzed Bank-full Discharge Return

Period Increasing (Petit & Pauquet, 1997)

Decreasing (0.12) Channel slope (+), large woody debris (+), boundary resistance (-), hydrological regime (+/-)

Bank-full Velocity Increasing (Knighton, 1987) Increasing (0.35) Not analyzed Channel and Floodplain Sedimentology

Channel Bed Alluvium Thickness

Increasing? (n.a.) Increasing (0.28) Channel slope (-), discharge (-), sediment supply (+), geomorphic history (+/-), large woody debris (+)

Exposed Clay Percentage Decreasing (Montgomery & Buffington, 1997)

None (0.00) Channel slope (+), discharge (+), sediment supply (-), geomorphic history (+/-), large woody debris (-)

Bed Material Size Decreasing (Knighton, 1987) Increasing (0.02-0.37) Channel slope (+), glacial inheritance (+), large woody debris (-)

Sorting Coefficient Decreasing (Knighton, 1980) Decreasing (0.51) Channel slope (-), glacial inheritance (+/-), large woody debris (+)

Overbank Alluvium Thickness Increasing (Church, 2002) None (0.00) Channel slope (-), sediment supply (+), large woody debris (+)

Silt/Clay Percentage Increasing? (n.a.) Decreasing (0.35) Channel slope (-), large woody debris (+) Organic Content (Fp age) Increasing (Brown, 1990) Decreasing (0.86)** Channel slope (-), large woody debris (+), boundary

resistance (+)

Page 120: Downstream Variability of Fluvial Form, Process, and Character in a

109

Table 6.1 (previous page) - Observed downstream trends in various channel and floodplain characteristics compared against their expected trends. N.A. – No study was found suggesting any preferential downstream trend. Accordingly, a trend is indicated based on theoretical expectations. * – Regression slope significant at the 95% confidence interval. ** – Regression slope significant at the 99% confidence interval. Controls are listed with most important control first. + – An increase in the control variable results in an increase in the response variable. - – An increase in the control variable results in a decrease in the response variable. +/- – An increase or a qualitative change in the control variable results in either an increase or decrease in the response variable. See Appendix F for a full explanation of the controls.

The conventional controls on each of the characteristics in Table 6.1 were

examined in detail, the results of which are presented in Appendix F. To summarize that

analysis, channel slope acts as either a primary or secondary control for all of the

variables except bank-full cross-sectional area. The morphological and hydrological

variables tend to show the most conformity with expected downstream trends due to

discharge acting as a strong control, which increases downstream in watersheds of humid

and temperate climates regardless of how the watershed zones are organized. Increasing

channel slope along the lower half of the channel results in a reinforcement of channel

width and counteracts the positive effect that discharge has on depth. As a result, channel

width tends to show a strong downstream increase while channel depth shows no trend.

The effect of increasing channel slope downstream can be seen in the width-depth ratio

and channel sinuosity, which are strongly related to channel slope. Channel boundary

resistance and large woody debris play smaller roles in morphological and hydrological

control.

Sedimentological variables, on the other hand, show the most scatter and

departure from expected trends due to a suite of various co-dominant controls including

channel slope, upstream sediment supply, and geomorphic history, among others. The

complexity in the spatial arrangement of sediment sources, the degree to which the

channel is incised into them, and the variability of channel slope can easily be seen in

Figure 6.3. The effect of these various co-dominant controls is most pronounced in the

Page 121: Downstream Variability of Fluvial Form, Process, and Character in a

110

channel bed alluvium thickness and exposed glacial clay percentage. The remaining

variables are more dominantly controlled by channel slope, such that the channel bed

becomes coarser and the floodplain becomes thinner, coarser, and younger with

increasing channel slope, though scatter remains in some variables due to secondary

controls, such as upstream sediment supply, glacially inherited bed material, and large

woody debris.

Page 122: Downstream Variability of Fluvial Form, Process, and Character in a

111

Chapter Seven: Discussion and Conclusions 7.1 Reappraisal of the ‘Graded’ River Model for Non-Alpine Deglaciated

Watersheds

The traditional exponentially decreasing channel slope characteristic of graded

river channels is based on the idea that river channels originate in steep, mountainous

terrain and flow towards the ocean or some other spatially extensive base-level. This

rather generalized model is most applicable for large, high-relief watersheds where such

features often exist at the upper and lower ends of the watershed respectively, but not

applicable for many small, non-alpine, deglaciated watersheds where watershed relief is

often much lower and the channel slope is controlled by the spatial distribution of glacial

and geological landforms, bedrock outcroppings, and post-glacial processes, such that a

variety of non-graded longitudinal profile shapes are possible.

As the findings of this study reinforce the importance of channel slope in

controlling the spatial distribution of fluvial forms and the downstream trends of various

channel and floodplain characteristics, many of the results presented here may not be

easily transferable to watersheds within southern Ontario with different longitudinal

profiles than that of the Little Rouge River. Such watersheds likely exhibit their own

downstream trends in accordance with variations in channel slope, which may be further

confounded by watershed-specific variations in glacial sediment sources and other

controls. Analysis of other southern Ontario watersheds may reveal different downstream

trends and different channel and floodplain forms, but it is probable that the variability

within the watershed would be directly proportional to the complexity of the longitudinal

profile.

Page 123: Downstream Variability of Fluvial Form, Process, and Character in a

112

7.2 Summary of Findings

The existence of five channel-floodplain types are proposed for the Little Rouge

River watershed based on specific stream power and bed sediment calibre. Laterally

stable, low energy types tend to develop their floodplains through lateral accretion of

sandy and/or gravely bed-load accompanied by oblique accretion of sandy-silt deposits.

More dynamic, higher energy types are dominated by lateral accretion of sandy-gravel

point bars and vertical accretion of lateral and medial gravel bars, overlain by vertically

accreted sand. Their spatial distribution is broadly related to the distribution of specific

stream power, which is highest in the lower half of the watershed and lowest in the upper

half, counter to what is expected in conventional ‘graded’ rivers. Variations in specific

stream power are a result of high slopes downstream driven by base-level fluctuations

and low slopes upstream where the channel flows over glaciolacustrine deposits. The

spatial distribution of specific stream power and progressive downstream channel

entrenchment suggests that the watershed is organized in a spatially variable, but

generally inverted manner with the production zone near the lower end of the watershed

and the transport and deposition zones dominating the upper half of the watershed.

As a consequence of this inverted organization, downstream trends in various

channel and floodplain characteristics are found to be highly variable, but are more

conformable with expected trends for morphological and hydrological variables, whereas

sedimentological variables are inconsistent with conventional downstream trends.

Morphological and hydrological variables are strongly controlled by channel slope in

addition to discharge, though large woody debris and boundary resistance play smaller

roles. The large scatter observed for sedimentological variables is attributed to the highly

Page 124: Downstream Variability of Fluvial Form, Process, and Character in a

113

variable nature of glacial sediment sources and the variable degree to which the channel

is incised into these deposits, on top of the conventional relationships with channel slope.

Lastly, using a rational regime relation, it is suggested that the channel is in

various states of equilibria with the present hydrological regime, channel slope, and bed

material size. Four sites were found to be in equilibrium, while two are thought to be in a

state of persistent passive equilibrium where the bed material is too coarse for the channel

slope.

7.3 Future Research Considerations

Compiling data from additional study sites, with particular emphasis on the upper

half of the watershed, would allow for increased interpretative power for many of the

relationships and inferences made here. Furthermore, the longitudinal profile of the Little

Rouge River exhibits three incidences of the conventional graded profile separated by

two prominent inflection points. By increasing the number of data points, downstream

trends within each sub-profile can be assessed in addition to the full profile, which would

allow insight into how the various characteristics respond under graded conditions in

deglaciated watersheds. By adding additional sites, the range of, and variability within,

the channel-floodplain types may be broadened, which would allow for relative

abundances of each type to be calculated for the watershed to see if any particular types

are dominant. Examination of other southern Ontario watersheds may broaden the

inventory of fluvial styles in deglaciated watersheds.

Reconstruction of channel response to environmental change remains poorly

detailed for southern Ontario watersheds, particularly for smaller watersheds like the

Little Rouge. Reconstructing the chronologies of the Little Rouge River in both the

Page 125: Downstream Variability of Fluvial Form, Process, and Character in a

114

downstream and upstream ends of the watershed would be useful considering the variety

of depositional environments observed.

As has been described here previously, downstream fining trends in the bed

material are often thought to be disrupted by lateral inputs from glacial sediment sources.

However, the contribution of tributary inputs relative to glacial inputs in this disruption

remains unresolved for non-alpine, deglaciated watersheds. A further research

consideration might investigate how the grain size distribution of the bed material

changes downstream of confluences under various slope, drainage area, and bed material

conditions, and to what extent glacial valley side inputs alter these trends.

Page 126: Downstream Variability of Fluvial Form, Process, and Character in a

115

References

Adamowski, K., Liang, G.-C., and Patry, G.G. 1998. Annual maxima and partial duration

flood series analysis by parametric and non-parametric methods. Hydrological

Processes, 12, 1685-1699.

Allmendinger, N.E., Pizzuto, J.E., Potter, N., Johnson, T.E., and Hession, W.C. 2005.

The influence of riparian vegetation on stream width, eastern Pennsylvania, USA.

Geological Society of America Bulletin, 117, doi: 10.1130/B25447.1.

Anderson, T.W. and Lewis, C.F.M. 2012. A new water-level history for Lake Ontario

basin: evidence for a climate-driven early Holocene lowstand. Journal of

Paleolimnology, 47, doi: 10.1007/s10933-011-9551-8.

Andrews, E.D. 1984. Bed-material entrainment and hydraulic geometry of gravel-bed

rivers in Colorado. Geological Society of America Bulletin, 95, 371-378.

Annable, W.K. 1996. Database of Morphologic Characteristics of Watercourses in

Southern Ontario. Ontario Ministry of Natural Resources.

Annable, W.K., Lounder, V.G., and Watson, C.C. 2010. Estimating channel-forming

discharge in urban watercourses. River Research and Applications, 27, doi:

10.1002/rra.1391.

Arbogast, A.F., Bookout, J.R., Schrotenboer, B.R., Lansdale, A., Rust, G.L., and Bato,

V.A. 2008. Post-glacial fluvial response and landform development in the upper

Muskegon River valley in North-central Lower Michigan, U.S.A.

Geomorphology, 102, doi: 10.1016/j.geomorph.2008.06.008.

Ashmore, P. 1993. Contemporary erosion of the Canadian landscape. Progress in

Physical Geography, 17, 190-204.

Page 127: Downstream Variability of Fluvial Form, Process, and Character in a

116

Barnett, P.J., Sharpe, D.R., Russell, H.A.J., Brennand, T.A., Gorrell, G., Kenny, F., and

Pugin, A. 1998. On the origin of the Oak Ridges Moraine. Canadian Journal of

Earth Science, 35, 1152-1167.

Bettess, R. and White, W.R. 1987. Extremal hypotheses applied to river regime. In:

Thorne, C.R., Bathurst, J.C., and Hey, R.D. (Eds.). Sediment Transport in Gravel-

bed Rivers. John Wiley & Sons Ltd: Chichester. 767-789.

Bluck, B.J. 1971. Sedimentation in the meandering River Endrick. Scottish Journal of

Geology, 7, doi: 10.1144/sjg07020093.

Bray, D.I. 1979. Estimating average velocity in gravel-bed rivers. Journal of the

Hydraulics Division, 105, 1103-1122.

Brewer, P., Johnstone, E., and Macklin, M. 2009. River dynamics and environmental

change in Wales. In: Williams, D.D. and Duigan, C.A. (Eds.). The Rivers of

Wales. Backhuys Publishers: Leiden. 17-34.

Brierley, G.J. and Hickin, E.J. 1985. The downstream gradation of particle sizes in the

Squamish River, British Columbia. Earth Surface Processes and Landforms, 10,

597-606.

Brierley, G.J. 1989. River planform facies models: the sedimentology of braided,

wandering, and meandering reaches of the Squamish River, British Columbia.

Sedimentary Geology, 61, 17-35.

Brierley, G.J. and Hickin, E.J. 1991. Channel planform as a non-controlling factor in

fluvial sedimentology: a case of the Squamish River floodplain, British Columbia.

Sedimentary Geology, 75, 67-83.

Page 128: Downstream Variability of Fluvial Form, Process, and Character in a

117

Brown, A.G. 1990. Holocene floodplain diachronism and inherited downstream

variations in fluvial processes: a study of the river Perry, Shropshire, England.

Journal of Quaternary Science, 5, 39-51.

Buffington, J.M. and Montgomery, D.R. 1997. A systematic analysis of eight decades of

incipient motion studies, with special reference to gravel-bedded rivers. Water

Resources Research, 33, 1993-2029.

Campo, S.H. and Desloges, J.R. 1994. Sediment yield conditioned by glaciation in a rural

agricultural basin of southern Ontario, Canada. Physical Geography, 15, 495-515.

Carson, M.A. 1984a. Observations on the meandering-braided river transition, the

Canterbury Plains, New Zealand: Part Two. New Zealand Geographer, 40, 89-99.

Carson, M.A. 1984b. The meandering-braided river channel threshold: a reappraisal.

Journal of Hydrology, 73, 315-334.

Carson, M.A. 1986. Characteristics of high-energy “meandering” rivers: The Canterbury

Plains, New Zealand. Geological Society of America Bulletin, 97, doi:

10.1130/0016-7606(1986)97<886:COHMRT>2.0.CO;2d.

Charlton, R. 2008. Fundamentals of Fluvial Geomorphology. Routledge: London. 234

pp.

Chew, L.C. and Ashmore, P.E. 2001. Channel adjustment and a test of rational regime

theory in a proglacial braided stream. Geomorphology, 37, 43-63.

Church, M. 2002. Geomorphic thresholds in riverine landscapes. Freshwater Biology, 47,

541-557.

Page 129: Downstream Variability of Fluvial Form, Process, and Character in a

118

Church, M. 2006. Bed material transport and the morphology of alluvial river channels.

Annual Review of Earth and Planetary Science, 34, doi:

10.1146/annurev.earth.33.092203.122721.

Church, M. and Rood, K. 1983. Catalogue of Alluvial River Channel Regime Data.

Department of Geography, University of British Columbia, Vancouver.

Church, M., Ham, D., Hassan, M., and Slaymaker, O. 1999. Fluvial clastic sediment yield

in Canada: scaled analysis. Canadian Journal of Earth Science, 36, 1267-1280.

Church, M., Kellerhals, R., and Day, T.J. 1989. Regional clastic sediment yield in British

Columbia. Canadian Journal of Earth Science, 26, 31-45.

Church, M.A., McLean, D.G., and Wolcott, J.F. 1987. River bed gravels: sampling and

analysis. In: Thorne, C.R., Bathurst, J.C., and Hey, R.D. (Eds.). Sediment

Transport in Gravel-bed Rivers. John Wiley & Sons Ltd: Chichester. 43-88.

Church, M. and Slaymaker, O. 1989. Disequilibrium of Holocene sediment yield in

glaciated British Columbia. Nature, 337, 452-454.

Croke, J.C. 1994. Floodplain change in the Glenmalure Valley, South-east Leinster. Irish

Geography, 27, 122-134.

Dade, W.B. 2000. Grain size, sediment transport, and alluvial channel pattern.

Geomorphology, 35, 119-126.

Davidson, S.K. and Hey, R.D. 2011. Regime equations for natural meandering cobble-

and gravel-bed rivers. Journal of Hydraulic Engineering, 137, doi:

10.1061/(ASCE)HY.1943-7900.0000408.

Davies, T.R.H. and Sutherland, A.J. 1983. Extremal hypotheses for river behavior. Water

Resources Research, 19, 141-148.

Page 130: Downstream Variability of Fluvial Form, Process, and Character in a

119

Dean, W.E. 1974. Determination of carbonate and organic matter in calcareous sediments

and sedimentary rocks by loss on ignition: comparison with other methods.

Journal of Sedimentary Petrology, 44, 242-248.

Desloges, J.R. and Church, M. 1987. Channel and floodplain facies in a wandering

gravel-bed river. In: Ethridge, F.G., Flores, R.M., and Harvey, M.D. (Eds.).

Recent Developments in Fluvial Sedimentology. Society of Economic

Paleontologists and Mineralogists, 99-110.

Dury, G.H., Sinker, C.A., and Pannett, D.J. 1972. Climatic change and arrested meander

development on the River Severn. Area, 4, 81-85.

Eaton, B.C. and Church, M. 2004. A graded stream response relation for bed load-

dominated streams. Journal of Geophysical Research, 109, F03011, doi:

10.1029/2003JF000062.

Eaton, B.C. and Church, M. 2007. Predicting downstream hydraulic geometry: a test of

rational regime theory. Journal of Geophysical Research, 112, F03025, doi:

10.1029/2006JF000734.

Eaton, B.C. and Church, M. 2009. Channel stability in bed load-dominated streams with

nonerodible banks: inferences from experiments in a sinuous flume. Journal of

Geophysical Research, 114, F01024, doi: 10.1029/2007JF000902.

Eaton, B.C., Church, M., and Millar, R.G. 2004. Rational regime model of alluvial

channel morphology and response. Earth Surface Processes and Landforms, 29,

doi: 10.1002/esp.1062.

Eaton, B.C. and Millar, R.G. 2004. Optimal alluvial channel width under a bank stability

constraint. Geomorphology, 62, doi: 10.1016/j.geomorph.2004.02.003.

Page 131: Downstream Variability of Fluvial Form, Process, and Character in a

120

Eaton, B.C., Millar, R.G., and Davidson, S. 2010. Channel patterns: braided,

anabranching, and single-thread. Geomorphology, 120, doi:

10.1016/j.geomorph.2010.04.101.

Eyles, N. and Boyce, J.I. 1991. Earth Science Survey of the Rouge Valley Park. Ontario

Ministry of Natural Resources, Central Region: Aurora, Ontario. Open File

Geological Report 9113. 82 pp.

Ferguson, R.I. 1971. Channel pattern and sediment type. Area, 5, 38-41.

Ferguson, R.I. 1981. Channel forms and channel changes. In: Lewin, J. (Ed.). British

Rivers. George Allen & Unwin Ltd: London. 90-125.

Ferguson, R.I. 1986. Hydraulics and hydraulic geometry. Progress in Physical

Geography, 10, 1-31.

Ferguson, R. 1987. Hydraulic and sedimentary controls of channel pattern. In: Richards,

K. (Ed.). River Channels: Environment and Process. Basil Blackwell Ltd:

Oxford. 129-158.

Ferguson, R. and Ashworth, P. 1991. Slope-induced changes in channel character along a

gravel-bed stream: the Allt Dubhaig, Scotland. Earth Surface Processes and

Landforms, 16, 65-82.

Ferguson, R., Hoey, T., Wathen, S., and Werritty, A. 1996. Field evidence for rapid

downstream fining of river gravels through selective transport. Geology, 24, 179-

182.

Folk, R.L. and Ward, W.C. 1957. Brazos River bar: a study in the significance of grain

size parameters. Journal of Sedimentary Petrology, 27, 3-26.

Page 132: Downstream Variability of Fluvial Form, Process, and Character in a

121

Forbes, D.L. 1983. Morphology and sedimentology of a sinuous gravel-bed channel

system: lower Babbage River, Yukon coastal plain, Canada. In: Collinson, J.D.

and Lewin, J. (Eds.). Modern and Ancient Fluvial Systems. International

Association of Sedimentologists: Oxford. 195-206

Foster, G.E. 1998. Morphology of three semi-alluvial stream channels in southern

Ontario. Unpublished Masters Thesis. The University of Western Ontario. 170

pp.

Glaves, R. and Waylen, P.R. 1997. Regional flood frequency analysis in southern Ontario

using L-moments. The Canadian Geographer, 41, 178-193.

Griffiths, G.A. 1981. Stable-channel design in gravel-bed rivers. Journals of Hydrology,

52, 291-305.

Griffiths, G.A. 1984. Extremal hypotheses for river regime: an illusion of progress. Water

Resources Research, 20, 113-118.

Gurnell, A.M. 1998. The hydrogeomorphological effects of beaver dam-building activity.

Progress in Physical Geography, 22, doi: 10.1177/030913339802200202.

Henderson, F.M. 1963. Stability of alluvial channels. Transactions of the American

Society of Civil Engineers, 128, 657-720.

Hession, W.C., Pizzuto, J.E., Johnson, T.E., and Horwitz, R.J. 2003. Influence of bank

vegetation on channel morphology in rural and urban watersheds. Geology, 31,

147-150.

Hey, R.D. 1979. Flow resistance in gravel-bed rivers. Journal of the Hydraulics Division,

105, 365-379.

Page 133: Downstream Variability of Fluvial Form, Process, and Character in a

122

Hey, R.D. and Thorne, C.R. 1986. Stable channels with mobile gravel beds. Journal of

Hydraulic Engineering, 112, 671-689.

Hickin, E.J. 1983. River channel changes: retrospect and prospect. In: Collinson, J.D. and

Lewin, J. (Eds.). Modern and Ancient Fluvial Systems. Blackwell Publishing Ltd:

Oxford. 61-83.

Jackson, R.G. 1976. Depositional model of point bars in the lower Wabush River.

Journal of Sedimentary Petrology, 46, 579-594.

Jackson, R.G. 1981. Sedimentology of muddy fine-grained channel deposits in

meandering streams of the American middle west. Journal of Sedimentary

Petrology, 51, 1169-1192.

Javelle, P., Ouarda, T.B.M.J., and Bobeé, B. 2003. Spring flood analysis using the flood-

duration-frequency approach: application to the provinces of Quebec and Ontario,

Canada. Hydrological Processes, 17, doi: 10.1002/hyp.1349.

Jeffries, R., Darby, S.E., and Sear, D.A. 2003. The influence of vegetation and organic

debris on flood-plain sediment dynamics: case study of a low-order stream in the

New Forest, England. Geomorphology, 51, 61-80.

Jiongxin, X. and Yunxia, Y. 2005. Scale effects on specific sediment yield in the Yellow

River basin and geomorphological explanations. Journal of Hydrology, 307, doi:

10.1016/j.jhydrol.2004.10.011.

Karrow, P.F. 1967. Pleistocene Geology of the Scarborough Area. Geological Report 46.

Ontario Department of Mines. 108 pp.

Keller, E.A. and Swanson, F.J. 1979. Effects of large organic material on channel form

and fluvial processes. Earth Surface Processes, 4, 361-380.

Page 134: Downstream Variability of Fluvial Form, Process, and Character in a

123

Kidson, R. and Richards, K.S. 2005. Flood frequency analysis: assumptions and

alternatives. Progress in Physical Geography, 29, 392-410.

Kleinhans, M.G. and van den Berg, J.H. 2011. River channel and bar patterns explained

and predicted by an empirical and a physics-based method. Earth Surface

Processes and Landforms, 36, doi: 10.1002/esp.2090.

Knighton, A.D. 1980. Longitudinal changes in size and sorting of stream-bed material in

four English rivers. Geological Society of America, 91, 55-62.

Knighton, A.D. 1987. River channel adjustment – the downstream dimension. In:

Richards, K. (Ed.). River Channels: Environment and Process. Basil Blackwell

Ltd: Oxford. 95-128.

Knighton, D. 1998. Fluvial Forms and Processes: A New Perspective. Arnold: London.

383 pp.

Langbein, W.B. 1949. Annual floods and the partial-duration flood series. Transactions

of the American Geophysical Union, 30, 879-881.

Langbein, W.B. and Leopold, L.B. 1964. Quasi-equilibrium states in channel

morphology. American Journal of Science, 262, 782-794.

Lee, J.-S. and Julien, P.Y. 2006. Downstream hydraulic geometry of alluvial channels.

Journal of Hydraulic Engineering, 132, doi: 10.1061/(ASCE)0733-

9429(2006)132:12(1347).

Leopold, L.B. and Bull, W.B. 1979. Base level, aggradation, and grade. Proceedings of

the American Philosophical Society, 123, 168-202.

Page 135: Downstream Variability of Fluvial Form, Process, and Character in a

124

Leopold, L.B. and Maddock Jr., T. 1953. The hydraulic geometry of stream channels and

some physiographic implications. United States Geological Survey Professional

Paper 252.

Leopold, L.B. and Wolman, M.G. 1957. River channel patterns: braided, meandering,

and straight. United States Geological Survey Professional Paper 282-B.

Lewin, J. 1976. Initiation of bed forms and meanders in coarse-grained sediment.

Geological Society of America Bulletin, 87, doi: 10.1130/0016-

7606(1976)87<281.IOBFAM>2.0.CO;2.

Mackin, J.H. 1948. Concept of the graded river. Geological Society of America Bulletin,

59, 463-512.

Mahaney, W.C. and Hancock, R.G.V. 1993. Late Quaternary stratigraphy and

geochemistry of the R47 section, Rouge River basin, south-central Ontario,

Canada: correlation with Scarborough Bluffs. Journal of Quaternary Science, 8,

167-178.

Mahaney, W.C. and Sanmugadas, K. 1986. Soil development as a function of time in the

Rouge River basin, south-central Ontario. Géographie physique et Quaternaire,

40, 207-216.

Mahaney, W.C., Terasmae, J., Melville, H., and Boyer, M.G. 1989. Very late Holocene

overbank sedimentation and 14C ages of buried soils in the Rouge River basin,

south-central Ontario. Acta Geologica Hungarica, 32, 325-348.

Massong, T.M. and Montgomery, D.R. 2000. Influence of sediment supply, lithology,

and wood debris on the distribution of bedrock and alluvial channels. Geological

Society of America Bulletin, 112, 591-599.

Page 136: Downstream Variability of Fluvial Form, Process, and Character in a

125

Miall, A.D. 1996. The Geology of Fluvial Deposits: Sedimentary Facies, Basin Analysis,

and Petroleum Geology. Springer: Berlin. 582 pp.

Millar, R.G. 2000. Influence of bank vegetation on alluvial channel patterns. Water

Resources Research, 36, 1109-1118.

Millar, R.G. 2005. Theoretical regime equations for mobile gravel-bed rivers with stable

banks. Geomorphology, 64, doi: 10.1016/j.geomorph.2004.07.001.

Montgomery, D.R. and Buffington, J.M. 1997. Channel-reach morphology in mountain

drainage basins. Geological Society of America Bulletin, 109, 596-611.

Nanson, G.C. and Croke, J.C. 1992. A genetic classification of floodplains.

Geomorphology, 4, 459-486.

Nanson, G.C. and Young, R.W. 1981. Downstream reduction of rural channel size with

contrasting urban effects in small coastal streams of southeastern Australia.

Journal of Hydrology, 52, 239-255.

Navratil, O., Albert, M-B., Hérouin, E., and Gresillon, J-M. 2006. Determination of

bankfull discharge magnitude and frequency: comparison of methods on 16

gravel-bed river reaches. Earth Surface Processes and Landforms, 31, doi: 10.

1002/esp.1337.

Page, K.J., Nanson, G.C., and Frazier, P.S. 2003. Floodplain formation and sediment

stratigraphy resulting from oblique accretion on the Murrumbidgee River,

Australia. Journal of Sedimentary Research, 73, 5-14.

Park, C.C. 1977. World-wide variations in hydraulic geometry exponents of stream

channels: an analysis and some observations. Journal of Hydrology, 33, 133-146.

Page 137: Downstream Variability of Fluvial Form, Process, and Character in a

126

Parker, G. 1979. Hydraulic geometry of active gravel rivers. Journal of the Hydraulics

Division, 105, 1185-1201.

Parker, G., Wilcock, P.R., Paola, C., Dietrich, W.E., and Pitlick, J. 2007. Physical basis

for quasi-universal relations describing bankfull hydraulic geometry of single-

thread gravel bed rivers. Journal of Geophysical Research, 112, F04005, doi:

10.1029/2006JF000549.

Petit, F. and Pauquet, A. 1997. Bankfull discharge recurrence interval in gravel-bed

rivers. Earth Surface Processes and Landforms, 22, 685-693.

Phillips, R.T.J. and Robert, A. 2005. River incision in relation to post-glacial events in

the Humber River basin, Ontario. Géographie physique et Quaternaire, 59, 17-30.

Phillips, R.T.J. and Robert, A. 2007. Hydrologic control of waveforms on small

meandering rivers. Earth Surface Processes and Landforms, 32, doi:

10.1002/esp.1483.

Reinfelds, I., Cohen, T., Batten, P., and Brierley, G. 2004. Assessment of downstream

trends in channel gradient, total, and specific stream power: a GIS approach.

Geomorphology, 60, doi: 10.1016/j.geomorph.2003.10.003.

Reinfelds, I. and Nanson, G. 1993. Formation of braided river floodplains, Waimakariri

River, New Zealand. Sedimentology, 40, 1113-1127.

Renwick, W.H. 1992. Equilibrium, disequilibrium, and nonequilibrium landforms in the

landscape. Geomorphology, 5, 265-276.

Rice, S. and Church, M. 1996. Sampling surficial fluvial gravels: the precision of size

distribution percentile estimates. Journal of Sedimentary Research, 66, 654-665.

Page 138: Downstream Variability of Fluvial Form, Process, and Character in a

127

Sambrook Smith, G.H. and Ferguson, R.I. 1995. The gravel-sand transition along river

channels. Journal of Sedimentary Research, A65, 423-430.

Schumm, S.A. 1963. Sinuosity of alluvial rivers on the Great Plains. Geological Society

of America, 74, 1089-1100.

Schumm, S.A. 1993. River response to baselevel change: implications for sequence

stratigraphy. The Journal of Geology, 101, 279-294.

Schumm, S.A. and Khan, H.R. 1972. Experimental study of channel patterns. Geological

Society of America Bulletin, 83, 1755-1770.

Shugar, D., Kostaschuk, R., Ashmore, P., Desloges, J., and Burge, L. 2007. In situ jet-

testing of the erosional resistance of cohesive streambeds. Canadian Journal of

Civil Engineering, 34, doi: 10.1139/L07-024.

Smith, N.D. and Smith, D.G. 1984. William River: an outstanding example of channel

widening and braiding caused by bed-load addition. Geology, 12, 78-82.

Toronto and Region Conservation Authority. 2007a. Study Area and Physical Setting.

Rouge River State of the Watershed Report. 15 pp.

Toronto and Region Conservation Authority. 2007b. Surface Water Quantity. Rouge

River State of the Watershed Report. 33 pp.

Thorn, C.E. and Welford, M.R. 1994. The equilibrium concept in geomorphology.

Annals of the Association of American Geographers, 84, 666-696.

Weninger, J.M. and McAndrews, J.H. 1989. Late Holocene aggradation in the lower

Humber River valley, Toronto, Ontario. Canadian Journal of Earth Science, 26,

1842-1849.

Page 139: Downstream Variability of Fluvial Form, Process, and Character in a

128

Williams, G.P. 1978. Bank-full discharge of rivers. Water Resources Research, 14, 1141-

1154.

Wolcott, J. and Church, M. 1990. Strategies for sampling spatially heterogeneous

phenomena: the example of river gravels. Journal of Sedimentary Petrology, 61,

534-543.

Wolman, M.G. 1954. A method of sampling coarse river-bed material. Transactions of

the American Geophysical Union, 35, 951-956.

Wohl, E. and Achyuthan, H. 2002. Substrate influences on incised-channel morphology.

The Journal of Geology, 110, 115-120.

Wolman, M.G. and Leopold, L.B. 1957. River floodplains: some observations on their

formation. United States Geological Survey Professional Paper 282-C.

Wong, M. and Parker, G. 2006. Reanalysis and correction of bed-load relation of Meyer-

Peter and Müller using their own database. Journal of Hydraulic Engineering,

132, doi: 10.1061/(ASCE)0733-9429(2006)132:11(1159).

Page 140: Downstream Variability of Fluvial Form, Process, and Character in a

129

Appendix A: Channel Morphology and Sedimentology

A.1 Bank-full Channel Width

Site Cross-Section 1 2 3 4u 4d 5 6

1 4.77 10.46 10.60 13.20 13.10 15.36 18.48 2 5.67 8.93 9.00 13.20 14.75 11.60 19.26 3 6.60 6.70 8.90 14.06 14.70 14.97 24.41 4 7.60 10.06 7.94 13.04 10.97 13.40 19.66 5 4.39 13.00 7.25 14.70 28.87 12.41 6 4.83 10.35 10.06 15.65 22.52 7 11.06 13.26 15.89 26.80 8 6.20 5.90 16.04 9 8.60 15.00 10 8.80 27.57 11 17.60

Average 5.64 9.42 9.11 13.38 13.64 17.45 20.51 Standard Deviation 1.24 2.03 2.25 0.46 1.65 5.55 4.67

Table A.1 - Bank-full channel width measurements in metres for each site.

A.2 Bank-full Channel Depth

Site Cross-Section 1 2 3 4u 4d 5 6

1 0.78 0.55 1.43 1.02 0.85 0.36 1.2 2 0.71 0.54 1.22 0.45 0.78 0.86 0.86 3 0.66 0.63 1.00 0.67 0.89 0.72 0.56 4 0.70 0.66 1.03 0.79 0.46 0.82 0.97 5 0.51 0.70 0.63 0.66 0.61 0.74 6 0.72 0.64 1.26 0.62 0.91 7 0.71 1.00 0.65 0.44 8 0.85 0.73 0.54 9 0.56 0.97 10 0.64 0.78 11 0.77

Average 0.68 0.65 1.04 0.73 0.73 0.70 0.81 Standard Deviation 0.09 0.09 0.27 0.24 0.18 0.17 0.26

Table A.2 - Bank-full channel depth measurements in metres for each site.

Page 141: Downstream Variability of Fluvial Form, Process, and Character in a

130

A.3 Channel Bed Alluvium Thickness

Site 1 2 3* 4u 4d 5 6*

# of Points 112 259 120 230 91 370 185 Total

Thickness 4238.66 11596.51 7017.7 8421.57 2378.95 18571.56 16820.45

Average 37.85 44.77 58.48 36.62 26.14 50.19 90.92 Standard Deviation

24.31 15.97 32.88 18.51 21.22 24.42 36.29

Maximum Value

122.00 85.00 143.30 84.30 80.40 104.97 164.60

Minimum Value

1.30 1.98 7.20 0.20 0.65 1.05 1.31

Table A.3 - Alluvium thickness measurements in centimetres for each site. 'Total thickness' refers to the sum of all points for all cross-sections at a site. Sites marked with an asterisk are known to possess large under-estimates.

A.4 Channel Bed Gravel Thickness

Site 1 2 3* 4u 4d 5 6

# of Points 112 259 120 230 91 370 155 Total

Thickness 3434.86 11596.51 7017.7 8421.57 2378.95 18571.56 5717.35

Average 30.67 44.77 58.48 36.62 26.14 50.19 36.89 Standard Deviation

21.97 15.97 32.88 18.51 21.22 24.68 22.63

Maximum Value

94.00 85.00 143.30 84.30 80.40 104.97 86.70

Minimum Value

1.30 1.98 7.20 0.20 0.65 1.05 1.18

Table A.4 - Gravel thickness measurements in centimetres for each site. Sites marked with an asterisk are known to possess large under-estimates.

A.5 Miscellaneous Alluvium Strata Thickness

Site (Sediment Unit) 1 (Organic Sand) 3 (Sandy Gravel) 3 (Gravel)* 6 (Sand)*

Number of Points 73 116 32 185 Total Thickness 803.80 5644.40 1268.40 11103.1

Average 11.01 48.66 39.64 60.02 Standard Deviation 11.48 25.13 5.47 37.00 Maximum Value 59.20 98.60 44.70 164.60 Minimum Value 0.20 1.00 33.90 1.31

Table A.5 - Alluvium thickness for strata that appear above or below the channel gravels, or channel gravels that can be differentiated. Sites marked with an asterisk are known to possess large under-estimates.

Page 142: Downstream Variability of Fluvial Form, Process, and Character in a

131

A.6 Percent Glacial Clay in Bank-full Channel

Site Cross-Section 1 2 3 4u 4d 5 6

1 37.74% U

0.00% U

0.00% U

3.79% P

15.27% P

0.00% R

0.00% R

2 0.00% U

0.00% U

0.00% U

0.00% R

0.00% R

42.59% P

0.00% P

3 15.15% U

0.00% U

41.57% U

0.00% R

53.74% P

5.08% R

0.00% R

4 0.00% U

0.00% U

0.00% U

0.00% P

20.05% R

26.64% P

0.00% P

5 0.00% U

0.00% U

0.00% U

0.00% R

40.82% P

0.00% R

0.00% R

6 37.27% U

0.00% U

0.00% U

40.52% P

29.84% P

7 0.00% U

0.00% U

9.44% R

0.00% R

8 41.95% P

0.00% U

24.28% P

9 0.00% R

6.67% R

10 0.00% P

7.98% P

11 13.07% R

Average 15.03% 4.19% 5.20% 0.76% 25.98% 16.02% 4.26% Standard Deviation

18.37% 13.26% 14.70% 1.69% 21.31% 15.25% 11.28%

Table A.6 - Percentage of the bank-full channel that is occupied by glacial clay/bedrock.

R = riffle, P = pool, and U = undifferentiated bed-form.

A.7 Channel Bed Material Grain Size Characteristics Site

1 2 3 4u 4d 5 6 Dmax 1110 940 340 960 1260 2500 670 D95 55.72 157.59 36.00 142.02 132.51 187.4 61.82 D84 29.04 97.01 19.84 102.54 93.70 107.63 48.84 D50 7.46 35.51 3.36 61.82 56.10 53.82 30.91 D16 0.32 8.00 0.30 30.06 32.90 24.25 17.15 D5 0.07 4.76 0.10 14.83 21.11 12.13 8.00

Table A.7 - Grain size characteristics of the bed material. Values are in millimetres.

Page 143: Downstream Variability of Fluvial Form, Process, and Character in a

132

A.8 Grain Size Distribution of the Channel Bed Material

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

0.01 0.1 1 10 100

Grain Size (mm)

Per

cen

t F

iner

Composite

Sample 1

Sample 2

Sample 3

Figure A.1- Grain size distribution for Site 1 showing individual sample distributions and the composite distribution.

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

1 10 100 1000 10000

Grain Size (mm)

Per

cen

t F

iner

Composite

Sample 1

Sample 2

Sample 3

Sample 4

Sample 5

Figure A.2 - Grain size distribution for Site 2 showing individual sample distributions and the composite distribution.

Page 144: Downstream Variability of Fluvial Form, Process, and Character in a

133

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

0.01 0.1 1 10 100

Grain Size (mm)

Per

cen

t F

iner

Composite

Sample 1

Sample 2

Sample 3

Sample 4

Figure A.3 - Grain size distribution for Site 3 showing individual sample distributions and the composite distribution.

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

1 10 100 1000

Grain Size (mm)

Per

cen

t F

iner

Downstream

Upstream

Sample 1

Sample 2

Sample 3

Sample 4

Sample 5

Figure A.4 - Grain size distribution for Site 4d and 4u showing individual sample distributions and the composite distribution.

Page 145: Downstream Variability of Fluvial Form, Process, and Character in a

134

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

1 10 100 1000 10000

Grain Size (mm)

Per

cen

t F

iner Composite

Sample 1

Sample 2

Sample 3

Sample 4

Sample 5

Sample 6

Figure A.5 - Grain size distribution for Site 5 showing individual sample distributions and the composite distributions.

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

1 10 100 1000

Grain Size (mm)

Per

cen

t F

iner

Composite

Sample 1

Sample 2

Sample 3

Sample 4

Figure A.6 - Grain size distribution for Site 6 showing individual sample distributions and the composite distribution.

Page 146: Downstream Variability of Fluvial Form, Process, and Character in a

135

Appendix B: Floodplain Sedimentology

B.1 Floodplain Overbank Alluvium Thickness

Site Pit/Auger Hole 1 2 3 4u 4d 5 6

1 103* 100 111 54 51 33 95 2 94* 51 157 82* 31 36 292* 3 97* 72 165 41 68 19 4 92* 98 41 81 25 5 120* 97 13 6 88* 8 7 9 8 5 9 10

Bank Section 1 72 145 97* 58 44 262*† 2 94 22 255*†

Core 1 372*† 2 372*†

Average 99.00 81.67 144.50 63.00 64.00 20.36 274.67Standard Deviation 11.45 19.83 23.8 25.33 22.31 15.56 102.05

Table B.1 - Thickness of overbank sediments from pits, auger boreholes, bank analyses, and sediment cores in centimetres. Values marked by an asterisk give probable over-estimates for overbank thickness as overbank deposits here are often indistinguishable from channel deposits or represent multiple channel-overbank sequences. Refusal, which formed the cut-off for what was considered possible overbank sediment, for these values was medium gravel, which is too large to be recovered in the ESP cores. † - refusal not reached.

B.2 Grain Size Analysis of Floodplain Samples

Sample Depth from Floodplain Surface % Gravel % Sand % Silt S1-AH1-A1 2.0 0.00% 26.49% 73.51% S1-AH1-A2 93.5 0.00% 53.19% 46.81% S1-AH3-A1 13.5 0.00% 14.53% 85.47% S1-AH3-A2 48.0 0.00% 37.30% 62.70% S1-AH3-A3 91.0 0.00% 66.29% 33.71% S1-AH5-A1 4.5 0.00% 33.81% 66.19% S1-AH5-A2 67.5 0.00% 23.41% 76.59% S1-AH5-A3 113.0 0.00% 37.72% 62.28%

Site 1 Average 0.00% 36.59% 63.41% Standard Deviation 0.00% 16.60% 16.60%

S2-B1-A1 30.0 0.00% 38.38% 61.62% S2-B1-A2 56.0 0.00% 50.57% 49.43%

S2-AH1-A1 4.5 0.00% 44.02% 55.98%

Page 147: Downstream Variability of Fluvial Form, Process, and Character in a

136

S2-AH1-A2 50.0 0.00% 41.86% 58.14% S2-AH1-A3 79.5 0.00% 63.31% 36.69% S2-AH5-A1 4.0 0.00% 19.53% 80.47% S2-AH5-A2 43.0 0.00% 18.46% 81.54% S2-AH5-A3 93.0 2.61% 41.99% 55.40%

Site 2 Average 0.33% 39.76% 59.91% Standard Deviation 0.92% 14.95% 15.03%

S3-B1a 47.0 0.00% 34.07% 65.93% S3-B1b 87.0 0.00% 28.85% 71.15% S3-B1c 130.0 0.00% 27.71% 72.29%

S3-AH3-A1 4.5 0.00% 14.18% 85.82% S3-AH3-A2 79.5 0.00% 13.75% 86.25% S3-AH3-B1 127.5 0.00% 18.02% 81.98% S3-AH3-C1 192.0 0.00% 32.76% 67.24% S3-AH3-C2 259.0 0.00% 17.73% 82.27%

Site 3 Average 0.00% 23.38% 76.62% Standard Deviation 0.00% 8.36% 8.36%

S4u-B2-A1 13.0 0.00% 37.90% 62.10% S4u-B2-A2 40.0 0.00% 29.98% 70.02% S4u-B2-B 73.0 0.00% 56.44% 43.56%

S4u-AH4-A1 4.5 0.00% 38.32% 61.68% S4u-AH4-A2 51.5 4.48% 67.96% 27.56% S4u-AH5-A1 8.0 0.00% 52.25% 47.75% S4u-AH5-A2 78.0 0.00% 63.94% 36.06%

Site 4u Average 0.64% 49.54% 49.82% Standard Deviation 1.69% 14.41% 15.43%

S4d-P1-A1 14.0 2.12% 69.79% 28.09% S4d-P1-A2 43.0 0.27% 66.83% 32.90% S4d-B1-A1 14.0 0.00% 34.02% 65.98% S4d-B1-A2 50.0 0.18% 60.72% 39.11%

S4d-AH2-A1 4.5 0.43% 46.58% 52.99% S4d-AH2-A2 64.0 18.43% 33.65% 47.92%

S4d-B3-A 18.0 0.00% 36.56% 63.44% S4d-B3-B 51.0 0.00% 40.70% 59.30% S4d-B3-C 78.0 0.32% 65.43% 34.25%

Site 4d Average 2.42% 50.48% 47.11% Standard Deviation 6.04% 15.11% 14.14%

S5-B1-A1 9.0 0.00% 49.55% 50.45% S5-B1-A2 41.0 3.51% 64.71% 31.78% S5-AH8 4.0 1.24% 56.30% 42.46%

S5-P2-A1 8.0 0.00% 61.63% 38.37% S5-P2-A2 26.0 0.42% 67.36% 32.22% S5-P4-A1 10.0 0.00% 74.35% 25.65% S5-P4-A2 34.0 1.04% 50.30% 48.66%

Site 5 Average 0.89% 60.60% 38.51% Standard Deviation 1.27% 9.12% 9.23%

S6-1a 13.0 0.00% 34.27% 65.73% S5-1b 50.0 0.00% 53.26% 46.74% S6-1c 73.0 0.00% 15.51% 84.49% S6-1d 101.0 0.00% 30.83% 69.17%

Page 148: Downstream Variability of Fluvial Form, Process, and Character in a

137

S6-1e 137.0 0.00% 66.15% 33.85% S6-1f 162.0 0.00% 41.17% 58.83% S6-2a 18.0 0.00% 69.18% 30.82% S6-2b 90.0 0.00% 62.13% 37.87% S6-2c 124.0 0.00% 56.74% 43.26%

Site 6 Average 0.00% 47.69% 52.31% Standard Deviation 0.00% 18.26% 18.26%

Table B.2 - Crude grain size distribution of selected floodplain samples. Depth values are in centimetres.

B.3 Percent Organics in Floodplain Samples

Sample Depth from Floodplain Surface % OrganicsS1-AH1-A1 2.0 7.64% S1-AH1-A2 93.5 5.11% S1-AH3-A1 13.5 8.29% S1-AH3-A2 48.0 4.65% S1-AH3-A3 91.0 4.96% S1-AH5-A1 4.5 8.54% S1-AH5-A2 67.5 5.04% S1-AH5-A3 113.0 9.35%

Site 1 Average 6.70% Standard Deviation 1.94%

S2-B1-A1 30.0 5.42% S2-B1-A2 56.0 4.73%

S2-AH1-A1 4.5 8.19% S2-AH1-A2 50.0 4.42% S2-AH1-A3 79.5 2.49% S2-AH5-A1 4.0 6.91% S2-AH5-A2 43.0 5.07% S2-AH5-A3 93.0 1.79%

Site 2 Average 4.88% Standard Deviation 2.10%

S3-B1a 47.0 5.32% S3-B1b 87.0 2.94% S3-B1c 130.0 1.86%

S3-AH3-A1 4.5 11.30% S3-AH3-A2 79.5 3.71% S3-AH3-B1 127.5 3.05% S3-AH3-C1 192.0 2.83% S3-AH3-C2 259.0 3.25%

Site 3 Average 4.28% Standard Deviation 3.00%

S4u-B2-A1 13.0 6.78% S4u-B2-A2 40.0 5.10% S4u-B2-B 73.0 1.32%

S4u-AH4-A1 4.5 9.09%

Page 149: Downstream Variability of Fluvial Form, Process, and Character in a

138

S4u-AH4-A2 51.5 1.62% S4u-AH5-A1 8.0 4.80% S4u-AH5-A2 78.0 2.51%

Site 4u Average 4.46% Standard Deviation 2.86%

S4d-P1-A1 14.0 2.10% S4d-P1-A2 43.0 2.39% S4d-B1-A1 14.0 7.89% S4d-B1-A2 50.0 2.01%

S4d-AH2-A1 4.5 4.43% S4d-AH2-A2 64.0 2.59%

S4d-B3-A 18.0 6.42% S4d-B3-B 51.0 3.56% S4d-B3-C 78.0 1.48%

Site 4d Average 3.65% Standard Deviation 2.20%

S5-B1-A1 9 1.52% S5-B1-A2 41 1.15% S5-P2-A1 8 1.72% S5-P2-A2 26 1.71% S5-P4-A1 10 0.81% S5-P4-A2 34 1.92%

Site 5 Average 1.47% Standard Deviation 0.41%

S6-1a 13.0 2.65% S6-1b 50.0 1.17% S6-1c 73.0 2.38% S6-1d 101.0 1.74% S6-1e 137.0 0.84% S6-1f 162.0 1.95% S6-2a 18.0 1.24% S6-2b 90.0 1.09% S6-2c 124.0 1.19%

Site 6 Average 1.58% Standard Deviation 0.63%

Table B.3 - Percent organic content in selected floodplain samples based on loss-on-ignition method. Depth values are in centimetres.

Page 150: Downstream Variability of Fluvial Form, Process, and Character in a

139

Appendix C: Bank-full Discharge Formulae Analysis

The formulae presented in Chapter Four were tested against meandering gravel

and cobble-bed river data from Bray (1979), Church and Rood (1983), and Annable

(1996) to analyze how well each relation estimated bank-full discharge and if there was

any relation that was consistently better than the others. The results of this analysis are

presented in Tables C.1 – C.3. It should be noted that for the Bray (1979) data, D90 was

used instead of D84 and the values were tested against the 2-year discharge.

Williams (1978)

Lacey Limerinos-Manning

Darcy-Weisbach

Mean Error 15.08% 10.16% -5.25% 3.42% Standard Deviation 29.54% 31.25% 27.17% 27.70% Mean Absolute Error 22.25% 22.52% 22.67% 22.49% Standard Deviation 24.51% 23.80% 15.65% 16.29% Median Error 8.07% 1.32% -9.36% 7.66% Median Absolute Error 13.26% 13.98% 21.26% 20.54% Maximum Error 94.62% 121.21% 62.74% 61.98% Minimum Error -64.92% -59.57% -62.17% -65.79% 20% Error 67.16% 64.18% 46.27% 46.27% Underestimation 29.85% 47.76% 67.16% 34.33% Overestimation 70.15% 52.24% 32.84% 65.67%

Table C.1 - Statistics on errors for Bray (1979) data (n = 67).

Williams

(1978) Lacey Limerinos-

Manning Darcy-

Weisbach Mean Error 3.28% 12.38% 12.86% 15.09% Standard Deviation 29.93% 34.74% 46.93% 47.83% Mean Absolute Error 21.72% 27.19% 34.41% 35.15% Standard Deviation 20.73% 24.79% 34.24% 35.63% Median Error 0.02% 10.00% 1.92% 4.08% Median Absolute Error 14.76% 23.35% 25.77% 25.13% Maximum Error 105.47% 116.56% 190.60% 196.32% Minimum Error -52.52% -64.74% -79.51% -79.07% 20% Error 59.34% 46.15% 42.86% 42.86 Underestimation 49.45% 37.36% 48.35% 48.35% Overestimation 50.55% 62.64% 51.65% 51.65%

Table C.2 - Statistics on errors for Church and Rood (1983) data (n = 91).

Page 151: Downstream Variability of Fluvial Form, Process, and Character in a

140

Williams (1978)

Lacey Limerinos-Manning

Darcy-Weisbach

Mean Error 23.35% 33.36% 25.24% 27.71% Standard Deviation 37.44% 44.58% 51.88% 52.92% Mean Absolute Error 31.18% 38.76% 39.39% 40.56% Standard Deviation 30.94% 39.83% 41.83% 43.56% Median Error 15.48% 17.16% 8.53% 10.50% Median Absolute Error 20.41% 20.69% 22.90% 23.96% Maximum Error 118.42% 137.67% 157.41% 162.37% Minimum Error -30.90% -39.04% -54.00% -53.09% 20% Error 48.57% 48.57% 45.71% 42.86% Underestimation 22.86% 22.86% 40% 40% Overestimation 77.14% 77.14% 60% 60%

Table C.3 - Statistics on errors for Annable (1996) data (n = 35).

The data in all the tables have been stratified to determine which relation is the

best predictor. Bolded black numbers represent the best value, followed by blue, and

green. Red represents the worst value. It is a necessary requirement that the predictor

relation possesses the lowest errors in its estimates. Since mean and median errors can be

deceiving in that the underestimates tend to cancel out the overestimates, it was opted to

use the mean absolute error as a criterion for choosing the best relation. In all three

databases the Williams (1978) bank-full relation has the lowest error. However, only the

Church and Rood (1983) database demonstrates significant differences in mean absolute

errors between the four relations at the 95% confidence level, with the Williams (1978)

relation being significantly different from all the relations with the exception of the Lacey

relation (Tables C.4-C.7). Consequently, despite being the best predictor in this regard,

the Williams (1978) relation is not appreciably better consistently. Moreover, using the

median absolute errors in order to remove the effects of outliers reveals that the median

value in all three databases is 20% or less for the Williams (1978) relation, which is

comparable to the Lacey relation and less than the Limerinos-Manning and Darcy-

Weisbach relations.

Page 152: Downstream Variability of Fluvial Form, Process, and Character in a

141

Sum of Squares df Mean Square F Sig. Level Between groups 0.001 3 0.000 0.005 1.000 Within groups 11.069 264 0.042

Total 11.070 267

Table C.4 - Results of an analysis of variance test on the Bray (1979) dataset based on mean absolute error.

Sum of Squares df Mean Square F Sig. Level

Between groups 1.109 3 0.370 4.244 0.006 Within groups 31.367 360 0.087

Total 32.476 363

Table C.5 - Results of an analysis of variance test on the Church and Rood (1983) dataset based on mean absolute error.

Sum of Squares df Mean Square F Sig. Level Between groups 0.191 3 0.064 0.410 0.746 Within groups 21.050 136 0.155

Total 21.240 139

Table C.6 - Results of an analysis of variance test on the Annable (1996) dataset based on mean absolute error.

Relation 1 Relation 2 Significance Level

Williams (1978) Lacey 1.000 Limerinos-Manning 0.024 Darcy-Weisbach 0.014

Lacey Limerinos-Manning 0.599 Darcy-Weisbach 0.419

Limerinos-Manning Darcy-Weisbach 1.000

Table C.7 - Results of a Bonferroni post-hoc test on Church and Rood (1983) dataset.

Furthermore, one must take into consideration the variability in the errors with the

lowest variability being the most desirable. Many possible variables could be looked at

from the tables including standard deviation, the estimate extremes, or the percentage of

the values that are under-predicted versus over-predicted. The criterion selected for

variability was the percentage of the values within 20% of the reported discharge. On

average, between the three databases, 58% of the values for the Williams (1978) relation

fall within 20% above or below the reported discharge value, which is higher than the

other relations (Lacey: 53%, Limerinos-Manning: 45%, Darcy-Weisbach: 44%). In

Page 153: Downstream Variability of Fluvial Form, Process, and Character in a

142

regards to the other categories on variability mentioned above, the Williams (1978)

relation tends to, but not always, possess the best values such as the lowest extremes, the

lowest standard deviations, and the lowest under-/over-estimates, which help to further

solidify the selection of this relation. Moving forward with the Williams (1978) relation

as the bank-full discharge estimator for the current study, it should be considered that the

estimates will most likely be over-predictions given that the relation in all three databases

has a positive mean error and the majority of the distribution is overestimated. Using the

Williams (1978) relation, the average bank-full width and depth, and the surveyed

channel slope, the following bank-full discharge estimates were obtained, which compare

favourably with those predicted using the other relations (Table C.8).

Site 1 2 3 4u 4d 5 6 Width (m) 5.64 9.42 9.11 13.38 13.64 17.45 20.51 Depth (m) 0.68 0.65 1.04 0.73 0.73 0.70 0.81 Channel Slope 0.0002 0.0072 0.0002 0.0017 0.0043 0.0069 0.0035 Discharge (m3s-1) 1.87 8.95 5.57 10.59 14.02 20.55 24.64 Lacey Relation 1.66 8.85 5.47 9.73 13.45 19.14 22.95 Manning Relation 1.32 8.89 4.79 8.04 12.37 19.76 26.68 Darcy Relation 1.34 9.07 4.88 8.20 13.63 20.15 27.18

Table C.8 - Bank-full discharge estimates for the study sites. The channel slope of 0.0028 for Site 1 was found to be unacceptable as an estimate of the energy conditions and was replaced by an estimate of 0.0002 based on the surveyed water level, which was near bank-full stage.

Page 154: Downstream Variability of Fluvial Form, Process, and Character in a

143

Appendix D: Flood Frequency Analysis After initially screening 46 gauging stations in watersheds between the Niagara

Escarpment, Peterborough, Lake Ontario, and Lake Simcoe, 11 stations were chosen

within the vicinity of the Little Rouge watershed that were believed to produce similar

hydrological responses to the Little Rouge River. These 11 stations were chosen by

regressing their 2-year return period discharge, selected from a distribution-fitted annual

maximum series curve, against drainage area and selecting stations that form a trend with

the Little Rouge station. The chosen stations form the nucleus of the flood frequency

analysis and their basic characteristics are summarized in Table D.1.

Water Survey of Canada (WSC) Station Name

WSC Station ID

Drainage Area (km2)

Record Period Length (years)

Lynde Creek near Whitby 02HC018 106.0 1959 – 2010 52 Duffins Creek above Pickering 02HC019 93.5 1960 – 2009 50 Rouge River near Markham 02HC022 186.0 1961 – 2010 50 West Duffins Creek at Green River 02HC026 98.1 1963 – 1988 26 Little Rouge Creek at Locust Hill 02HC028 77.7 1963 – 2010 48 West Duffins Creek above Green River 02HC038 52.0 1974 – 2010 37 Duffins Creek at Ajax 02HC049 251.0 1989 – 2010 22 Bowmanville Creek at Bowmanville 02HD006 82.9 1959 – 2010 52 Oshawa Creek at Oshawa 02HD008 95.8 1959 – 2010 52 Wilmot Creek near Newcastle 02HD009 82.6 1965 – 2010 46 Harmony Creek at Oshawa 02HD013 41.6 1980 – 2010 31

Table D.1 - Gauging stations used in the flood frequency analysis.

Annual discharge data for each of the above stations were ordered from highest to

lowest and then fitted to the General Extreme Value (GEV) distribution for the annual

maximum series (AMS) daily and instantaneous values and the General Pareto (GP)

distribution for the partial duration series (PDS) maximum daily values (Kidson and

Richards, 2005). The formulae for each distribution are outlined below, summarized from

Glaves and Waylen (1997) and Adamowski et al. (1998).

Page 155: Downstream Variability of Fluvial Form, Process, and Character in a

144

General Extreme Value Distribution

The General Extreme Value distribution estimates discharge quantiles based on

the cumulative probability function and three site-specific parameters that characterize

the shape (skew), scale (variation), and the location (mean) of the distribution. The

formula for the distribution is outlined as

k

GEV

β 1- -ln pQ = ζ +

k

for k 0 D.1a

= ζ - βln 1 - ln(p) for k = 0 D.1b

where QGEV is the discharge, is the location parameter, is the scale parameter, k is the

shape parameter, and p is the cumulative probability of the quantile (ie. 1-exceedance

probability). The parameters can be estimated using L-moments, which are linear

combinations of ranked data (ie. annual floods) that are stated to be unbiased estimates

compared to product-moments since they do not need to be cubed or squared

(Adamowski et al., 1998). The formulae for calculating the parameters through L-

moments are

1

β Γ 1 + k - 1ζ = λ +

k

D.2

2-k

λ kβ =

1 - 2 Γ 1 + k D.3

2k = 7.859c + 2.9554c D.4

where 1 is the first L-moment, 2 is the second L-moment, is the gamma function, and

c is equal to

3 2

2 log(2) -

λ /λ + 3 log(3) D.5

Page 156: Downstream Variability of Fluvial Form, Process, and Character in a

145

with 2 and 3 representing the second and third L-moments respectively. The four L-

moments can be summarized as follows:

1 0λ = b D.6

2 1 0= 2b – bλ D.7

3 2 1 0= 6b – 6b + bλ D.8

4 3 2 1 0λ = 20b – 30b + 12b – b D.9

where 1-4 are respectively the mean, standard deviation, skewness, and kurtosis and, b0-3

are the probability weighted moments for the sample data. The probability-weighted

moments can be defined as:

n

jj=1

0

x

b = n

D.10

n - 1

1 jj = 1

n - jb = x

n n - 1

D.11

n - 2

2 jj = 1

n - j n - j - 1b = x

n n - 1 n - 2

D.12

n - 3

3 jj = 1

n - j n - j - 1 n - j - 2b = x

n n - 1 n - 2 n - 3

D.13

where xj is a value from the sample with rank j when the sample values are ranked from

highest to lowest, and n is the total number of values in the sample.

Page 157: Downstream Variability of Fluvial Form, Process, and Character in a

146

General Pareto Distribution

The General Pareto distribution is very similar to the GEV distribution above in

that it estimates discharge quantiles based on a cumulative probability function and three

parameters that characterize the distribution. The difference between the two is that the

General Pareto distribution is useful for partial duration series data as opposed to the

annual maximum series. In defining the database for the partial duration series, it is

necessary to establish a cut-off point for the inclusion of discharge values, which was

taken as the lowest discharge occurring on the annual maximum series. In this way, at

least one discharge value from each year in the record period will appear in the database

(Langbein, 1949). The General Pareto distribution can be defined as:

k

GP

α 1 - 1 - pQ = μ +

k

for k ≠ 0 D.14a

= μ – αln 1 – p for k = 0 D.14b

where QGP is the discharge estimate, μ is the location parameter, α is the scale parameter,

k is the shape parameter, and p is the cumulative probability of the discharge occurring

(ie. 1 – exceedance probability). The three parameters can be defined as:

12

2 + kμ = λ –

λ D.15

2α = 1 + k 2 + k λ D.16

3 2

3 2

1 - 3 λ /λk =

1 + λ /λ D.17

where λ1-3 are the first, second, and third L-moments from Equations D.6-D.8.

Page 158: Downstream Variability of Fluvial Form, Process, and Character in a

147

Following this, discharges for various return periods ranging from 0.5 years to 20

years were calculated for each station based on their maximum instantaneous (AMS) and

maximum daily (AMS and PDS) distributions. These values were then compiled and

regressed against drainage area to formulate regional discharge-drainage area rating

curves for each return period. The coefficients and exponents are listed in Table D.2.

The relations were tested against the Little Rouge at Locust Hill (used in the

development of the relations) and Duffins Creek at Pickering (not used in the

development; station ID 02HC006) stations to see how they performed. The results of the

gauging station analysis are mixed at best with the discharges of the Little Rouge River

being better predicted by the relations than for Duffins Creek (Tables D.3 and D.4).

Generally, the relations under-predict the discharges with the only exception being the

maximum instantaneous values for the Little Rouge. For the Little Rouge data, the best

predictor was the partial duration series relations, which had the lowest error values while

the maximum instantaneous relations generated the highest errors. While all three series

relations had very similar average error values for the Duffins Creek data, the partial

duration series relations tended to be more consistent than the other two. Despite the

predictions, the Little Rouge data typically lay within one standard error of the estimate

while the Duffins Creek data lay outside, particularly for values between 20 and 80 m3s-1

(Figs. D.1 and D.2). Whether or not the Duffins Creek station is an anomalous occurrence

is not known, but the results from this station are not as encouraging.

As to which series should be used for the estimate of bank-full discharge return

period, one must consider the reliability of the series and the range of return periods that

can be estimated. The preceding discussion already alluded to the partial duration series

Page 159: Downstream Variability of Fluvial Form, Process, and Character in a

148

as likely the most reliable. The value of the partial duration series over the annual series

is increased when the range of available return periods is considered, such that three extra

return periods are gained using the partial duration series. Furthermore, it was discovered

Series Return Period a b Standard Error of Estimate r2 AMS-MD 1.05 0.699 0.463 0.246 0.499

1.10 0.520 0.580 0.180 0.745 1.50 0.393 0.754 0.140 0.890 2.00 0.407 0.793 0.143 0.897 2.50 0.433 0.807 0.144 0.898 3.00 0.460 0.812 0.145 0.898 5.00 0.557 0.813 0.150 0.892 7.50 0.659 0.805 0.157 0.880 10.00 0.745 0.796 0.165 0.868 15.00 0.893 0.780 0.179 0.842 20.00 1.017 0.767 0.191 0.819

PDS-MD 0.50 0.279 0.697 0.255 0.677 0.75 0.296 0.749 0.198 0.800 1.00 0.318 0.770 0.182 0.834 1.05 0.322 0.773 0.181 0.837 1.10 0.326 0.775 0.179 0.840 1.50 0.356 0.789 0.172 0.856 2.00 0.389 0.797 0.168 0.863 2.50 0.417 0.801 0.166 0.868 3.00 0.442 0.803 0.164 0.871 5.00 0.520 0.806 0.161 0.876 7.50 0.592 0.804 0.159 0.877 10.00 0.649 0.802 0.160 0.876 15.00 0.738 0.798 0.163 0.871 20.00 0.808 0.794 0.167 0.864

AMS-MI 1.05 0.159 0.852 0.524 0.427 1.10 0.053 1.153 0.499 0.600 1.50 0.018 1.506 0.551 0.678 2.00 0.015 1.593 0.593 0.670 2.50 0.014 1.630 0.621 0.660 3.00 0.014 1.652 0.641 0.651 5.00 0.014 1.689 0.694 0.625 7.50 0.014 1.705 0.733 0.604 10.00 0.015 1.712 0.759 0.589 15.00 0.016 1.717 0.794 0.568 20.00 0.017 1.718 0.818 0.554

Table D.2 - Exponents and coefficients for the discharge-area rating curves of various return periods (years). AMS = annual maximum series, PDS = partial duration series, MD = maximum daily values, MI = maximum instantaneous values. The rating curves take on the form of Q = aAd

b, where Q is the discharge for the chosen return period (m3s-1), Ad is the drainage area (km2), and a and b are the values from the table that correspond to the chosen series and return period. Standard error of the estimate values are in ln units.

Page 160: Downstream Variability of Fluvial Form, Process, and Character in a

149

% Difference % Absolute Difference Return Period

AMS–MD AMS–MI PDS–MD AMS–MD AMS–MI PDS–MD

0.5 6.70% 6.70%0.75 -5.31% 5.31%

1 -8.99% 8.99%1.05 -27.75% 11.25% -9.31% 27.75% 11.25% 9.31%1.1 -24.62% 10.39% -9.62% 24.62% 10.39% 9.62%1.5 -17.36% 10.25% -11.42% 17.36% 10.25% 11.42%2 -13.48% 12.42% -12.06% 13.48% 12.42% 12.06%

2.5 -11.11% 13.65% -12.19% 11.11% 13.65% 12.19%3 -8.83% 15.04% -12.11% 8.83% 15.04% 12.11%5 -2.88% 19.32% -10.48% 2.88% 19.32% 10.48%

7.5 2.05% 22.60% -9.15% 2.05% 22.60% 9.15%10 5.31% 25.13% -7.59% 5.31% 25.13% 7.59%15 10.52% 29.04% -5.02% 10.52% 29.04% 5.02%20 14.10% 31.91% -3.24% 14.10% 31.91% 3.24%

Average -6.73% 18.27% -7.84% 12.55% 18.27% 8.80%Standard Deviation 13.83% 7.82% 5.04% 8.24% 7.82% 2.87%

Table D.3 - Relative error values as a percent difference between the predicted and observed discharges for the Little Rouge gauging station.

% Difference % Absolute Difference Return Period

AMS–MD AMS–MI PDS–MD AMS–MD AMS–MI PDS–MD

0.5 14.06% 14.06%0.75 -9.68% 9.68%

1 -16.01% 16.01%1.05 24.54% -57.98% -16.64% 24.54% 57.98% 16.64%1.1 -17.49% -43.68% -17.20% 17.49% 43.68% 17.20%1.5 -30.75% -23.97% -20.48% 30.75% 23.97% 20.48%2 -28.81% -18.17% -22.16% 28.81% 18.17% 22.16%

2.5 -26.89% -15.93% -22.98% 26.89% 15.93% 22.98%3 -24.75% -14.35% -23.48% 24.75% 14.35% 23.48%5 -18.86% -11.00% -23.39% 18.86% 11.00% 23.39%

7.5 -13.92% -9.42% -23.36% 13.92% 9.42% 23.36%10 -10.66% -8.38% -22.78% 10.66% 8.38% 22.78%15 -5.68% -6.96% -21.63% 5.68% 6.96% 21.63%20 -2.34% -6.00% -20.94% 2.34% 6.00% 20.94%

Average -14.15% -19.62% -17.62% 18.61% 19.62% 19.63%Standard Deviation 15.87% 16.62% 9.93% 9.55% 16.62% 4.25%

Table D.4 - Relative error values as a percent difference between the predicted and observed discharges for the Duffins Creek gauging station.

that beyond the 1.5-year return period, the values between the partial duration series and

annual series (both using maximum daily values) are quite similar (Fig. D.3), which

Page 161: Downstream Variability of Fluvial Form, Process, and Character in a

150

means that the difference between the two series lies in the discharges occurring more

frequently than 1.5-years. Since the same cut-off value was used in constructing both the

partial duration and annual series, the values below a return period of 1.5 years have to fit

into five return period categories for the partial duration series versus two for the annual

series, which suggests that the discharges for return periods below 1.5 years on the annual

series may be under-estimated compared to the same return period on the partial duration

series.

0

10

20

30

40

50

60

70

0 5 10 15 20 25 30 35 40

Observed Discharge (m3s-1)

Pre

dic

ted

Dis

char

ge

(m3 s

-1)

AMS - MD AMS - MI PDS - MD 1:1 Line

Figure D.1 - Plot of observed discharges from the Little Rouge River gauging station against the predicted discharges for the various series. Error bars represent one standard error of the estimate.

Given that the partial duration series is generally more accurate, is very similar to

the annual series for return periods over 1.5-years, and has greater predictive capacity

below 1.5-years leading to more accurate representation of the return period, the partial

duration series is advocated as being the best predictor of the bank-full discharge return

periods for this study. However, it should be noted that as the partial duration series is for

maximum daily values, it over-estimates the return period of a given discharge.

Maximum instantaneous values should be considered in conjunction with the

Page 162: Downstream Variability of Fluvial Form, Process, and Character in a

151

0

20

40

60

80

100

120

140

0 20 40 60 80 100 120

Observed Discharge (m3s-1)

Pre

dic

ted

Dis

char

ge

(m3 s

-1)

AMS - MD AMS - MI PDS - MD 1:1 Line

Figure D.2 - Plot of observed discharges from the Duffins Creek gauging station against the predicted discharges for the various series. Error bars represent one standard error of the estimate.

0.8

1

1.2

1.4

1.6

1.8

2

2.2

0 5 10 15 20 25

Return Period (years)

Dis

char

ge

(PD

S -

MD

) :

Dis

char

ge

(AM

S -

MD

)

Figure D.3 - Average ratio of the discharge from the partial duration series and the annual series for a given return period for the 11 gauging stations. On average, for a given return period below 1.5-years, discharges are 1.5-2 times larger on the partial duration series than the annual series. Discharges on the two series are about equal at the 1.5-year return period and stabilize at between 0.90 and 0.98 by the 2-year return period. Error bars represent one standard error. partial duration series, but the lower limit of its usefulness is likely reached at the 1.5-

year return period. The rating curves presented in Table D.2 for the partial duration series

and maximum annual series (instantaneous values) were applied to the study sites to

Page 163: Downstream Variability of Fluvial Form, Process, and Character in a

152

estimate the return period of bank-full discharge. The results of this are displayed

graphically in Figures D.4-D.10 and summarized in Table D.5.

1

10

100

0.1 1 10 100

Return Period (years)

Dis

char

ge

(m3 s-1

)

PDS - MD Bank-full Discharge AMS - MI

Figure D.4 - Bank-full discharge flood frequency estimate for Site 1. Error bars represent one standard error of the estimate.

1

10

100

0.1 1 10 100Return Period (years)

Dis

char

ge

(m3s-1

)

PDS - MD Bank-full Discharge AMS - MI

Figure D.5 - Bank-full discharge flood frequency estimate for Site 2. Error bars represent one standard error of the estimate.

Page 164: Downstream Variability of Fluvial Form, Process, and Character in a

153

1

10

100

0.1 1 10 100Return Period (years)

Dis

char

ge

(m3s-1

)

PDS - MD Bank-full Discharge AMS - MI

Figure D.6 - Bank-full discharge flood frequency estimate for Site 3. Error bars represent one standard error of the estimate.

1

10

100

0.1 1 10 100Return Period (years)

Dis

char

ge

(m3 s-1

)

PDS - MD Bank-full Discharge AMS - MI

Figure D.7 - Bank-full discharge flood frequency estimate for Site 4u. Error bars represent one standard error of the estimate.

Page 165: Downstream Variability of Fluvial Form, Process, and Character in a

154

1

10

100

0.1 1 10 100Return Period (years)

Dis

char

ge

(m3 s-1

)

PDS - MD Bank-full Discharge AMS - MI

Figure D.8 - Bank-full discharge flood frequency estimate for Site 4d. Error bars represent one standard error of the estimate.

1

10

100

0.1 1 10 100Return Period (years)

Dis

char

ge

(m3 s-1

)

PDS - MD Bank-full Discharge AMS - MI

Figure D.9 - Bank-full discharge flood frequency estimate for Site 5. Error bars represent one standard error of the estimate.

Page 166: Downstream Variability of Fluvial Form, Process, and Character in a

155

1

10

100

0.1 1 10 100Return Period (years)

Dis

char

ge

(m3 s-1

)

PDS - MD Bank-full Discharge AMS - MI

Figure D.10 - Bank-full discharge flood frequency estimate for Site 6. Error bars represent one standard error of the estimate.

Site 1 2 3 4u 4d 5 6 Drainage Area (km2)

16.99 21.70 73.14 86.53 86.65 107.42 114.12

Discharge (m3s-1) 1.87 8.95 5.57 10.59 14.02 20.55 24.64 PDS-MD Return Period (years)

0.5 20 0.5 1.1 2 3 5

Range <0.5-0.5 10->20 <0.5-0.5 1-1.5 1.5-3 2.5-5 5-7.5 AMS-MI Return Period (years)

<0.5 1.5 <0.5 <1.1 <1.5 <1.5 1.5

Range <0.5 <1.5-2.5 <0.5 <1.1 <1.5 <1.5-1.5 <1.5-2

Table D.5 - Summary of bank-full return periods for the maximum daily partial duration series and the maximum instantaneous annual series rating curves for each site.

Page 167: Downstream Variability of Fluvial Form, Process, and Character in a

156

Appendix E: Downstream Trends in Channel and Floodplain Characteristics

E.1 Width-Depth Ratio and Sinuosity

The interaction between bank-full width and depth is reflected in the width-depth

ratio (Fig. E.1), which shows a moderate downstream increase. Sites 1 and 3 possess the

lowest width-depth ratios, which, in the case of Site 3, is a result of increased bank-full

depth and a concomitant decrease in width (see Chapter Six, section 6.3.1). The higher

values in the downstream end of the watershed are brought on by channel width

increasing against a constant to slightly decreasing channel depth.

Figure E.1 - Downstream variation in bank-full width-depth ratio (w:d) and channel sinuosity (P).

Channel sinuosity (P) shows a very weak downstream decrease. All values are

below 1.30 except for sites 1 and 3, which are coincident with the lowest width-depth

ratios. It should be noted that sinuosity is a somewhat arbitrary value as the start and end

points for measurements can vary between users. Moreover, sinuosity does not stay

constant over a selected distance and can change quite rapidly over a short horizontal

distance. Taking these realities into consideration, the sinuosity values discussed above

represent general approximations for each site.

Page 168: Downstream Variability of Fluvial Form, Process, and Character in a

157

E.2 Bank-full Discharge Return Period

It is often expected that return periods of bank-full discharge along a channel

fluctuate between 1 and 2 years on the annual maximum series (Wolman and Leopold,

1957). This, however, has been shown to be a general approximation at best with return

periods in the data compilation of Williams (1978) ranging from one year to over 50

years on the annual series. When the partial duration series is considered, the range is

even larger. The downstream variability in return periods has not been well studied, but

has been shown to possess a positive relationship with drainage area by Petit and Pauquet

(1997). The opposite is observed as an overall trend in the Little Rouge watershed using

maximum daily flow values from the partial duration series, which is driven by the

comparatively high value at Site 2 (Fig. E.2) (for derivation of these values and possible

ranges, see Appendix D). The remaining sites exhibit the expected positive trend.

Figure E.2 – Downstream variation in bank-full discharge return period.

Page 169: Downstream Variability of Fluvial Form, Process, and Character in a

158

E.3 Channel Bed Alluvium Thickness

Channel bed alluvium thickness shows a very weak downstream increase, though

the slope of the regression line is not significantly different from zero (Fig. E.3).

Thickness values generally fall within the range of 25-60 cm with the only departure

from this being Site 6 (91 cm), which is responsible for the impression of a downstream

increasing trend. In half of the cross-sections at Site 6 basal bedrock was never reached.

Thus, the actual alluvial thickness at this site is under-estimated. Site 3 is also under-

estimated but not likely to the same extent. Based on the limited data, two smaller trends

can be discerned from the overall trend where the average alluvial bed thickness

increases between sites 1 and 3, and 4 and 6. These two smaller trends are separated by a

decrease in thickness of up to 55%. The maximum alluvial thickness tends to follow the

trend observed for average alluvium thickness and values vary between 80 and 165 cm,

though sites 3 and 6, which represent the thickest sites, are under-estimates.

The average percentage of glacial clay (or clay-bedrock in the case of Site 6) that

outcrops within the bank-full channel varies between 0% and 24%, and appears to show

no relationship with distance (Fig. E.4). Since the values in Figure E.6 are site-averages

based on riffle and pool cross-sections, the large error bars about each average are a

reflection of the variability in clay exposure within the two bed-forms. Clay was most

often exposed in channel pools where the channel became coupled with valley or terrace

sides, though instances were also observed when the channel was decoupled from the

valley sides, while riffles were fully composed of gravel in many cases.

Page 170: Downstream Variability of Fluvial Form, Process, and Character in a

159

Figure E.3 - Downstream variation in maximum and average channel bed alluvium thickness. Error bars represent one standard error.

Figure E.4 - Downstream variation in exposed glacial clay/bedrock within the bank-full channel. Error bars represent one standard error.

Page 171: Downstream Variability of Fluvial Form, Process, and Character in a

160

E.4 Silt/Clay Content and Organic Content

The general downstream pattern of the depth-averaged silt/clay content within the

floodplain is similar to that described for overbank thickness (Fig. E.5). Overall, there is a

weakly defined negative trend with a slope not significantly different from zero. Silt/clay

percentages are above 45% in all of the sites except for Site 5 where silt and clay

compose 39% of the floodplain sediment. Sites upstream of Site 4 tend to be finer with

silt/clay percentages about 60% or greater, while the other sites have less than 55%

silt/clay indicating slightly sandier floodplains downstream.

Depth-averaged organic content tends to show a strongly decreasing downstream

trend (Fig. E.6). This trend is realized because sites 5 and 6 possess significantly less

organic-rich floodplains than at Site 4 and farther upstream.

Figure E.5 - Downstream variation in depth-averaged silt/clay content. Error bars represent one standard error.

Page 172: Downstream Variability of Fluvial Form, Process, and Character in a

161

Figure E.6 - Downstream variation in depth-averaged organic content. Error bars represent one standard error.

Page 173: Downstream Variability of Fluvial Form, Process, and Character in a

162

Appendix F: Glacial Control on Downstream Trends F.1 Introduction

The objective of the following analysis is to explain the downstream trends

described in the preceding appendix in relation to the relevant physical controls (ie.

discharge, valley slopes) and sediment inputs. While it is most desirable to explain the

downstream trends as a function of these variables, it is understood that fluvial systems

often involve more complicated interactions with intrinsic and extrinsic influences. In

some instances, discharge, valley slope, and sediment inputs may not be the only

controlling parameters behind the observed trends. In which case, other intrinsic or

extrinsic forces may be sought to explain these occurrences.

F.2 Channel Morphology, Hydrology, and Sedimentology

F.2.1 Channel Dimensions and Cross-Sectional Area

In equilibrium channels, the bank-full dimensions represent a balance between the

prevailing hydrological regime (often represented by bank-full discharge), channel slope,

supply rate of bed-load sediment, the bed material size, and bank-strength (see Table

2.1). Separating these variables apart is difficult since their inherent downstream

variation makes controlling for them unrealistic in many situations. However, by dividing

the bank-full dimensions by the upstream drainage area at each site and plotting these

‘relative dimensions’ against bed material size, discharge, and channel slope, some

inferences can be made on the downstream adjustment in the channel dimensions.

It is often regarded that bank-full discharge exerts the strongest control over the

channel dimensions (refer to Table 2.1), but since bank-full discharge was derived from

the dimensions themselves, plotting the dimensions per unit drainage area (hereafter

Page 174: Downstream Variability of Fluvial Form, Process, and Character in a

163

referred to as relative width, depth, and area) against bank-full discharge would lead to

spurious correlations (Ferguson, 1986). Instead, as it is theoretically thought that river

channels are adjusted to a flow of a return frequency of 1-2 years (Wolman and Leopold,

1957), the 2-year discharge as derived from the relations in Appendix D is used to replace

bank-full discharge. If the 2-year discharge is the strongest control over the bank-full

channel dimensions, then the downstream variation in the relative channel dimensions

can be modelled using simple power relations. Leopold and Maddock (1953) show that

bank-full channel width (wbf) and depth (dbf) vary with bank-full discharge (Qbf)

respectively as

0.50bf bfw Q F.1a

0.40bf bfd Q F.1b

with bank-full cross-sectional area (Abf) necessarily varying with discharge as

0.90bf bfA Q F.1c

A relation that relates the 2-year discharge (Q2) as a function of drainage area (Ad)

is given as (R. Phillips, personal communication)

0.912 dQ A F.2

Substituting Equation F.2 into Equation F.1 yields

0.82

bf dA A F.3a

0.46

bf dw A F.3b

0.36

bf dd A F.3c

Page 175: Downstream Variability of Fluvial Form, Process, and Character in a

164

Dividing each relation by drainage area yields the theoretical downstream adjustment of

relative width, depth, and cross-sectional area as a function of drainage area (Eq. F.4) and

the Q2 (Eq. F.5), given as

-1 -0.18bf d dA A A F.4a

-1 -0.54bf d dw A A F.4b

-1 -0.64bf d dd A A F.4c

-1 -0.20bf d 2A A Q F.5a

-1 -0.59bf d 2w A Q F.5b

-1 -0.70bf d 2d A Q F.5c

Any deviation from Equation F.5 can be interpreted in the context of grain size, channel

slope, bank strength, and other controlling variables.

The exponent for relative cross-sectional area plotted against the 2-year discharge

is –0.53, which is steeper than predicted by Equation F.5a by 165% (Fig. F.1a) and

implies that cross-sectional area is increasing slower downstream than discharge, which

may be a result of relatively wide and deep channels upstream or narrow and shallow

channels downstream. Multiplying the regime relations developed by Millar (2005),

bank-full cross-sectional area (Abf) can be expressed as

0.86 -0.15 0.02 -0.46bf bf 50 CA Q D S μ’ F.6

where Qbf is the bank-full discharge (m3s-1), D50 is the median bed grain size (m), SC is

the channel slope (-), and ’ is the relative bank strength compared to the channel bed (-).

Equation F.6 suggests that discharge exerts the most control over cross-sectional area and

that increasing grain size results in a decrease in cross-sectional area and channel slope

Page 176: Downstream Variability of Fluvial Form, Process, and Character in a

165

has no effect, both of which are reflected, albeit very poorly, in Figure F.1b and c. The

effects of bank vegetation cannot be quantitatively assessed here. The fact that bank-full

cross-sectional area cannot adequately be explained by discharge, grain size, or channel

slope alludes to external controls.

Figure F.1 - Relative cross-sectional area as a function of discharge (a), grain size (b), and channel slope (c). Sites are given for reference.

Wohl and Achyuthan (2002) demonstrate how channels flowing over resistant

substrates tend to possess smaller cross-sectional areas than channels flowing over

alluvium. The sites in Figure F.1 can be segregated into glacial clay and alluvium, which

represent the two dominant substrate types. Site 6 is carved into a thick sequence (>2 m)

of sand, while all sites upstream of Site 6 are cut into glacial clays (tills or

glaciolacustrine clays), which possess varying erodibilities that resist morphological

Page 177: Downstream Variability of Fluvial Form, Process, and Character in a

166

adjustment (Shugar et al., 2007) compared to the relatively erodible alluvium. As a result,

the cross-sectional area of the channel at Site 6 is larger per unit drainage area than the

upstream sites (Fig. F.1), and may be manifest as a wider and deeper channel.

While differential substrate erodibility adequately explains the downstream

increase in cross-sectional area between sites 5 and 6, this rationale cannot be extended to

sites 1 or 2, which are underlain by glacial clay and possess very large relative cross-

sectional areas. It is believed that the large channel area present at Site 2 is the result of

bed and bank scouring caused by the presence of debris dams along the channel, which

are absent in the downstream sites (Fig. F.2). Keller and Swanson (1979) note that flow

deflection caused by debris dams can cause bank erosion citing a study in which such a

scenario induced a 230% increase in channel width. Furthermore, flow deflection towards

the banks by mid-channel bars accreted downstream of dams may also widen the

channels (Fig. F.3). Local scouring in the pool immediately downstream of the dam and

scouring of the bed when the dam breaks may cause increases in channel depth as well,

such that, after the dam breaks, the legacy of the debris dam remains as an over-widened

and deepened channel.

Additionally, and perhaps equally important, the large width relative to the

drainage area at Site 2 may be a result of the vegetation lining the banks. Hession et al.

(2003) and Allmendinger et al. (2005) note that channels flowing through forested

floodplains are wider than channels flowing through grassy floodplains, which

Allmendinger et al. (2005) attribute to the trapping of sediment by grassy vegetation

along the inner banks of the channel. Such vegetation is typically absent on forested

floodplains due to the forest canopy blocking out much of the sunlight necessary for

Page 178: Downstream Variability of Fluvial Form, Process, and Character in a

167

grasses to grow. The channel at Site 2 flows through a dense forest of coniferous trees

with a ground cover consisting of sparse and low-lying herbaceous scrubs and minimal

grasses. Due to the lack of light along the floodplain, the over-widened channel cannot be

reclaimed by colonizing vegetation, and, as a result, the wide channel and large cross-

sectional area persists.

Lastly, as Site 1 flows through a grass-dominated floodplain, the preceding

explanation may not necessarily apply to this site. Instead, it is believed that the

occurrence of several relic beaver dams along the channel has caused the channel to

become wider and deeper (Fig F.4). Little empirical information is available on the

effects of beaver dam construction on channel dimensions to support why this might be.

However, since beaver dams are, in essence, debris dams (Gurnell, 1998), it is

conceivable that the widening effects of debris dams mentioned previously may be

extended to Site 1.

Figure F.2 - Debris dam at Site 2 during low-water stage. Flow direction given by the arrow.

Page 179: Downstream Variability of Fluvial Form, Process, and Character in a

168

Figure F.3 – Bar-form developed behind fallen trees in the channel. Notice the downstream gradation of sediment along the bar and lack of colonizing vegetation. Coarse gravel is at the bar-head while sand is at the bar-tail. Flow direction given by the arrow.

Figure F.4 - Beaver dam looking downstream at Site 1. The crest of the dam is at approximately bank-full stage.

Page 180: Downstream Variability of Fluvial Form, Process, and Character in a

169

Figure F.5a and b show how relative channel width and depth vary downstream as

a function of the 2-year discharge. The exponents are –0.65 and –1.14 for width and

depth respectively. While the width exponent is comparable to that in Equation F.5b,

scatter remains about the trend line. Moreover, there appears to be an increasing trend

between sites 3 and 6, meaning that the channel width is increasing faster downstream

than the 2-year discharge. The large apparent gap between sites 2 and 3 is a consequence

of the structure of the watershed where almost half of the total drainage area enters

between these two sites raising the 2-year discharge. The downstream trend in channel

depth shows very little scatter but the exponent of the trendline is 63% steeper than that

predicted in Equation F.4c, meaning that channel depth is increasing much slower than

discharge is downstream. These deviations from the expected relations, through either

scatter or different exponents, allude to other controls on bank-full channel morphology.

The trends in Figure F.5a and b have been described by the variation in channel

slope (Fig. F.5c and d) and median grain size of the bed material (Fig. F.5e and f). In

Figure F.5c-f, the relations are poorly defined with no relationship being statistically

significant, which is a consequence of sites 1 and 2 plotting well above the other sites.

Nevertheless, despite the lack of statistically defined relations, basic visual trends can be

discerned from the data. Considering sites 1 and 2 as their own group (affected by debris

dams) and sites 3-6 (not affected by debris dams) as another, it appears that when channel

slope increases, relative width also increases while relative depth decreases. In addition,

channels with median grain sizes above 10 mm are typically wider and shallower than

channels with median sizes below 10 mm, though these grain size trends are not

particularly well defined.

Page 181: Downstream Variability of Fluvial Form, Process, and Character in a

170

Figure F.5 - Relative width and depth as a function of discharge (a and b), channel slope (c and d), and bed material size (e and f). Sites numbers are given for reference.

The slope-width and slope-depth relations are in agreement with rational regime

relations (Parker, 1979; Griffiths, 1981; Millar, 2005) but are in disagreement with the

semi-theoretical relations of Lee and Julien (2006) and the empirically derived hydraulic

geometry relations of Davidson and Hey (2011). Results from a study by Nanson and

Page 182: Downstream Variability of Fluvial Form, Process, and Character in a

171

Young (1981) support the relationships between slope and width, and slope and depth

described here. Since slope is traditionally thought to be proportional to bed-load

concentration (Schumm and Khan, 1972), the observed increase in width and decrease in

depth may be related to increasing bed-load concentration within the channel. Channels

with high bed-load concentrations are typically reported to be wide and shallow

compared to low bed-load concentration channels, which are narrow and deep (Leopold

and Maddock, 1953; Schumm and Khan, 1972). Furthermore, since sand-sized particles

can be more easily transported in suspension than gravel, the positive and inverse

relationships grain size has with relative width and depth respectively may be indicative

of increasing bed-load transport relative to the total sediment load.

Since channel slope increases between sites 3 and 5, based on Figure F.5c and d,

the relative channel width should increase while the relative depth should decrease.

Given width and depth increase with discharge downstream, the increase in channel slope

should magnify the downstream increase in channel width and counteract the increase in

depth. This phenomenon broadly explains why relative channel width increases

downstream of Site 3 (Fig. F.5a), while channel depth decreases faster than predicted

(Fig. F.5b).

F.2.2 Width-Depth Ratio and Sinuosity

Dividing the width and depth relations of Millar (2005), the width-depth ratio can

be expressed as,

0.13 -0.33 2.28 -2.88bf 50w:d Q D C* μ’ F.7

where C* is the concentration of bed-load sediment discharge per unit water discharge.

From this relation, bed-load sediment concentration and relative bank-strength exert the

Page 183: Downstream Variability of Fluvial Form, Process, and Character in a

172

greatest control over the width-depth ratio. By contrast, bank-full discharge and median

bed-material size are relatively ineffective. The implication of this relation is that as bed-

load concentration increases the width-depth ratio will increase, but can be impeded by

increasingly resistant banks, which is consistent with past studies (Schumm, 1963;

Schumm and Khan, 1972; Eaton and Church, 2004; Eaton and Church, 2009). While

there is no physical relation describing the controls on sinuosity, based on the work of

Schumm (1963) and Schumm and Khan (1972), it is probable that sinuosity is controlled

in much the same way as width-depth ratio is.

The degree to which the bed-material composition can be divided into bed-load

and suspended-load sediment is presented in Table F.1, and was calculated using the

dimensionless shear stress (τ*) given as,

* h C

s

ρgR Sτ =

(ρ - ρ)gD F.8

where ρ is the density of water (1000 kgm-3), ρs is the density of sediment (2650 kgm-3),

g is the gravitational acceleration (9.81 ms-2), Rh is the hydraulic radius of the channel

(m), SC is the channel slope (-), and D is a grain size of a given calibre on the channel bed

(m). The initiation of suspended sediment transport is given as a dimensionless shear

stress of 1 (Dade, 2000; Church, 2006) and using the slope and hydraulic radius of each

site, the maximum grain size that can be transported in suspension at bank-full flow is

calculated (Table F.1).

The suspended sediment data is poorly detailed with many of the sites not capable

of transporting any bed material sediment in suspension, which may be a result of the

methods employed to sample the bed material. Bulk sampling was undertaken at sites 1

and 3, which tends to emphasize the fine tail of the distribution while the remaining sites

Page 184: Downstream Variability of Fluvial Form, Process, and Character in a

173

Site Max. Grain Size Transported in Suspension (mm)

% of Grain Size Distribution

Width-Depth Ratio

Sinuosity

1 0.07 5 8.31 1.71 3 0.10 5 8.81 1.70 2 2.49 0 14.55 1.14

4u 0.68 0 18.30 1.15 4d 1.72 0 18.72 1.18 5 2.71 0 24.89 1.17 6 1.59 0 25.29 1.26

Table F.1 - Percent of the bed material of the study sites mobilized in suspension compared to the width-depth ratio and sinuosity.

were sampled using a pebble count that emphasizes the coarser tail. Nevertheless, there

does appear to be an emerging trend where the two sites with the highest sinuosities and

lowest width-depth ratios do transport some of their bed-material in suspension, albeit a

very low amount. The reason for such a low percentage is that these beds are poorly

sorted and consist of sediments that the river is incompetent to transport. Mobile

sediments represent between 33% and 51% of the grain size distribution at sites 1 and 3

respectively. As a percentage of the total mobile sediment distribution, the sediment in

Table F.2 represents, respectively for the two sites, 15% and 10%, with the remainder

being transported as bed-load, though saltation is probably pervasive at these sites due to

the fine-calibre nature of the bed material. The results agree with the notion that narrow,

deep, and sinuous channels transport a greater proportion of their sediment load in

suspension than wide and shallow channels.

To further explain the various width-depth ratios, bank-full bed-load

concentration data were sought. While bed-load data were not collected at the sites,

volumetric discharges can be estimated from a bed-load formulae and related to bank-full

discharge to estimate a bed-load sediment concentration. A modified Meyer-Peter and

Müller equation was used (Wong and Parker, 2006) given as,

Page 185: Downstream Variability of Fluvial Form, Process, and Character in a

174

bi*i 3

i

=q

q(G-1)gD

F.9

* * * 1.6i i Crq =4.93(τ -τ ) F.10

where qi* is the dimensionless unit bed-load transport rate for a given grain size fraction i,

qbi is the bed-load transport rate of a given grain size fraction i per unit channel width

(m2s-1), G is the specific gravity of sediment taken as 2.65, g is the gravitational

acceleration (9.81 ms-2), Di is the grain size of a given grain size fraction i (m), τi* is the

dimensionless shear stress for a given grain size fraction i (Eq. F.8), and τCr* is the critical

dimensionless shear stress taken as 0.03, which is generally regarded as the lower limit

for sediment mobility (Buffington and Montgomery, 1997). Using Equations F.9 and

F.10, the fractional volumes for the D95, D90, D84, D75, D50, D25, D16, D10, and D5 were

calculated and summed to give the total bank-full transport rate per unit channel width.

This, in-turn, was multiplied by the active gravel width to yield bank-full volumetric

transport rates (in m3s-1). Active gravel widths were chosen instead of bank-full widths

since bank-full widths incorporate glacial clays, bare sand exposures, and vegetated

surfaces, in addition to the active gravel surfaces. Thus, including these surfaces would

over-estimate the transport volumes. Bed-load concentrations (logC*) were derived by

dividing the bank-full bed-load discharge by the bank-full discharge (Table F.2). Since

hiding and protrusion effects, which are important in sediment transport were not

considered here, these estimates are considered high.

The results of this analysis are shown in Figure F.6. While there is considerable

scatter within the plot, there appears to be a positive correlation between bed-load

sediment concentration and width-depth ratio. Since bed-load transport is a function of

Page 186: Downstream Variability of Fluvial Form, Process, and Character in a

175

the channel slope, discharge, and sediment inputs from banks and valley sides, the scatter

in Figure F.6 is a reflection of the interactions between these variables.

Site qb Wg Qb Qbf LogC* w:d 1 0.0000224 5.64 0.000126 1.87 -4.17 8.31 3 0.000051 9.11 0.000465 5.57 -4.08 8.81 2 0.008694 9.11 0.079202 8.95 -2.05 14.55

4u 0.0000605 10.11 0.000612 10.59 -4.24 18.30 4d 0.001423 5.99 0.008524 14.02 -3.22 18.72 5 0.006765 10.58 0.071574 20.55 -2.46 24.89 6 0.002904 10.75 0.031218 24.64 -2.90 25.29

Table F.2 - Bed-load concentration (logC*) estimates for each study site compared against width-depth ratios (w:d). qb is the bed-load transport rate per unit width (m2s-1), Wg is the average active gravel width (m), Qb is the bank-full bed-load transport rate (m3s-1), and Qbf is the bank-full discharge (m3s-1).

Figure F.6 – Bank-full width-depth ratios of the study sites as a consequence of bed-load concentration segregated by dominant sediment sources. Sites are labelled for reference.

To reflect the variability in sediment inputs, the data points were segregated by

the dominant source of bed-load sediment. Sites that derive all of their bed-load sediment

from glaciolacustrine clays or gravel-poor tills possess the lowest bed-load

Page 187: Downstream Variability of Fluvial Form, Process, and Character in a

176

concentrations, which is a result of the lack of sediment supply (Site 4d), the lack of

transport capacity manifest as low channel slopes relative to the bed material size (Site

4u), and the size of the bed material (sites 1 and 3). Site 2 is an exception since its

estimated bed-load concentration is comparatively high but still maintains a low width-

depth ratio. This may be a reflection of both the strength of the channels banks, which are

root stabilized, and the bimodal grain size distribution of the channel bed, where much of

the sediment load comes from the finer mode. The reason for the bimodality and why the

bed-load sediment concentration may likely not be an accurate reflection of the channel’s

sediment transport regime is explored in section F.2.4.3.

Site 4u is an interesting case, as it possesses the lowest bed-load concentration,

yet maintains a width-depth ratio similar to that of Site 4d, which has a bed-load

concentration an order of magnitude higher. The speculated reason for this is that the

banks at Site 4u are relatively more erodible than the channel bed, which results in

preferential erosion of the banks with minimal bed-load transport as was suggested in

Chapter Five.

Sites 5 and 6 exhibit the highest width-depth ratios and coincide with the highest

bed-load concentrations after Site 2, which are a result of the high transport capacity (ie.

high slopes versus grain size) of the channel and the coarse glaciofluvial sediments and

gravel-rich tills exposed in the valley sides, which act to increase the active gravel widths

compared to gravel-starved sites upstream (ie. Site 4d). It is speculated that the reason

why Site 6 has a similar width-depth ratio as Site 5 but a smaller bed-load concentration

is that the banks at Site 5 may be stronger due to the presence of glacial clay resisting

further widening compared to the more erodible alluvial boundaries of Site 6.

Page 188: Downstream Variability of Fluvial Form, Process, and Character in a

177

F.2.3 Bank-full Discharge Return Period

The return period of bank-full discharge is a function of the hydrological regime

of the watershed, the bank-full cross-sectional area, and the average bank-full velocity,

which is proportional to channel slope. Figure F.7 shows how return period, using

maximum daily flow values on the partial duration series, varies with the channel slope

and the relative cross-sectional area. Generally, there appears to be a statistically

significant positive relationship between channel slope and return period, which is in

accordance with Williams’ (1978) findings. Though Nanson and Young (1981) did not

explicitly look at return periods in their study, they found that a downstream decrease in

channel slope was in part responsible for a downstream decrease in bank-full discharge,

which, in watersheds where discharge of a given frequency increases downstream, would

result in a downstream decrease in return period as well. This relationship between slope

and return period can be explained as follows: high slopes typically result in high

velocities and high discharges for a given cross-sectional area. Thus, a large magnitude

discharge would need to occur in order to overtop the banks. Such a discharge, however,

would be expected to occur infrequently. Conversely, the low slope present at Site 3 leads

to low velocities and hydraulic damming, which causes the water to backup and over-

flow the banks with higher frequencies. In the case of Site 1, the over-bank flows are

driven by the presence of beaver dams that reduce the water surface slope. Over-bank

flows are common at this site after summer convective storms.

The relationship between relative cross-sectional area and return period is much

weaker, but its importance can be summarized by comparing sites 4d, 5, and 6, where

Site 6 has a larger return period than the other two sites despite having a lower channel

Page 189: Downstream Variability of Fluvial Form, Process, and Character in a

178

slope, and can be attributed to its larger relative cross-sectional area. This suggests that

channels with fully alluvial boundaries (Site 6) may be less prone to overbank flooding

than channels with mixed alluvial-non-alluvial boundaries (sites 4d and 5) due to its

ability to more freely adjust the channel dimensions to accommodate peak flows.

The nature of the hydrological regime plays an important role in determining the

return periods, which are different depending on whether maximum daily or maximum

instantaneous flow values are used. Table F.3 gives the return periods using the

maximum instantaneous flow values and the maximum daily flow values.

Figure F.7 - Return period of bank-full discharge (max. daily) plotted against relative cross-sectional area and channel slope. Sites are labelled for reference.

In all of the sites, the return period for bank-full discharge is either 1.5 years or

less using instantaneous values. This is particularly interesting for Site 2 considering its

return period for maximum daily values is much higher. Javelle et al. (2003) report that

Page 190: Downstream Variability of Fluvial Form, Process, and Character in a

179

rivers within south-central Ontario are dominated by spring floods characterized by

moderate flood peaks that occur over short time periods. Accordingly, flood peaks are

high enough to reach bank-full stage with a high degree of regularity. However, they are

not high enough, nor are they persistent enough, to generate a daily average equivalent to

bank-full discharge (hereafter referred to as sustained bank-full discharge). This

reasoning is less true for sites with low slopes since they reach bank-full frequently

regardless of which set of data is used.

Site Return Period (Maximum Instantaneous) Return Period (Maximum Daily) 1 <0.5 years 0.5 years 2 1.5 years 20 years 3 <0.5 years 0.5 years

4u <1.1 years 1.1 years 4d <1.5 years 2 years 5 <1.5 years 3 years 6 1.5 years 5 years

Table F.3 - Return periods for the study sites (see Appendix D for full derivation of these values).

It is interesting to note the disparity between sites 2 and 6. Both have maximum

instantaneous values of 1.5 years, however, their maximum daily values differ by a factor

of 4, which may have to do with their locations within the watershed. As Site 2 is located

in the upper portion of the watershed, it is limited in the catchment size from which its

floodwaters are supplied. As a result, flood-waves are likely to be very flashy and given

the size of the channel cross-sectional area, sustained bank-full flows are not probable.

Site 6, on the other hand, has a smaller relative cross-sectional area and larger drainage

area, which permit a sustained bank-full discharge to occur more frequently.

F.2.4 Bed Material Size and Sorting

Bed sediment size typically varies with channel slope since channels with high

slopes are competent to carry larger sediment particles while channels with low slopes

Page 191: Downstream Variability of Fluvial Form, Process, and Character in a

180

are able to only transport small sediment particles. Grain size and sorting data were re-

plotted against channel slope and segregated by the dominant non-alluvial sediment

sources at each site (Fig. F.8). In general, grain size tends to correlate well with channel

slope and the expected direct relationship is apparent in all of the size fractions. However,

scatter is visible within each fraction, and there is no apparent relationship with channel

slope on slopes above 0.001, indicative of glacial conditioning of the bed material.

F.2.4.1 Size-Selective Transport of the Bed Material

The direct relationship between channel slope and grain size is predicated on the

notion that channels of a given slope can transport the majority of the sediment along the

channel bed. Theoretically, any increase in channel slope will result in an increase in

transport capacity and coarsening of the grain size composition of the channel bed. This

idea is tested on the data to see how effective size selective transport is as an explanation

for the observed relationships in Figure F.8.

Using Equation F.8 and a dimensionless shear stress of 0.03, the maximum grain

size entrainable at each site were calculated and are presented in Table F.4. At the various

study sites, the channel is competent to transport between 10% and 87% of the bed

material at bank-full flow. However, these are estimates, as hiding and protrusion effects

have not been taken into account. Sites 2, 5, and 6 are the most competent sites capable of

transporting over three-quarters of the bed material. The remaining bed sediment not

mobilized during bank-full flows can be considered to be glacially inherited lag deposits.

From a sediment transport perspective, those sites with the most mobile sediment can be

thought of as the least affected by the last glaciation.

Page 192: Downstream Variability of Fluvial Form, Process, and Character in a

181

Figure F.8 - Sorting and grain sizes of various fractions of the bed sediment plotted against channel slope. Datapoints are segregated based on dominant sediment sources and site numbers are given for reference.

Page 193: Downstream Variability of Fluvial Form, Process, and Character in a

182

Site Max. Grain Size Transported (mm) % of Grain Size Distribution 1 2.21 33 2 83.08 79 3 3.42 51

4u 22.60 10 4d 57.28 51 5 90.33 77 6 53.08 87

Table F.4 - Percent of the bed material of the study sites potentially mobilized during bank-full flows.

Of particular note is Site 6, which transports the most amount of sediment of the

study sites and nearly all of its bed material. Figure F.8a-d show that the coarse fractions

at Site 6 are comparatively finer than the other gravel-cobble based sites, with the

disparity increasing with the size fraction. In addition, between sites 5 and 6 there is a

downstream decrease in the sorting coefficient coinciding with a downstream decrease in

channel slope. These two patterns are consistent with conventional size-selective

transport models (Knighton, 1980; Brierley and Hickin, 1985) and it is proposed that this

site has derived its bed-load sediment almost exclusively from downstream transport with

little non-alluvial sediment input or reworking of previously laid-down deposits.

In support of this, the very recent radiocarbon date retrieved for this site (Table

5.4 and Fig. 5.12a) and the thick sand deposits underlying the present channel suggest

that gravel did not exist at the current channel elevation several hundred years ago, which

means that the gravel had to have been transported in from upstream within this recent

time period or supplied directly from the valley sides. Since valley slope-channel

coupling is not extensive and the valley slopes are stable and forested, it is unlikely that

much of the gravel deposits have been derived from them. Size-selective transport

remains the best explanation for the grain size characteristics at Site 6.

Page 194: Downstream Variability of Fluvial Form, Process, and Character in a

183

F.2.4.2 Glacial Inheritance of the Bed Material

The remaining sites are reworking previously laid-down deposits derived from

incision into glacial sediments along the channel length, which account for much of the

variance within Figure F.8a-f. Following the data in Table F.4 and the preceding

discussion on size-selective transport, it is interesting to compare the sorting values and

the channel competence between sites 4d, 4u, and 6. The sorting values for these sites are

respectively 0.78, 0.94, and 0.82, which show a high degree of similarity, especially

between sites 4d and 6 (Fig. F.8g). However, their capacities to move the bed material are

quite different. As mentioned, Site 6 is capable of moving 87% of the bed material at

bank-full discharge. Sites 4u and 4d, on the other hand, can only transport approximately

10% and 51% respectively. It is speculated that the vast disparity between competence

with relatively similar sorting values between sites 4 and 6 may be indicative of a past

discharge regime at Site 4, which would necessarily be larger than present such that the

channel could transport most of the bed material. Paleochannels with discharges 8-10

times larger than present have been identified in studies in southern Ontario (Phillips and

Robert, 2005) and Michigan (Arbogast et al., 2008), while large paleochannels were

observed in the upper Little Rouge watershed. Subsequent waning discharge with climate

change appears to have led to a largely immobile channel bed.

Compared to the other gravel-cobble based sites, Site 5 is coarser in terms of the

Dmax and the D95 and D84 fractions, which result in a higher sorting coefficient (Fig. F.8a-

c, and g). This likely results from the channel receiving sediment from gravel-rich glacial

outwash deposits exposed in the valley sides (Mahaney and Hancock, 1993) while the

Page 195: Downstream Variability of Fluvial Form, Process, and Character in a

184

other sites derive their gravel from gravel-poor silty-clayey tills and glaciolacustrine silts

(Fig. F.9).

The high sorting coefficient at Site 3 results from the low channel slope, which

forces deposition of sand previously passed through the channel at higher slopes. These

sands are mixed with gravel, which account for 57% of the bed material by weight,

deposited from lateral inputs and left over from more competent prior channel states. It is

this combination of sand and gravel that results in the inverse relationship between

sorting and channel slope, counter to what is typically expected in conventional

downstream fining models.

F.2.4.3 External Controls on the Bed Material

On top of the glacially inherited sediment, sites 1 and 2 experience external

controls that dictate the grain size distribution of the bed material. In the case of Site 1,

the presence of beaver dams reduces the water surface slope of the bank-full channel to

approximately 0.0002 (the channel bed slope is an order of magnitude higher) causing

sand to be trapped and deposited out over gravel and cobbles inherited from the pre-dam

channel state. The result is the same as Site 3, where the mixed composition of the

channel bed leads to a very high sorting coefficient and the inheritance of a grain size

distribution from a prior channel state leads to low channel competence.

The high sorting coefficient for Site 2 exists due to bimodality in the grain size

distribution, which is the result of the presence of debris dams along the channel. Several

grain size analyses were done above a debris dam and compared against an analysis done

on a mid-channel bar downstream of the dam (Fig. F.10). Comparing the two

distributions shows that the upstream distribution is strongly bimodal while the

Page 196: Downstream Variability of Fluvial Form, Process, and Character in a

Figure F.9 - Bed material characteristics at each study site in relation to the channel slope and sediment inputs along the channel length. Processes of bed sediment origin are listed and elaborated on in the text. Vertical exaggeration = 78x.

Page 197: Downstream Variability of Fluvial Form, Process, and Character in a

186

downstream distribution is unimodal and far better sorted. The reason for this

phenomenon is the bank-full water surface slope upstream of the debris dam is about 70-

75% lower than that of the channel bed slope, which results in a lower transport capacity.

Using Equation F.8, a critical dimensionless shear stress of 0.03, and inputting the

hydraulic radius (0.57 m) and a slope of 0.002 (based on the water surface slope

measured in the field), the maximum grain size that can be entrained upstream of the

debris dam is 23 mm or 52% of the bed material. Conversely, using a gradient of 0.0072

yields a maximum entrainable grain size of 83 mm or about 95% of the bed material on

the mid-channel bar. A grain size of 23 mm or less represents 16% of the bed material on

the mid-channel bar meaning that there is an excess of this finer material upstream of the

debris dam. The upper half of the upstream grain size distribution is immobile, while the

smaller grain size fractions are brought in from farther upstream and stored on the

channel bed. Overtime, these smaller fractions progressively make up a greater

proportion of the bed material until the debris dam breaks or the channel is diverted and

the water surface slope steepens up flushing out the smaller fractions and remobilizing

the larger fractions. This process of sediment trapping and expulsion has implications for

the bed-load concentration data presented in Figure F.6. The calculation was made using

the combined upstream and downstream grain size distribution for which the upstream

distribution dominates due to more sampling locations, and a slope of 0.0072. In essence,

the datapoint for Site 2 in Figure F.6 represents a rare scenario after the debris dam

breaks or the channel is diverted around it where the excess fine sediment is flushed

through the system. The processes described in the preceding paragraphs and their spatial

distributions are summarized in Figure F.9.

Page 198: Downstream Variability of Fluvial Form, Process, and Character in a

187

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

1 10 100 1000 10000

Grain Size (mm)

Per

cen

t F

iner

Upstream of Debris Dam

Downstream of Debris Dam

With damWithout dam

Figure F.10 - Grain size distributions for Site 2 above (n = 405) and below (n = 128) the debris dam. Red dots represent the limits of competence at bank-full flow. F.2.5 Channel Bed Alluvium Thickness and Exposed Clay Percentage

Massong and Montgomery (2000) have shown that for mountain drainage basins,

alluvial and bedrock reaches can be discriminated through the use of a simple slope-

discharge plot (Fig. F.11). For a given discharge, bedrock channels occur on higher

slopes unless the channel is forced into an alluvial state by wood debris or through high

sediment supply relative to the transporting capacity of the channel. Likewise, a bedrock

channel that falls below the threshold line may be caused by a low sediment supply

relative to the transport capacity. The implication here is that these two substrates are

segregated based on their transport capacities relative to their sediment supply.

Since the dimensionless shear stress is a ratio between the transport capacity of

the channel and the grain size of the channel bed, it was compared against the average

percentage of glacial clay or bedrock exposed in the channel (Fig. F.12) and the average

channel bed alluvium thickness (Fig. F.13) to see if similar trends appear to that reported

by Massong and Montgomery (2000). It is expected that channels with dimensionless

Page 199: Downstream Variability of Fluvial Form, Process, and Character in a

188

shear stresses above critical (0.03) will possess high exposed clay percentages (under

supply-limited conditions) and low alluvium thickness values unless forced into an

alluvial state by excess sediment or external forces (transport-limited conditions). In

determining the dimensionless shear stress (Eq. F.8), the grain size was set as the D50.

Figure F.11 - Theoretical discrimination between alluvial and bedrock channels based on slope and drainage area (after Massong and Montgomery, 2000).

Figures F.12 and F.13 show that the expected trends do not appear to hold, as a

positive trend is observed between both dimensionless shear stress, percent clay exposed,

and alluvium thickness, though the relationships are quite poor and in neither case is the

regression slope significant. For percent clay exposed, the data were subdivided into four

classifications based on sediment mobility and sediment supply conditions. The boundary

between immobile and mobile sediment is given as a dimensionless shear stress value of

0.03 above which sediment is often considered mobile. The boundary between supply-

and transport-limited conditions is arbitrarily defined but given the separation between

sites 1, 4d, and 5 from the other sites, the approximate location of the boundary is

justified. Despite exhibiting up to 5% clay exposure, on average, the lower sites are

considered transport-limited because clay was only exposed in one cross-section at each

Page 200: Downstream Variability of Fluvial Form, Process, and Character in a

189

site. Thus, an alluvial (transport-limited) state was considered the average condition for

these sites.

While the data in Figures F.12 and F.13, in general, do not support the hypothesis

that increasing transport capacity (ie. dimensionless shear stress) results in thinner

channel bed alluvium and more exposed glacial clay, this is a consequence of the diverse

channel conditions and geomorphic histories within the watershed. Consider Site 4,

which is split into two sub-reaches on the basis of differing channel slopes. Within both

sub-reaches, all relevant controlling conditions (glacial inputs and discharge regime) are

held constant with channel slope varied from 0.0017 at the upstream sub-reach (4u) to

0.0043 at the downstream sub-reach (4d). As a consequence of this change in slope, the

Figure F.12 - Average percent clay exposed in the bank-full channel as a function of dimensionless shear stress. Critical dimensionless shear stress of 0.03 is shown as the lower limit of sediment mobility. Sites are given for reference.

Page 201: Downstream Variability of Fluvial Form, Process, and Character in a

190

Figure F.13 - Average channel bed alluvium thickness as a function of dimensionless shear stress.

dimensionless shear stress increases from 0.011 to 0.031 and the channel moves from an

immobile bed state to a mobile state. This increase in transport capacity given a low

sediment supply strips away bed material and increases the percent clay exposed in the

channel from 0.76% to 26% and decreases the average thickness of the channel bed from

37 cm to 26 cm (Figs F.12 and F.13). A t-test revealed that these values are significantly

different at the 90% confidence level.

The role that glacial sediment inputs have on the exposure of clay and channel

bed alluvium thickness is evident when comparing sites 4d and 5. While the channel at

Site 5 has a more mobile bed, the contribution of sediment from outwash gravel deposits

exposed in the valley sides and terraces helps to conceal the glacial clay substrate and

increase the alluvium thickness. Conversely, Site 4d derives its alluvium from a silt-clay

Page 202: Downstream Variability of Fluvial Form, Process, and Character in a

191

till and thus does not have the necessary sediment supply to form a thick alluvial layer,

nor hide as much of the underlying clay.

Despite having a dimensionless shear stress below critical and similar to that of

Site 4u, glacial clay is exposed in 15% of the surveyed channel cross-sections at Site 1. It

is proposed that this site represents a condition where the majority of the bed sediment is

immobile (Table F.4) while possessing a limited supply of bed sediment. As the channel

bed slope is an order of magnitude higher than the present bank-full water surface stage,

the presence of glacial clay may be indicative of a former channel state (and transport

capacity) in existence prior to the initiation of beaver dam construction along the channel,

where the channel may have been in a mobile-bed supply-limited state competent to

transport at least half of its coarse (> 2 mm) bed material.

Sites 2, 3, and 6, despite being competent to transport at least half of the channel

sediment, are transport limited due to a greater influx of sediment than transport capacity.

This is particularly true for Site 3, which, due to a reduced channel slope, and thus

transport capacity, much of the fine sediment derived from upstream settles out,

blanketing the channel bed and concealing much of the exposed clay (see Fig. F.9). In

addition, the newly deposited sand covers relic channel gravels increasing the thickness

of the channel bed alluvium compared to the other study sites (Fig. F.13).

While the same general process of deposition described for Site 3 happens in Site

1, the build up of fine alluvium within the channel is not to the same extent, which may

be due to Site 1’s location farther upstream where fine sediment recruitment is less

available due to the channel’s smaller erosive power upstream and lack of valley sides to

contribute sediment. Conversely, Site 3 is located downstream of a large glaciolacustrine

Page 203: Downstream Variability of Fluvial Form, Process, and Character in a

192

sand deposit that the channel has cut into, which supplies the site with more bed-calibre

sediment (Fig. F.9).

Site 2 is forced into an alluvial state due to the presence of a debris dam. Figure

F.14 shows the effect that the observed debris dam has on the alluvium thickness and

exposure of glacial clay within the channel. Upstream of the dam, fine gravel and

granules are transported in, increasing the thickness of the gravel deposits and smoothing

out the channel bed resulting in a bed structure more comparable to a run than a riffle-

pool morphology. Immediately downstream of the dam, the increased gradient and

turbulence leads to greater bed scouring and, as a result, gravel thickness is reduced and

glacial clay is exposed in the bed. Farther downstream the channel remains fully alluvial

with a thick (approximately 40 cm) layer of gravel, which is thought to be due to the

recurrence period of bank-full discharge. According to the flood frequency analysis

presented in Appendix D, bank-full discharge has an exceedance probability of

approximately 67% in any given year using maximum instantaneous values, which is

comparable to other sites. However, instantaneous measurements only reflect a 15-minute

interval and do not transport much sediment within this time period. While a sustained

bank-full discharge (ie. bank-full discharge averaged over a day) is capable of much

greater sediment transport, it has a probability of occurring in any given year between 5-

10%. Thus, competent discharges necessary to erode into the accumulated sediment are

either too short or too infrequent to expose clay.

Page 204: Downstream Variability of Fluvial Form, Process, and Character in a

193

Figure F.14 - Longitudinal profile of Site 2 showing the effects the debris dam has on the downstream variation in gravel thickness and exposed clay. Clay surface inferred based on selected auger points that reached the surface (beige circles) and points that did not (grey circles). Note the reduced water surface slope upstream of the debris dam compared to the channel bed slope, which results in preferential transport and deposition of fine gravel leading to a bimodal grain size distribution and thick gravel deposits.

Lastly, as previously mentioned, the channel at Site 6 flows over thick sand deposits

such that bedrock is only exposed during channel-valley side coupling. These thick sand

deposits account for the high datapoint for channel bed alluvium thickness shown in

Figure F.13.

F.3 Floodplain Sedimentology

F.3.1 Overbank Alluvium Thickness and Silt/Clay Content

Overbank alluvium thickness and the silt content of floodplains are typically

thought to increase downstream as sediment is conveyed from upstream, high gradient

source regions and deposited along low gradient reaches (Church, 2002). The

relationships between channel slope, overbank alluvium thickness, and the silt content are

tested and shown in Figures F.15 and F.16. There appears to be good agreement in the

two trends as the floodplain tends to get finer and thicker as channel slope decreases,

which conforms to conventional knowledge that increasingly erosive overbank flows

Page 205: Downstream Variability of Fluvial Form, Process, and Character in a

194

prevent deposition of fine sediment. Overbank alluvium thickness tends to show far more

scatter than the silt content though, suggesting that factors other than channel slope are

important in determining the thickness of overbank deposits. One such factor is the

sediment supply from upstream. While fine-grained glacial sediments are contributed to

the channel along much of its length, the amount of fine sediment is determined by the

degree to which the channel is incised into the glacial sediments and the type of sediment

the channel is incised into. Incision exposes more surface area for weathering and erosion

and the steep valley slopes transfer the sediment to the valley floor. Moreover, incised

channels often experience more coupling with the valley sides than channels that are not

incised. As a result, sediment can be directly transferred to the channel instead of sitting

on the valley floor.

In the context of the Little Rouge River, the channel is incised into glacial

deposits by up to 30-40 metres in its lower 20 kilometres. The middle reach (10-25 km) is

still incised, but to a more limited degree, while the upstream-most 10 km are not incised

at all (Fig. F.17). The impact of this is that fine sediment production is limited in the

upper watershed and active in the lower watershed. This explains the difference in

sediment storage between sites 1 and 3. Site 3 is downstream of glaciolacustrine sand

deposits that the channel is cut into while Site 1 is downstream of glaciolacustrine clay

deposits that the channel flows over but is not directly cut into. Moreover, as the erosive

potential of the channel is lower upstream than it is downstream, the ability to recruit

sediment is further hindered above Site 1.

The importance of sediment supply is again demonstrated when comparing sites 4

and 6. Site 6 is downstream of a highly energetic and deeply incised channel cut into sand

Page 206: Downstream Variability of Fluvial Form, Process, and Character in a

195

diamicts (Fig. F.17). As a result, sediment is recruited from the valley sides and

transported downstream with little storage on the floodplain (which explains the thin

nature of Site 5) and deposited at Site 6. Conversely, Site 4 is downstream of a major

sediment sink in Site 3 where fine sediment is prevented from moving farther

downstream. Consequently, starvation of sediment leads to thinner overbank deposits at

Site 4.

Figure F.15 – Average floodplain overbank alluvium thickness as a function of channel slope.

Figure F.16 - Depth-averaged silt/clay content of the floodplain as a function of channel slope.

Page 207: Downstream Variability of Fluvial Form, Process, and Character in a

Figure F.17 - Spatial variation of floodplain sedimentological characteristics and their relation to the longitudinal profile and glacial sediment sources. See text for discussion. Vertical exaggeration = 78x.

Page 208: Downstream Variability of Fluvial Form, Process, and Character in a

197

It is interesting to note that despite possessing a channel slope an order of

magnitude higher, Site 6 has an overbank thickness comparable to that of Site 3. This is

likely caused by a combination of greater sediment supply from the valley sides upstream

of Site 6 and a greater efficiency of downstream sediment conveyance towards Site 6.

Lastly, despite having similar channel slopes, Site 2 is significantly thicker and

finer than Site 5 at the 99% confidence interval. Jeffries et al. (2003) describe a case

where sediment deposition on the floodplain of a low-order gravel-bed stream is strongly

conditioned by the presence of in-channel debris dams. They observe that impoundment

of water behind debris dams forces much of the would-be channelized discharge onto the

floodplain. Suspended sediment is then transferred onto the floodplain through diffusion

and deposited due to low overbank velocities. The frequency of overbank flows, and thus

sediment deposition, increases upstream of these dams as well. Jeffries et al. (2003) note

that during the period of their study, overbank flows occurred 41.5% of the time upstream

of the dam, while only occurring 0.22% of the time downstream. It is suspected that Site

2 may behave in much the same manner. Indeed, since debris dams force low-energy

conditions upstream by ponding water, assuming a bank-full water surface slope of 0.002

upstream of the debris dam (based on survey measurements) is an appropriate value to

characterize upstream energy conditions, the observed overbank alluvium thickness and

the silt/clay content of Site 2 plot much closer to the trendlines in Figures F.15 and F.16.

Given that sites 1 and 2 both possess similar sediment supplies due to their close

proximity and both possess woody dams spanning their channels, why their overbank

alluvium thickness differ so dramatically remains uncertain. It is speculated that some of

the sediment, which is considered overbank alluvium at Site 2, is stored on the channel

Page 209: Downstream Variability of Fluvial Form, Process, and Character in a

198

bed at Site 1. The average thickness of the sub-2mm sequence at Site 1 is 99 cm, which is

not statistically different from the overbank alluvium thickness at Site 2 at the 95%

confidence level.

F.3.2 Floodplain Age

Absolute ages of floodplain surfaces could not be attained for all sites due to the

lack of suitable dating material available. As a substitute, relative ages are proposed

based on depth-averaged percent organics sampled from multiple points within the

floodplain. The premise behind this is that as soils increase in age and become more

mature, the organic content increases as well such that older floodplain surfaces will

contain greater organic levels (Mahaney and Sanmugadas, 1986). While this method is

very limited in terms of dating control, it can help elucidate relative trends in floodplain

age, and thus channel stability. Figure F.18 shows organic content against channel slope

and Figure F.17 divides the Little Rouge River into two zones with sites 5 and 6

potentially being significantly younger than the other sites. This inference is supported by

statistical tests and two radiocarbon dates retrieved from sites 3 and 6. An analysis of

variance and subsequent Fisher’s Least Significance Difference post-hoc test reveals that

the data can be partitioned into three groups. Sites 2, 3, and 4, and 5 and 6 represent two

distinct groups while Site 1 represents a third group. The radiocarbon dates suggest that

the floodplain at Site 4u and upstream may be at least 3100 years old, while sites 5 and 6

are modern in age (Fig. F.17). The value for Site 4d is over-estimated because of mixing

between old low-lying terraces (organic content of floodplain surface between 6% and

8%, comparable to Site 4u in age) and newer inset floodplain surfaces (organic content =

Page 210: Downstream Variability of Fluvial Form, Process, and Character in a

199

2-4%, comparable to Site 6 in age if not older). Whether Site 1, which possesses the

highest organic content, is significantly older than Site 3 is unknown.

Since the radiocarbon dates were acquired from the outer bends of the channel,

the dates can be thought to represent the residence period of floodplain sediments.

Floodplain sediments at, and upstream of Site 4u appear to be recycled more than an

order of magnitude slower than sites 5 and 6, which imply a more laterally active channel

downstream of Site 4 and a very stable channel upstream.

Organic content was plotted against channel slope demonstrating a weak inverse

relationship (Fig. F.18). The results suggest energetic channels are more capable of

floodplain turnover, thus being younger, than less energetic channels. Brown (1990) drew

similar conclusions. It is thought that the reason Site 2 plots much higher (ie. more stable)

than Site 5 given similar slopes is a combination of low total stream power due to low

discharge and debris dams reducing the water surface slope, as well as root-stabilized

banks that resist floodplain erosion.

Figure F.18 - Depth-averaged organic content of the floodplain in relation to channel slope. Sites are given and uncalibrated radiocarbon dates are attached for reference.

Page 211: Downstream Variability of Fluvial Form, Process, and Character in a

200

Appendix G: Statistical Tests

Sum of Squares df Mean Square F Sig. Level Between groups 1188.348 6 198.058 16.435 0.000 Within groups 530.232 44 12.051

Total 1718.580 50

Table G.1 - Analysis of variance test results of differences in bank-full channel width.

Site Site Significance Level

1 2 0.041 3 0.071 4u 0.001 4d 0.000 5 0.000 6 0.000

2 3 0.855 4u 0.060 4d 0.031 5 0.000 6 0.000

3 4u 0.051 4d 0.027 5 0.000 6 0.000

4u 4d 0.909 5 0.051 6 0.002

4d 5 0.048 6 0.002

5 6 0.075

Table G.2 - Fisher's least significance difference post-hoc test results on bank-full channel width.

Sum of Squares df Mean Square F Sig. Level

Between groups 0.846 6 0.141 3.910 0.003 Within groups 1.587 44 0.036

Total 2.433 50

Table G.3 - Analysis of variance test results of differences in bank-full channel depth.

Site Site Significance Level

1 2 0.746 3 0.001 4u 0.671 4d 0.678 5 0.837 6 0.220

2 3 0.000

Page 212: Downstream Variability of Fluvial Form, Process, and Character in a

201

4u 0.456 4d 0.446 5 0.534 6 0.088

3 4u 0.012 4d 0.006 5 0.000 6 0.026

4u 4d 0.972 5 0.771 6 0.511

4d 5 0.786 6 0.457

5 6 0.231

Table G.4 - Fisher's least significance difference post-hoc test results on bank-full channel depth.

Sum of Squares df Mean Square F Sig. Level

Between groups 690.748 6 115.125 9.395 0.000 Within groups 539.194 44 12.254

Total 1718.580 50

Table G.5 - Analysis of variance test results of differences in bank-full channel area.

Site Site Significance Level

1 2 0.227 3 0.003 4u 0.012 4d 0.005 5 0.000 6 0.000

2 3 0.032 4u 0.080 4d 0.042 5 0.000 6 0.000

3 4u 0.990 4d 0.865 5 0.155 6 0.001

4u 4d 0.895 5 0.261 6 0.006

4d 5 0.292 6 0.005

5 6 0.022

Table G.6 - Fisher's least significance difference post-hoc test results on bank-full channel area.

Page 213: Downstream Variability of Fluvial Form, Process, and Character in a

202

t-statistic df Significance Level 4.293 319 0.000

Table G.7 - T-test results of differences in channel bed alluvium thickness between sites 4u and 4d.

t-statistic df Significance Level

-2.638 4.051 0.057

Table G.8 - T-test results of differences in exposed clay percentage between sites 4u and 4d.

t-statistic df Significance Level

7.761 15 0.000

Table G.9 - T-test results of differences in overbank alluvium thickness between sites 2 and 5.

t-statistic df Significance Level

3.259 13 0.006

Table G.10 - T-test results of differences in silt/clay percentage between sites 2 and 5.

t-statistic df Significance Level

1.854 10 0.093

Table G.11 - T-test results of differences in sub-2 mm floodplain alluvium thickness between sites 1 and 2.

Sum of Squares df Mean Square F Sig. Level

Between groups 157.850 6 26.308 5.953 0.000 Within groups 212.133 48 4.419

Total 369.982 54

Table G.12 - Analysis of variance test results of differences in organic content.

Site Site Significance Level

1 2 0.090 3 0.026 4u 0.045 4d 0.004 5 0.000 6 0.000

2 3 0.574 4u 0.703 4d 0.236 5 0.004 6 0.002

3 4u 0.871 4d 0.540 5 0.017 6 0.011

Page 214: Downstream Variability of Fluvial Form, Process, and Character in a

203

4u 4d 0.450 5 0.014 6 0.009

4d 5 0.055 6 0.042

5 6 0.920

Table G.13 - Fisher's least significance difference post-hoc test results on organic content.

Page 215: Downstream Variability of Fluvial Form, Process, and Character in a

204

Curriculum Vitae

Name James Benjamin Thayer Place of Birth Newmarket, Ontario Year of Birth 1988 Post-Secondary Education 2010-2012 The University of Toronto Master of Science in Geography

2006-2010 The University of Western Ontario Honours Bachelor of Science in Geography and Geology

Honours and Awards 2011 John Horner Graduate Scholarship in Geography

2010 The University of Toronto Fellowship Award 2010 The Canadian Association of Geographers Prize 2010 Certificate of Merit of Excellence 2010 NSERC USRA 2010 Dean’s Honour List 2009 NSERC USRA 2009 Dean’s Honour List 2008 NSERC USRA 2008 Dean’s Honour List 2007 Dean’s Honour List 2006 Western Scholarship of Distinction

Related Work Experience September 2010 – May 2012

Research Assistant to Dr. Joe Desloges at The University of Toronto

September 2010 – May 2012 Teaching Assistant at The University of Toronto May 2010 – September 2010

Research Assistant to Dr. Peter Ashmore at The University of Western Ontario

May 2009 – September 2009

Research Assistant to Dr. Peter Ashmore at The University of Western Ontario

May 2008 – September 2008

Research Assistant to Dr. Peter Ashmore at The University of Western Ontario

Page 216: Downstream Variability of Fluvial Form, Process, and Character in a

205

Publications & Thayer, J.B. (2010). Floodplain development for a semi- Conference Abstracts alluvial river. Canadian Association of Geographers

Ontario Division, 2010 Annual Meeting, October 15-16, 2010. Ryerson University, Toronto, Canada.

Thayer, J.B. (2010). Floodplain development for a semi- alluvial river. Undergraduate thesis manuscript. The University of Western Ontario, 73 pp.