Transcript
Page 1: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

Floodplain Hydrology and Biogeochemistry

Charles N. Jones

Dissertation submitted to the faculty of the Virginia Polytechnic Institute and State University in

partial fulfillment of the requirements for the degree of

Doctor of Philosophy

in

Biological Systems Engineering

Durelle T. Scott, Chair

W. Cully Hession

Erich T. Hester

Richard F. Keim

David J. Sample

July 30, 2015

Blacksburg, Virginia

Keywords: river-floodplain connectivity, inundation hydrology, solute fate and transport, ecological

engineering, biogeochemistry, nutrients, dissolved organic matter, stream restoration

Copyright © 2015 C. Nathan Jones

Page 2: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

FLOODPLAIN HYDROLOGY AND BIOGEOCHEMISTRY

C. Nathan Jones

ABSTRACT

River-floodplain connectivity is defined as the water mediated transfer of materials and energy

between a river or stream and its adjacent floodplain. It is generally accepted that restoring and/or

enhancing river-floodplain connectivity can reduce the downstream flux of reactive solutes such

as nitrogen (N) and phosphorus (P) and thus improve downstream water quality. However, there

is little scientific literature to guide ecological engineering efforts which optimize river-floodplain

connectivity for solute retention. Therefore, the aim of my dissertation research was to examine

feedbacks between inundation hydrology and floodplain biogeochemistry, with an emphasis on

analyzing variation experienced along the river continuum and the cumulative effects of river-

floodplain connectivity at the basin scale. This was completed through four independent

investigations. Field sites ranged from the Atchafalaya River Basin, the largest river-floodplain

system in the continental US, to the floodplain of a recently restored headwater stream in

Appalachia. We also developed a method to examine river-floodplain connectivity across large-

river networks and applied that methodology to US stream network. Largely, our results highlight

the role floodwater residence time distributions play in floodplain biogeochemistry. In headwater

streams, residence times restrict redox dependent processes (e.g. denitrification) and downstream

flushing of reactive solutes is the dominant process. However, in large-river floodplains, redox

dependent processes can become solute limited because of prolonged residence times and

hydrologic isolation. In these floodplains, the dominant process is often autochthonous solute

accumulation. Further, results from our modeling study suggest large-river floodplains have a

greater impact on downstream water quality than floodplains associated with smaller streams, even

when considering cumulative effects across the entire river network.

Page 3: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

iii

To Mom and Dad. This work would not have been possible without your love, support, and

occasional (but well deserved) kick-in-the-ass.

Page 4: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

iv

ACKNOWLEDGMENTS

While the African proverb, “It takes a whole village to raise a [gradstudent],” is overly quoted and

probably cliché, it accurately captures my graduate experience at Virginia Tech. I am very thankful

for the comradery and collaborations I experienced in the greater water science community, the

BSE family, and the Hydroecology Lab over the last five years. I would like to think this degree

is not just a piece of paper, but the sum total of our shared experiences and memories.

Specifically, I would like to thank my adviser Dr. Durelle “Scotty” Scott for his patience, shared

insight, and guidance. He continues to freely and enthusiastically share his ideas, collaborations,

and time with our lab group. I’ve been truly blessed to have him as an adviser, and hope to provide

the same caliber of mentoring to my future students.

I would also like to thank my dissertation committee and other collaborators:

Forest and Wetland Ecohydrology Lab: Drs. Richard Keim (Committee Member) and

Brandon Edwards have been incredibly supportive and generous with their ideas,

resources, and time. Both suffered through edits on my first manuscript, and at the very

minimum, I owe their lab one boat anchor.

Environmental Hydraulics Research Group: Dr. Erich Hester (Committee Member) and

Chris Guth were both critical to the execution and success of the StREAM Lab flood

experiments, and Erich’s insights in surface water-groundwater connectivity have been

invaluable to my understanding of river systems.

USGS Hydroecology of Flowing Water: Drs. Jud Harvey and Jesus Gomez-Velez have

been incredible resources and set a very high standard in our recent modeling work. I truly

appreciate their patience and look forward to future collaborations.

VT-BSE: Dr. Cully Hession (Committee Member) has been an integral part of my

education here at tech. He recruited me, was a co-advisor for my first year, and provided

the opportunity to conduct a pilot study of our StREAM Lab floodplain experiment through

the REU program. Dr. David Sample (Committee Member) has also been an integral part

of my experience at tech, providing insight into modeling studies and resources to conduct

those studies.

This list wouldn’t be complete if I didn’t thank the BSE Staff members: Denton Yoder, Laura

Lehman, and Kelly Peeler. None of this work would have been possible without these individual’s

expertise, sense of humor, and hard work.

Finally, I would like to thank my friends and family for their support and prayers these last five

years. Not many people have the support system I do, and during my roughest days, I could always

find someone to provide perspective and encouragement. Also, I want to acknowledge my

beautiful wife, Virginia Jones, for her part in this work. She’s listened to me talk about floodplains

and river networks for hours, proof read most of my papers, and continues to be an amazing friend

and wife.

Page 5: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

v

TABLE OF CONTENTS

Table of Contents ............................................................................................................................ v

List of Tables ............................................................................................................................... viii

List of Figures ................................................................................................................................ ix

Chapter 1. Introduction ................................................................................................................... 1

1.1 Food, Water, and Energy Nexus ........................................................................................... 1

1.2 Riverine Systems and Floodplains ........................................................................................ 2

1.3 Ecological Engineering and River-Floodplain Connectivity ................................................ 2

1.4 Works Cited .......................................................................................................................... 4

Chapter 2. Perirheic mixing and biogeochemical processing in flow-through and backwater

floodplain wetlands ......................................................................................................................... 8

2.1 Introduction ........................................................................................................................... 8

2.2 Methods ................................................................................................................................ 9

2.2.1 Site Description .............................................................................................................. 9

2.2.2 Synoptic Sampling Design and Analysis ..................................................................... 11

2.2.3 Residence Time Conceptualization.............................................................................. 12

2.2.4 Statistical Analysis ....................................................................................................... 12

2.3 Results ................................................................................................................................. 14

2.3.1 Ordination Analysis ..................................................................................................... 14

2.3.2 Cluster Analysis of Samples ........................................................................................ 15

2.3.3 Comparison between the Connected and Disconnected Classes ................................. 16

2.4 Discussion ........................................................................................................................... 17

2.4.1 Conceptual Model ........................................................................................................ 17

2.4.2 Classification of Hydrologically Connected and Disconnected Locations .................. 18

2.4.3 Carbon, Nitrogen, and Phosphorus dynamics in the ARB........................................... 19

2.4.4 Temporal and Spatial Variability in Floodplain Processes .......................................... 20

2.4.5 Management Implications ............................................................................................ 20

2.5 Conclusions ......................................................................................................................... 21

2.6 Acknowledgements ............................................................................................................. 21

2.7 Works Cited ........................................................................................................................ 21

Page 6: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

vi

Chapter 3. Biogeochemical processing of dissolved organic matter and nitrogen in floodwaters

across a gradient of river-floodplain connectivity within the Atchafalaya River floodplain ....... 26

3.1 Introduction ......................................................................................................................... 26

3.2 Methods .............................................................................................................................. 27

3.2.1 Site Description ............................................................................................................ 27

3.2.2 Synoptic Sampling Design and Analysis ..................................................................... 30

3.3. Results ................................................................................................................................ 31

3.3.1 Hydrologic Tracer of Water ......................................................................................... 31

3.3.2 Reactive Solute Characterization ................................................................................. 32

3.3.3 DOM Characterization ................................................................................................. 32

3.4 Discussion ........................................................................................................................... 36

3.4.1 River-floodplain connectivity ...................................................................................... 36

3.4.2 Biogeochemical Processing of Nitrogen within the Floodplain .................................. 36

3.4.3 Temporal variation in Dissolved Organic Matter ........................................................ 38

3.5. Conclusion ......................................................................................................................... 39

3.6 Works Cited ........................................................................................................................ 39

Chapter 4. Seasonal Variation in Floodplain Biogeochemical Processing in a Restored

Headwater Stream ......................................................................................................................... 45

4.1 Introduction ......................................................................................................................... 46

4.2 Methods .............................................................................................................................. 48

4.2.1 Site Description ............................................................................................................ 48

4.2.2 Experimental Design .................................................................................................... 49

4.2.3 Flood Experiment Analyses ......................................................................................... 49

4.2.4 Reach-Scale Modeling ................................................................................................. 50

4.3 Results and Discussion ....................................................................................................... 50

4.3.1 Inundation Hydrology .................................................................................................. 50

4.3.2 Dissolved Organic Matter Export ................................................................................ 53

4.3.3 Initial Flush of Reactive Solutes .................................................................................. 53

4.3.4 Nitrate Removal ........................................................................................................... 55

4.3.5 Modeled Load Reductions ........................................................................................... 57

4.3.6. Management Implications ........................................................................................... 58

Page 7: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

vii

4.4 Acknowledgements ............................................................................................................. 58

4.5 Works Cited ........................................................................................................................ 58

Chapter S4. Supplemental Information ......................................................................................... 65

S4.1 Climatic Data Acquisition ................................................................................................ 65

S4.2 Water Quality Sample Handling and Analysis ................................................................. 65

S4.3 Estimates of Reach-Scale Load Reductions ..................................................................... 65

S4.4 Works Cited ...................................................................................................................... 69

Chapter 5. The effect of hydrogeomorphic gradients and hydroclimatic setting on river-

floodplain connectivity at the continental scale ............................................................................ 71

5.1 Introduction ......................................................................................................................... 71

5.2 Methods .............................................................................................................................. 73

5.3 Results and Discussion ....................................................................................................... 76

5.3.1 Heterogeneity of River-Floodplain Connectivity across USGS Gaging Stations ....... 76

5.3.2 Effect of Catchment Area on River-Floodplain Connectivity ..................................... 78

5.3.3 Effect of Physiography on River-Floodplain Connectivity ......................................... 80

5.4 Conclusion .......................................................................................................................... 80

5.5 Works Cited ........................................................................................................................ 81

Chapter S5. Supplemental Material .............................................................................................. 85

S5.1 Data Acquisition ............................................................................................................... 85

S5.2 Analysis of USGS Gage Data ........................................................................................... 87

S5.2.1 Identifying the threshold for connectivity ................................................................. 87

S5.2.2 Estimation of Sub-daily hydrographs ........................................................................ 88

S5.2.3 Characterization of river-floodplain connectivity ..................................................... 90

S5.3 Application to the National Stream Network ................................................................... 95

S5.4 Works Cited ...................................................................................................................... 98

Chapter 6. Concluding remarks and questions for future study.................................................. 102

6.1 Works Cited ...................................................................................................................... 104

Page 8: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

viii

LIST OF TABLES

Table 4-1. Summary of residence time metrics ............................................................................ 50

Table 4-2. Summary of Spiraling metrics and associated statistics from tracer addition. ........... 56

Table S4-1. Regional Regression Coefficient Values .................................................................. 67

Table S4-2. Log Pearson Type 3 Parameters ............................................................................... 67

Table S5- 1. Main variables used in analysis ............................................................................... 86

Table S5-2. Median Flood Volume Ratio by Stream Order and Physiographic Division .......... 94

Table S5-3. Median Storage Index by Stream Order and Physiographic Division ..................... 95

Table S5-4. Spearman Correlation (RS) between FVRand Stream Order ................................... 95

Table S5-5. Estimated annual floodplain storage......................................................................... 96

Table S5-6. Cumulative floodplain storage and length weighted mean floodplain storage ........ 96

Table S5-7. Cumulative length normalized floodplain storage.................................................... 97

Page 9: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

ix

LIST OF FIGURES

Figure 2-1. Map of the Southern Atchafalaya River Basin (ARB) .............................................. 10

Figure 2-2. Constituent concentrations in ordination space ......................................................... 13

Figure 2-3. Evolution of the floodwater chemistry over the course of the annual flood pulse .... 14

Figure 2-4. Separation of Connected and Disconnected sites ...................................................... 15

Figure 2-5. Box plots of nitrogen speciation and the DOC:NO3- ratios throughout the flood ... 16

Figure 2-6. Relationship between A.) DOC and δ18O and B.) Fe and SRP ............................... 17

Figure 2-7. Conceptual sketch ..................................................................................................... 18

Figure 3-1. Map displaying the sampling locations ..................................................................... 28

Figure 3-2. Annual Hydrograph and River-Floodplain Hysteresis Plot ......................................29

Figure 3-3. Isotopic content of samples collected over the course of the flood .......................... 32

Figure 3-4. NO3-, TKN, and DOC:TKN ratio over the course of the flood ................................. 33

Figure 3-5. Dissolved carbon concentration and isotopic content across the flood. .................... 34

Figure 3-6. Spectral Indicies of dissolved organic matter (DOM) across the flood pulse ........... 35

Figure 3-7. Byplots showing relationships between reactive solutes. ......................................... 36

Figure 3-8. Conceptual model of swamp channel hydrology and biogeochemistry. ................... 37

Figure 4-1. Conceptual model of processes within the experimental floodplain slough. ............ 47

Figure 4-2. Climatic data measured in the StREAM Lab floodplain. .......................................... 48

Figure 4-3. Inlet and outlet concentrations of dissolved reactive constituents. ........................... 52

Figure 4-4. The total load observed at the inlet (dark) and outlet (light) of the floodplain. ........ 54

Figure 4-5. Areal NO3- uptake ..................................................................................................... 57

Figure S4-1. Map of estimated inundation and flow-inundation relationship ............................. 67

Figure S4-2. Parameterized Flow Duration Curve (FDC) at the StREAM Lab watershed. ........ 68

Figure 5-1. Conceptual model displaying the threshold of connectivity ..................................... 73

Figure 5-2. Flood volume ratio and storage index approximations. ............................................ 76

Figure 5-3. Floodplain storage volume ........................................................................................ 78

Figure 5-4. The summation of length normalized floodplain storage volume............................. 80

Figure S5-1. Relationship between QBKF* and catchment area for 5,981 gaging stations ........... 88

Figure S5-2. Illustration of the Sangal [1983] approximation. .................................................... 84

Figure S5-3. Strong relationship experienced between QIPF and QMDF at Colombia River ......... 90

Figure S5-4. Annual hydrograph from Boulder Creek near Boulder ....................................................... 90

Figure S5-5. Estimate of UBKF at USGS gaging stations ............................................................. 92

Figure S5-6. Conceptualization of the Divided Channel Method (DCM) ................................... 93

Figure 6-1. Conceptual model of large scale controls on net solute removal in floodplains…..102

Page 10: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

1

CHAPTER 1. INTRODUCTION

“I recognize the right and duty of this generation to develop and use the natural

resources of our land; but I do not recognize the right to waste them, or to rob, by

wasteful use, the generations that come after us” – Theodore Roosevelt, 1910.

1.1 Food, Water, and Energy Nexus While the world’s population is expected to stabilize in the next 100 years [Lutz et al., 2001; Lutz

et al., 2014], the next generation of decision makers, engineers, and scientists are going to face

monumental challenges that involve complex interactions between social, economic, and

environmental drivers. The common denominator will likely be the interplay between food

security [Godfray et al., 2010], supply of freshwater [Vörösmarty et al., 2000], and energy use and

production [Lewis and Nocera, 2006]. Even today, these variables are drivers of unrest and change

across the world. This is most evident in the recent effects of global climate change [Walther et

al., 2002], which have been widely attributed to the increase in fossil fuel emissions which began

during the Industrial Revolution in the mid-1700s [Hegerl et al., 2007]. The Arab Spring, a series

of major protests, civil uprisings, and even civil wars in the Middle East and northern Africa, is

largely attributed to food shortages caused by global fluctuations in climate, among other

complicated social and economic factors [Davide et al., 2015]. Even here in the relatively stable

US, changes in climate have had adverse effects in the form of both drought and flood events.

During the 2012 drought in the Midwest and Northern Plains, commodity prices increased

anywhere from 2% (dairy) to 16% (beef) because of drought-impacted grain crop [Mallya et al.,

2013]. Ironically, we are simultaneously observing a general intensification of flooding in the

Midwest [Mallakpour and Villarini, 2015] and Eastern US [Villarini and Smith, 2010], the cost of

which will likely be disproportionate. Just in the last decade, flooding damage from Hurricane

Sandy (2012) and Hurricane Katrina (2005) was estimated at $50 Billion and $108 Billion,

respectively [Berg et al., 2013; Brown et al., 2005]. To put these costs into perspective, the cost of

Hurricane Katrina was approximately 4% of federal spending in 2005 [US Senate, 2005].

Another major challenge is balancing the intensification of agriculture with the loss of ecosystem

services. In the next 30 years alone, current research suggests we will have to double food

production to keep up with current consumption rates associated with increases in both population

and prosperity [Foley et al., 2011; Godfray et al., 2010]. Yet, at the current production levels,

agriculture landuse is already a major driver behind environmental change and degradation [Foley

et al., 2005; Power, 2010]. This degradation is particularly relevant when we examine water

quality across the United States, where over 1 million kilometers of our rivers and streams are

considered impaired [USEPA, 2011]. Much of this degradation is associated with the green

revolution and the subsequent threefold increase in agricultural production over the last 50-75

years, which was made possible by the technological advances in the production of nitrogen (N)

and phosphorus (P) fertilizers [Villalba et al., 2008].

Concentrating large amounts of fertilizers on relatively small areas of land to enhance production

essentially decouples the N and P cycles from carbon cycling [Sharpley et al., 2013], leading to

degradation of hydrologically connected aquatic ecosystems [Vitousek et al., 1997]. In undisturbed

systems, N, P, and carbon cycling is tightly coupled and is controlled by biological processing and

ecosystem stoichiometry. However, in agroecosystems elemental ratios are no longer coupled,

often exceeding assimilative thresholds and leading to the “leakage” of inorganic N and P to

Page 11: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

2

adjacent waterways [Dodds et al., 2010; Evans-White et al., 2009]. Elevated N and P

concentrations promote excessive algal growth and can lead to hypoxia, or suboxic conditions not

suitable for aquatic life [Rabalais et al., 2002], leading to the degradation of both the stream

network and receiving water bodies. Worldwide, over 400 receiving bodies experience hypoxia

annually, including the economically and culturally import Gulf of Mexico and Chesapeake Bay

[Diaz and Rosenberg, 2011].

1.2 Riverine Systems and Floodplains The ecological integrity and the associated ecosystem services of our riverine systems are directly

linked to the dynamic nature of the hydrologic cycle [Poff et al., 1997] and heterogeneity of our

landscapes [Ward et al., 2002]. Unlike a pipe, riverine systems have multiple methods of

conveying water in both the subsurface and surface environments. Subsurface flowpaths, or

hyporheic flow, can vary from short flowpaths on the order of centimeters to longer flowpaths on

the order of kilometers [Helton et al., 2012]. Hyporheic flow often occurs directly beneath the

stream bed, but can also be found in the subsurface adjacent to the stream [Poole et al., 2008]. As

stream order increases, the complexity and number of hyporheic flowpaths also increases [Stanford

and Ward, 1993], but the flowpaths convey a smaller portion of the total river flow [Wondzell,

2011]. During high flow events (e.g. floods), overbank flow is routed from the channel to adjacent

low-lying land, known as the floodplain [Poole, 2010]. As the stream network shifts from

headwater streams to a larger river, overbank flows transition from short, sporadic flows to longer,

sustained pulses [Tockner et al., 2000]. As river stage dynamically fluctuates, the stream

ecosystem expands and contracts as both hyporheic and overbank flowpaths dynamically connect

and disconnect from the main channel [Stanley et al., 1997]. Many of these alternate flowpaths are

mixing zones between different sources of water [Mertes, 1997], producing ecotonal gradients that

are often both ecologically and economically important [McClain et al., 2003; Vidon et al., 2010].

Of particular interest, floodplains are hotspots for solute retention and removal within riverine

networks. This is largely due to feedbacks between inundation hydrology and biogeochemical

processes that control solute retention and removal within floodplains [Junk et al., 1989; Tockner

et al., 2000]. More specifically, increased residence time of floodwaters increase the potential for

solute retention and removal [Forshay and Stanley, 2005; James et al., 2008]. During inundation,

floodwaters mix with local waters and come into contact with reactive floodplain sediments, often

forming redox gradients at the sediment-water interface which enhance redox-dependent processes

such as denitrification [Mertes, 1997; Noe and Hupp, 2005]. However, both short and prolonged

residence times can restrict redox-dependent processes, suggesting that a balance between river-

floodplain exchange and residence time distributions should be established [James et al., 2008;

Tockner et al., 1999]. However, complicated feedbacks between river-floodplain exchange and

inundation hydrology limit our understanding of how to achieve this balance at both the reach and

basin scale.

1.3 Ecological Engineering and River-Floodplain Connectivity Riverine floodplains provide many ecosystem services, ranging from increased

agriculture/silviculture production and flood peak attenuation, to biogeochemical processing of

excess agrochemicals during flood events [Tockner and Stanford, 2002]. In many human-impacted

basins, the majority of the solute balance occurs during discrete flood pulses when rivers are

hydrologically connected to their adjacent floodplains [Kaushal et al., 2014; Royer et al., 2006],

making river-floodplain connectivity an attractive management variable when engineering

solutions to reduce the downstream flux of reactive solutes (e.g. Mitsch et al. [2001]). The

Page 12: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

3

importance of river-floodplain connectivity is evident in recent efforts in the Mid-Atlantic and

North Eastern US associated with the Chesapeake Bay TMDL, where grants for stream restoration

are prioritized through functional assessments which emphasize river-floodplain connectivity

[Harman et al., 2012]. However, there is little scientific literature to guide ecological engineering

efforts which optimize river-floodplain connectivity for solute retention. Therefore, the overall

goal of this dissertation is to further investigate feedbacks between inundation hydrology

and floodplain biogeochemistry, specifically analyzing variation experienced along the

gradient from headwater streams to large rivers and the cumulative effects of river-

floodplain connectivity at the basin scale.

This dissertation is composed of four independent studies that spanned the river continuum,

investigating the effects of residence time, dissolved organic matter content, seasonality, and

hydrogeomorphic conditions on the downstream flux of solutes:

Chapter 2. Perirheic mixing and biogeochemical processing in flow-through and

backwater floodplain wetlands

Chapter 3. Biogeochemical processing of dissolved organic matter and nitrogen in

floodwaters across a gradient of river-floodplain connectivity within the Atchafalaya River

floodplain

Chapter 4. Seasonal variation in floodplain biogeochemical processing in a restored

headwater stream

Chapter 5. The effect of hydrogeomorphic gradients and hydroclimatic setting on river-

floodplain connectivity at the continental scale

“In the space of one hundred and seventy-six years the Lower Mississippi has shortened itself two

hundred and forty-two miles. That is an average of a trifle over one mile and a third per year.

Therefore, any calm person, who is not blind or idiotic, can see that in the Old Oolitic Silurian

Period, just a million years ago next November, the Lower Mississippi River was upwards of one

million three hundred thousand miles long, and stuck out over the Gulf of Mexico like a fishing-

rod. And by the same token any person can see that seven hundred and forty-two years from now

the Lower Mississippi will be only a mile and three-quarters long, and Cairo and New Orleans

will have joined their streets together, and be plodding comfortably along under a single mayor

and a mutual board of aldermen. There is something fascinating about science. One gets such

wholesale returns of conjecture out of such a trifling investment of fact.”

- Life on the Mississippi, Mark Twain

Page 13: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

4

1.4 Works Cited Berg, R. J., J. L. Beven Ii, E. S. Blake, J. P. Cangialosi, T. B. Kimberlain, and C. National

Hurricane (2013), United States National Oceanic and Atmospheric Administration's National

Weather Service.

Brown, D. P., R. D. Knabb, C. National Hurricane, and J. R. Rhome (2005), United States

National Oceanic and Atmospheric Administration's National Weather Service.

Davide, N., J. Aled Wynne, and B. Giangiacomo (2015), Quantitative Assessment of Political

Fragility Indices and Food Prices as Indicators of Food Riots in Countries, Sustainability, 7(4).

Diaz, R. J., and R. Rosenberg (2011), Introduction to Environmental and Economic

Consequences of Hypoxia, Int. J. Water Resour. Dev., 27(1), 71-82.

Dodds, W. K., W. H. Clements, K. Gido, R. H. Hilderbrand, and R. S. King (2010), Thresholds,

breakpoints, and nonlinearity in freshwaters as related to management, Journal of the North

American Benthological Society, 29(3), 988-997.

Evans-White, M. A., W. K. Dodds, D. G. Huggins, and D. S. Baker (2009), Thresholds in

macroinvertebrate biodiversity and stoichiometry across water-quality gradients in Central Plains

(USA) streams, Journal of the North American Benthological Society, 28(4), 855-868.

Foley, J. A., et al. (2005), Global Consequences of Land Use, Science, 309(5734), 570-574.

Foley, J. A., et al. (2011), Solutions for a cultivated planet, Nature, 478(7369), 337-342.

Forshay, K. J., and E. H. Stanley (2005), Rapid nitrate loss and denitrification in a temperate

river floodplain, Biogeochemistry, 75(1), 43-64.

Godfray, C. J., J. R. Beddington, I. R. Crute, L. Haddad, L. Davi, J. F. Muir, P. Jule, S.

Robinson, S. M. Thomas, and C. Toulmin (2010), Food Security: The Challenge of Feeding 9

Billion People, Science, 327(5967), 812-818.

Harman, W., R. Starr, M. Carter, K. Tweedy, M. Clemmons, K. Suggs, and C. Miller (2012), A

Function-Based Framework for Stream Assessment and Restoration ProjectsRep. EPA 843-K-12-

006, US Environmental Protection Agency, Office of Wetlands, Oceans, and Watersheds,

Washington, DC.

Hegerl, G., F. Zwiers, P. Braconnot, N. Gillet, and Y. Luo (2007), Understanding and attributing

climate change, Understanding and attributing climate change.

Helton, A. M., G. C. Poole, R. A. Payn, C. Izurieta, and J. A. Stanford (2012), Scaling flow path

processes to fluvial landscapes: An integrated field and model assessment of temperature and

dissolved oxygen dynamics in a river-floodplain-aquifer system, J. Geophys. Res.-Biogeosci.,

117, 13.

Page 14: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

5

James, W. F., W. B. Richardson, and D. M. Soballe (2008), Effects of residence time on summer

nitrate uptake in mississippi river flow‐ regulated backwaters, River Research and Applications,

24(9), 1206-1217.

Junk, W. J., P. B. Balyley, and R. E. Sparks (1989), The flood pulse concept in river-floodplain

systems., Can. Spec. Publ. Fish. Aquat. Sci., 106, 110-127.

Kaushal, S. S., P. M. Mayer, and P. G. Vidon (2014), Land use and climate variability amplify

carbon, nutrient, and contaminant pulses: a review with management implications, JAWRA

Journal of ….

Lewis, N. S., and D. G. Nocera (2006), Powering the planet: Chemical challenges in solar energy

utilization, Proceedings of the National Academy of Sciences, 103(43), 15729-15735.

Lutz, W., W. Sanderson, and S. Scherbov (2001), The end of world population growth, Nature,

412(6846), 543-545.

Lutz, W., W. Butz, S. KC, W. Sanderson, and S. Scherbov (2014), Population growth: Peak

probability, Science, 346(6209), 561.

Mallakpour, I., and G. Villarini (2015), The changing nature of flooding across the central

United States, Nature Clim. Change, 5(3), 250-254.

Mallya, G., L. Zhao, X. C. Song, D. Niyogi, and R. S. Govindaraju (2013), 2012 Midwest

Drought in the United States, J Hydrol Eng, 18(7), 737745.

McClain, M. E., et al. (2003), Biogeochemical Hot Spots and Hot Moments at the Interface of

Terrestrial and Aquatic Ecosystems, Ecosystems, 6(4), 301-312.

Mertes, L. A. K. (1997), Documentation and significance of the perirheic zone on inundated

floodplains, Water Resour. Res., 33(7), 1749-1762.

Mitsch, W. J., J. W. Day, J. W. Gilliam, P. M. Groffman, D. L. Hey, G. W. Randall, and N. M.

Wang (2001), Reducing nitrogen loading to the Gulf of Mexico from the Mississippi River

Basin: Strategies to counter a persistent ecological problem, Bioscience, 51(5), 373-388.

Noe, G. B., and C. R. Hupp (2005), Carbon, nitrogen, and phosphorus accumulation in

floodplains of Atlantic Coastal Plain rivers, USA, Ecological Applications, 15(4), 1178-1190.

Poff, N. L., J. D. Allan, M. B. Bain, J. R. Karr, K. L. Prestegaard, B. D. Richter, R. E. Sparks,

and J. C. Stromberg (1997), The natural flow regime, Bioscience, 47(11), 769-784.

Poole, G. C. (2010), Stream hydrogeomorphology as a physical science basis for advances in

stream ecology, Journal of the North American Benthological Society, 29(1), 12-25.

Page 15: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

6

Poole, G. C., S. J. O'Daniel, K. L. Jones, W. W. Woessner, E. S. Bernhardt, A. M. Helton, J. A.

Stanford, B. R. Boer, and T. J. Beechie (2008), Hydrologic spiralling: The role of multiple

interactive flow paths in stream ecosystems, River Research and Applications, 24(7), 1018-1031.

Power, A. G. (2010), Ecosystem services and agriculture: tradeoffs and synergies, Philosophical

Transactions of the Royal Society of London B: Biological Sciences, 365(1554), 2959-2971.

Rabalais, N. N., R. E. Turner, and W. J. Wiseman (2002), Gulf of Mexico hypoxia, aka "The

dead zone", Annual Review of Ecology and Systematics, 33.

Royer, T. V., M. B. David, and L. E. Gentry (2006), Timing of Riverine Export of Nitrate and

Phosphorus from Agricultural Watersheds in Illinois: Implications for Reducing Nutrient

Loading to the Mississippi River, Environ. Sci. Technol.

Senate, U. (2005), 2005 Public Budget Database, Fiscal Year 2005 Public Budget Database,

Receipts-Public Budget Database.

Sharpley, A., H. P. Jarvie, A. Buda, L. May, B. Spears, and P. Kleinman (2013), Phosphorus

Legacy: Overcoming the Effects of Past Management Practices to Mitigate Future Water Quality

Impairment, Journal of Environmental Quality, 42(5), 1308-1326.

Stanford, J. A., and J. V. Ward (1993), An Ecosystem Perspective of Alluvial Rivers:

Connectivity and the Hyporheic Corridor, Journal of the North American Benthological Society,

12(1), 48-60.

Stanley, E. H., S. G. Fisher, and N. B. Grimm (1997), Ecosystem expansion and contraction in

streams, Bioscience, 47(7), 427-435.

Tockner, K., and J. A. Stanford (2002), Riverine flood plains: present state and future trends,

Environmental Conservation, 29(3), 308-330.

Tockner, K., F. Malard, and J. V. Ward (2000), An extension of the flood pulse concept,

Hydrological Processes, 14(16-17), 2861-2883.

Tockner, K., D. Pennetzdorfer, N. Reiner, F. Schiemer, and J. V. Ward (1999), Hydrological

connectivity, and the exchange of organic matter and nutrients in a dynamic river-floodplain

system (Danube, Austria), Freshw. Biol., 41(3), 521-535.

USEPA (2011), Watershed Assessment, Tracking & Environmnetal Results, edited.

Vidon, P., C. Allan, D. Burns, T. P. Duval, N. Gurwick, S. Inamdar, R. Lowrance, J. Okay, D.

Scott, and S. Sebestyen (2010), Hot Spots and Hot Moments in Riparian Zones: Potential for

Improved Water Quality Management1, J. Am. Water Resour. Assoc., 46(2), 278-298.

Page 16: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

7

Villalba, G., Y. Liu, H. Schroder, and R. U. Ayres (2008), Global Phosphorus Flows in the

Industrial Economy From a Production Perspective, Journal of Industrial Ecology, 12(4), 557-

569.

Villarini, G., and J. A. Smith (2010), Flood peak distributions for the eastern United States,

Water Resources Research, 46(6), W06504.

Vitousek, P. M., H. A. Mooney, J. Lubchenco, and J. M. Melillo (1997), Human Domination of

Earth's Ecosystems, Science, 277(5325), 494-499.

Vörösmarty, C. J., P. Green, J. Salisbury, and R. B. Lammers (2000), Global Water Resources:

Vulnerability from Climate Change and Population Growth, Science, 289(5477), 284-288.

Walther, G.-R., E. Post, P. Convey, A. Menzel, C. Parmesan, T. J. C. Beebee, J.-M. Fromentin,

O. Hoegh-Guldberg, and F. Bairlein (2002), Ecological responses to recent climate change,

Nature, 416(6879), 389-395.

Ward, J. V., K. Tockner, D. B. Arscott, and C. Claret (2002), Riverine landscape diversity,

Freshw. Biol., 47(4), 517-539.

Wondzell, S. M. (2011), The role of the hyporheic zone across stream networks, Hydrological

Processes, 25(22), 3525-3532.

Page 17: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

8

CHAPTER 2. PERIRHEIC MIXING AND BIOGEOCHEMICAL

PROCESSING IN FLOW-THROUGH AND BACKWATER

FLOODPLAIN WETLANDS

C. Nathan Jones, Durelle T. Scott, Brandon L. Edwards, Richard. F. Keim

Submitted: March 2014

To: Water Resources Research

Status: Accepted for publication in September 2014

Abstract Inundation hydrology and associated processes control biogeochemical processing in floodplains.

To better understand how hydrologic connectivity, residence time, and intrafloodplain mixing vary

in floodplain wetlands, we examined how water quality of two contrasting areas in the floodplain

of the Atchafalaya River—a flow-through and a backwater wetland—responded to an annual flood

pulse. Large, synoptic sampling campaigns occurred in both wetlands during the rising limb, peak,

and falling limb of the hydrograph. Using a combination of conservative and reactive tracers, we

inferred three dominant processes that occurred over the course of the flood pulse: flushing (rising

limb), advective transport (peak), and organic matter accumulation (falling limb).

Biogeochemistry of the two wetlands was similar during the peak while the river overflowed into

both. However, during the rising and falling limbs, flow in the backwater wetland experienced

much greater residence time. This led to the accumulation of dissolved organic matter and

dissolved phosphorus. There were also elevated ratios of dissolved organic carbon to nitrate in the

backwater wetland, suggesting nitrogen removal was limited by nitrate transported into the

floodplain there. Collectively, our results suggest inclusion of a temporal component into the

perirheic concept more fully describes inundation hydrology and biogeochemistry in large river

floodplain.

2.1 Introduction Hydrologic connectivity of flowing waters and their adjacent floodplains is a key variable

controlling material flux from river basins. During floods, the river and adjacent floodplain become

hydrologically connected [Poole, 2010], allowing for surface exchange of materials between the

floodplain and river network [Junk et al., 1989; Tockner et al., 2000]. Both the physical

environment (e.g., topography and surface roughness) and antecedent conditions (e.g., soil

moisture) affect water flow and therefore control the movement of reactive solutes across the

floodplain surface [Mertes, 1997; Harvey et al., 2003; Welti et al., 2012]. In many floodplains, the

initial wetting of soils releases solutes attached to the soil surface leading to a first flush of reactive

solutes upon inundation [e.g., Bechtold et al., 2003]. These constituents fuel internal heterotrophic

activity and can also be flushed downstream through the river network [Noe et al., 2013]. Where

floodwaters meet locally derived water on the floodplain, an area of mixing known as the perirheic

zone forms, in part characterized by differences in chemical properties of the water [Mertes, 1997].

In many cases, steep redox gradients develop at the soil-water interface and across the perirheic

zone, allowing for rapid nutrient transformation [Forshay and Stanley, 2005; Racchetti et al., 2011;

Scaroni et al., 2011]. For example, during the 2011 flood within the Atchafalaya River Basin,

Page 18: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

9

Louisiana USA, periods of high hydrologic connectivity during the flood peak resulted in

significant nitrate (NO3-) removal [Scott et al., 2014].

In addition to hydrologic connectivity, residence time of water in the floodplain is a key variable

that affects biogeochemical processes [Cirmo and McDonnell, 1997]. A combination of regional

climate, local hydrology, topography, and feedbacks associated with internal complexity and river-

floodplain exchange, control residence time [Hughes, 1980; Welti et al., 2012]. Areas of the

floodplain that experience long residence times (e.g., backwaters) also experience the greatest

nitrogen removal rates [Misiti et al., 2011]. However, in many systems nitrogen removal can be

limited by NO3- transport [Forshay and Stanley, 2005; Racchetti et al., 2011]. Long-residence

waters begin to accumulate reactive solutes such as soluble reactive phosphorus (SRP),

ammonium, and labile dissolved organic carbon (DOC), posing a risk for downstream waters

[Banach et al., 2009; Chow et al., 2013]. In contrast, flow-through wetlands often have removal

rates as low as ~10% [McJannet et al., 2012]. Therefore, net floodplain function is a balance

between maximizing nutrient retention and minimizing export of other reactive constituents

[Tockner et al., 1999].

Inundation hydrology and associated processes (e.g., hydrologic connectivity, residence time, and

perirheic mixing) are critical for biogeochemical processing in floodplain ecosystems. Within the

Mississippi River Basin, and other large river systems, it has been proposed to restore many of

these processes to reduce the downstream flux of agrochemicals (e.g. Mitsch et al. [2001]).

However, there are few studies that have examined how these processes and the resulting

biogeochemical responses dynamically vary throughout the course of a flood and across the

floodplain. Furthermore, little information is available on how redox conditions in large floodplain

systems affect nitrogen removal and behavior of other reactive solutes. To better understand the

above processes, we conducted synoptic samplings across two contrasting areas of floodplain—a

flow through wetland and a backwater wetland—over the course of a large annual flood pulse.

Specifically, our goal was to examine differences in the evolution of the floodwaters in sites with

contrasting hydrologic regimes and infer the effect of residence time and hydrologic connectivity

on biogeochemical processes.

2.2 Methods 2.2.1 Site Description

The fifth largest river in North America, the Atchafalaya River flows 275 km though the

Atchafalaya River Basin (ARB)—a 3,391 km2 floodplain contained within levees—to its outlet in

the Gulf of Mexico [Ford and Nyman, 2011]. As a distributary system, 30% of the combined flows

of the Mississippi and Red rivers are routed into the Atchafalaya River.

The southern portion of the ARB, where we conducted this study, is made up of a series of

interconnected channels, canals, lakes, and sloughs that generally flow southward towards the Gulf

of Mexico (GOM) through a floodplain dominated by bottomland hardwoods and cypress-tupelo

swamp (Figure 2-1). Because the ARB is confined by artificial levees on the east and west, the

major water source is the Atchafalaya River, but there are several minor sources including

navigational locks and pumping stations for drainage from adjacent agricultural land. Much of the

permanently inundated portion of the basin is hydrologically isolated from the river [Scott et al.,

2014], and many of these backwater locations are low in dissolved oxygen and provide poor fish

Page 19: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

10

habitat [Kaller et al., 2011]. During large flood pulses, as much as 50% of river flow is routed into

the floodplain, reconnecting backwater sites to the main river [Scott et al., 2014]. However, it is

unclear how much floodplain-river exchange occurs after the initial filling of the backwater areas

[BryantMason et al., 2013].

Figure 2-1. Map of the Southern Atchafalaya River Basin (ARB) and the hydrograph at Butte La

Rose (USGS Gage 07381515).

Page 20: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

11

The ARB is divided into 13 distinct Water Management Units (WMUs), separated by internal

barriers to flow, as defined by the U.S. Army Corps of Engineers. For this study, the Buffalo Cove

and Henderson WMUs were chosen to represent flow-through and backwater wetland systems,

respectively. Flow exchanges with the main channel are highly variable at both sites over time.

The Buffalo Cove WMU, approximately 23,000 ha, was previously a lake-margin, deepwater

cypress swamp hydrologically connected to Grand Lake before becoming more isolated in recent

decades by sedimentation from the adjacent Atchafalaya River [Tye and Coleman, 1989].

Typically, the Buffalo Cove WMU only receives water through its southern border. However,

during the large 2011 flood when we sampled, river stage was high enough to route water through

cuts in the levees along the northern boundary and allow flow-through hydrologic conditions, or

flow from upstream to downstream. The Henderson WMU, approximately 29,000 ha, is also a

deepwater cypress swamp, but water is kept artificially deep by a low dam on its sole connection

to the Atchafalaya River near its southeastern (downstream) border [Tye and Coleman, 1989].

Representing backwater hydrology, floodwaters both fill and drain the wetland through this

connection, resulting in relatively hydrologically isolated areas of inundation in the northern

portion of the WMU.

2.2.2 Synoptic Sampling Design and Analysis

Three separate, week-long synoptic sampling campaigns were completed: the rising limb (52

samples), peak (39 samples), and falling limb (31 samples) of the 2011 flood. At each site, a 1 L

sample of near-surface water was collected from a boat using a peristaltic pump and filtered

through 0.45 μM Geotech capsule filters. Samples were stored on ice in the dark and were

transported to the laboratory within 24 hours. Subsamples for nutrient and metals analyses were

stored in 250 mL plastic bottles and frozen until analysis. Subsamples for dissolved organic carbon

and fluorescence analysis were stored in precombusted amber glass vials and refrigerated until

analysis within 5 days. Water isotope samples were collected separately and stored in 20 mL glass

vials with no headspace.

Analyses for ammonium (NH4+) and soluble reactive phosphorus (SRP) were completed on a

SEAL Analytical AA3 analyzer using the Berthelot reaction (salicylate alternative) and the

Murphy and Riley method, respectively. Chloride (Cl-), sulfate (SO42-), and nitrate (NO3-) were

measured through ion chromatography using a Dionex ICS-3000. Inductively couple plasma

atomic emission spectrometry was used to measure dissolved metals (Mn2+ and Fe2+). Dissolved

organic carbon (DOC) and total dissolved nitrogen (TDN) were measured using high temperature

oxidative combustion using Shimadzu TOC-V CPH. Dissolved organic nitrogen (DON) was

calculated by subtracting NO3- and NH4+ from TDN. Carbon fluorescence was measured using

Horiba Jobin Yvon Fluoromax-4 following the protocol outlined by Cory et al. [2010]. Excitation-

emission matrices were applied to the Cory and McKnight [2005] Parallel Factor Analysis

(PARAFAC) model to analyze DOM composition. Specifically, the redox coefficient was

estimated as RC = Q2/HQ, where Q2 and HQ are oxidized and reduced quinones of the DOM,

respectively, and large RC indicates oxidized conditions [Cory and McKnight, 2005]. Stable

isotopic composition, δ18O and δ2H (δD), was measured using cavity ring-down laser

spectrometry with a Picarro Water Isotope Analyzer. Deuterium excess (d) represents the deviation

from the meteoric water line (MWL), which indicates fractionation caused by evaporation

[Kendall and McDonnell, 1998].

Page 21: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

12

2.2.3 Residence Time Conceptualization

Our analysis investigated both the temporal and spatial variability of hydrologic residence time in

the ARB using a combination of conservative and reactive solutes. Stable isotopes of water have

been used extensively to define sources of water (e.g., Hunt et al. [1998]) and the age or residence

time of water (e.g., McGuire et al. [2005]) in a variety of systems. Because of strong latitudinal

gradients of isotopic composition of precipitation and thus river water across the Mississippi Basin

[Kendall and Coplen, 2001], river water was relatively low in δD and δ18O compared to rainfall-

derived floodplain water in the ARB [Scott et al. 2014]. In addition, δ18O was high in floodplain

water because of evaporative fractionation over time: as water evaporates, the light isotopologues

are preferentially evaporated, which increases δ18O in the residual water [Kendall and McDonnell,

1998]. Because there are two primary sources of water in the ARB, river water and rain water, and

because both fractionation of river water and addition of rainwater both increase δ18O relative to

river water, δ18O is a good proxy for residence time: high δ18O indicates higher residence time

and low δ18O indicates similarity to river water and therefore a shorter residence time in the ARB.

A similar approach was used by Scott et al. [2014] and Kaller et al. [in review], using δD to identify

regions of lower and higher residence times in the ARB for this event.

2.2.4 Statistical Analysis

A combination of ordination and cluster analyses was used to identify the evolution of water

through time, examine gradients from highly connected to hydrologically isolated regions, and

elucidate the effects of hydrologic connectivity and perirheic mixing on biogeochemical

processing. Nonmetric Multidimensional Scaling (NMDS) was chosen for the ordination analysis

for its ability to handle nonnormal and nonlinear datasets [Kruskal, 1964]. Euclidean distance of

the scaled data was used as the dissimilarity measure. The number of iterations in finding the least-

stress ordination was 10,000 and the first axis was rotated to the most dominant gradient. The

number of axes was chosen by minimizing both the dimensionality, the number of axis, and stress,

a goodness-of-fit measure that represents the deformation in the data as it is transformed into

ordination space [Clarke, 1993; Kruskal, 1964]. Linear environmental vectors were fit using least

squares to identify gradients of constituents in ordination space. All analyses were completed in

statistical software R using the Vegan package.

A Hierarchical Cluster Analysis (HCA) was run on the NMDS ordination scores to separate the

ordination into groupings. Ward’s Minimum Variance clustering technique was used in

conjunction with Euclidean distance as the dissimilarity measure [Orloci, 1967; Ward, 1963]. A

Multi-Response Permutation Procedure (MRPP) was then used to determine whether the clusters

were significantly different from each other [Mielke and Berry, 2001]. The MRPP was conducted

on the rank transformed water quality data associated with each cluster instead of the NMDS scores

used to develop the clusters. This was done to fully represent the data structure while also

mimicking the NMDS procedure [McCune et al., 2000].

Once statistically different groups were defined, pairwise Wilcoxon tests were used to test for

differences in the individual chemical analytes between each pair of different groups [Hodges and

Lehmann, 1963]. All Wilcoxon test were used in conjunction with Holm’s correction to control

for family-wise error rates at alpha = 0.05 [Holm, 1979].

Page 22: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

13

Figure 2-2. Constituent concentrations in ordination space. Bubble size is proportional to the

concentration of each constituent, and the length and direction of vector arrows indicate the

direction and strength of linear correlation of that constituent projected.

δ18O R2=0.87

NO3_

R2=0.63

SRP R2=0.48

Fe R2=0.66

RC R2=0.72

Cl-

R2=0.54

DOC R2=0.83

DON

R2=0.69

NMDS-1 NMDS-3

SO42-

R2=0.67

A.)

C.)

NH4+

R2=0.35

D.)

E.)

F.)

G.)

H.)

I.)

J.)

B.)

Page 23: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

14

2.3 Results 2.3.1 Ordination Analysis

A 3-axis NMDS model best represented the gradients in the dataset. There was high correlation

(R2 = 0.956) and relatively low stress (S=0.10) between the observed and modeled dissimilarity

(Figure 2-2). The gradient in NMDS-1 was associated with -DOC, -DON, -δ18O, +NO3, and +SO42-

(Figure 2-2a-d,j). Because DOM accumulates over time (e.g., Tockner et al. [1999]), electron

acceptors (e.g. NO3- and SO4

2-) typically decrease with increased residence time (e.g., Forshay and

Stanley, [2005]), and enriched (+) isotopic composition indicates evaporation or local rainfall (e.g.,

Scott et al., [2014]), NMDS-1 appears to represent (inverse) residence time. The gradient in

NMDS-2 was associated with -SRP, -Fe2+, and +RC (Figure 2-2e-g). SRP is often released when

redox sensitive ferric soils are reduced (e.g., Banach et al. [2009]). Therefore, NMDS-2 appears

to be an (inverse) redox gradient. Finally, the gradient in NMDS-3 was most closely associated

with -NH4+, -Cl-

, and –SO4- (Figure 2-2h-i); however, these linear gradients are weaker than those

of the previous two groups. Variation of these constituents, specifically with chloride, occurred

during the rising limb, which might have been an initial flush in the connected regions.

Figure 2-3. Evolution of the floodwater chemistry over the course of the annual flood pulse. Axis

labels are qualitative interpretations based on strength of correlation with constituents. Circles

(grey = flow-through, black = backwater) represent individual samples, while density ellipses

represent the location of 50% of samples from the backwater and flow-through wetlands,

respectively, from each synoptic sampling event. This figure corresponds to the first column in

Figure 2-2.

Page 24: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

15

There were temporal changes in water quality across the flood event in both wetlands (Figure 2-3).

Both wetlands started the event with water in a relatively high reduced state (low NMDS-2) and

long residence times (low NMDS-1), then moved to shorter residence times and oxidized states

during the peak of the flood (Figure 2-3). Following the peak, residence time increased in each

wetland but redox states remained relatively oxidized. Because the backwater wetland was less

hydrologically connected to the river, it experienced greater extremes in redox conditions and

residence times compared to the flow-through wetland. While full hysteresis is not represented in

this dataset, it appears likely that reducing conditions were returning to those similar to the rising

limb.

2.3.2 Cluster Analysis of Samples

The cluster analysis separated samples from across the two wetlands into two significantly

different, homogeneous categories (Figure 2-4a; p= 0.0001 and a= 0.14). The two-cluster model

accounted for 58% of the variance in the data. Because of differences in isotopic content

(Figure 2-4c), the two clusters were classified as connected and disconnected.

Figure 2-4. Separation of Connected and Disconnected sites. A.) Dendrogram resulting from the

Hierarchal Cluster Analysis (HCA) on NMDS scores. Results separated disconnected, orange,

from connected, blue, sites. B.) Locations of connected and disconnected sites in each of three

synoptic samples. C-G.) Boxplots of floodwater constituents. Grey diamonds indicate mean

concentration in river water [Scott et al. 2014]. Statistical differences between classes are denoted

by alphabetic groups

Page 25: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

16

Over the course of the storm, many sampling sites fluctuated between the connected and

disconnected classifications (Figure 2-4b). The greatest spatial variability occurred during the

rising limb, when sites in backwater and flow-through wetlands were primarily categorized as

disconnected and connected, respectively. During the peak, all were classified as connected, and

during the falling limb, sites once again separated into two distinct classes.

2.3.3 Comparison between the Connected and Disconnected Classes

Dissolved constituents varied significantly by connectivity class (Figure 2-4c-g). Disconnected

sites (enriched in δ18O) had significantly higher concentrations of DOC and Fe2+ and lower

concentrations of TDN. There was no significant difference in SRP by connectivity class.

Concentrations of nitrogen species varied by connectivity class and through the course of the

flooding event (Figure 2-5). During the rising limb of the flood, NO3- was significantly higher in

connected sites (Figure 2-5a), but there was no significant difference between classes during the

falling limb. In contrast, DON and the DOC:NO3- ratio were significantly higher in disconnected

sites over the course of the flood. There were no significant differences in NH4+ by connectivity

class over the course of the flood, though the direction of the effects was similar to that of NO3-.

Figure 2-5. Box plots of nitrogen speciation and the DOC:NO3- ratios throughout the flood,

separated by connected (blue) and disconnected (orange) sites as defined by the HCA (Fig 4).

Differences among groups are indicated by letters.

There was a strong, positive relationship between δ18O and DOC (Figure 2-6a), which suggests

residence time was a controlling factor on DOM accumulation in the two wetlands over the course

of the flood. There was separation between connected and disconnected classes, consistent with

the finding that δ18O and DOC were dominant variables in the NMDS and HCA. There was a

positive relationship between Fe2+ and SRP (Figure 2-6b), which suggests linkage by redox

process, consistent with the NMDS. The separation between classes is not as distinct, suggesting

site-specific factors were important for lower concentrations of Fe2+ and SRP, but the highest

concentrations of Fe2+ and SRP occurred in disconnected sites.

B.)

C.)

D.)

A.)

B.)

C.)

D.)

A.)

B.)

C.)

D.)

A.)

B.)

C.)

D.)

A.)

Page 26: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

17

Figure 2-6. Relationship between A.) DOC and δ18O and B.) Fe and SRP at connected (blue)

and disconnected (orange) sites.

2.4 Discussion 2.4.1 Conceptual Model

Our results suggest that perirheic exchange, or surficial mixing in the floodplain, and redox

conditions controlled the biogeochemistry of floodwaters during the flood (Figure 2-7). Sites were

disconnected, or hydrologically isolated, when river stage was relatively low, resulting in limited

exchange of water and materials between the river and floodplain. Biogeochemical cycling was

fueled by internal, autochthonous materials, and long residence times led to reducing conditions,

which allowed for the accumulation of organic matter and SRP. In connected sites, local floodplain

water mixed with floodwater, forming a transitional zone where allochthonous materials were

transported from oxic floodwaters and fueled denitrification.

The zone of intrafloodplain mixing, at the interface between floodplain water and floodwaters, is

important for biogeochemical processing and can be termed the biogeochemical perirheic zone.

This is an extension of hydrologic concept of the perirheic zone originally proposed by Mertes

[1997], which is the mixing zone between new water (i.e., river water) and old water (e.g.,

groundwater, flow from small tributary streams, and precipitation). Within the ARB, the main

source of water is the river. However, our results show that the residence time in floodplain

flowpaths changes the composition of that river water, essentially creating floodplain water as a

distinct end-member.

Adding a time dimension to the perirheic concept more fully describes inundation hydrology in

large river floodplains, where backwater residence times are long enough to substantially alter

water quality and solute concentrations. For example, mean water residence time within a

relatively pristine Amazon River floodplain was estimated to be 5 months during the seasonal

flood pulse [Bonnet et al., 2008]. Similarly, in a recently restored floodplain along the Danube

River, Schagerl et al. [2009] found mean water residence times up to 30 days and maximum

residence times up to 90 days on an annual basis. In both cases, the characteristics of the floodplain

water changed over time due to depositional processes and biological activity [Hein et al., 2004;

Bourgoin et al., 2007]. Within the ARB, Scott et al. [2014] found significant differences between

hydrologically isolated sites and river water, indicating biogeochemical processing in the

Page 27: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

18

floodplain. The results presented in our study confirm these findings, highlighting the observed

gradient from riverine floodwater to biogeochemically processed floodplain waters.

Figure 2-7. Conceptual sketch illustrating the differences between flow-through (connected) and

backwater (disconnected) locations within a floodplain.

2.4.2 Classification of Hydrologically Connected and Disconnected Locations

The results from the cluster analysis suggest both temporal and spatial variation in hydrologic

conditions (i.e., residence time of water) and biogeochemical processes (Figure 2- 4). We infer

that the connected and disconnected classes represent floodwater and floodplain water,

respectively, as defined by the conceptual model above. The chemistry associated with the

connected cluster most closely resembled river water. This is consistent with the observation that

all sites were classified as connected during the peak of the flood (Figure 4b), with floodwater

biogeochemistry. Conversely, there were both connected and disconnected sites during the rising

and falling limbs. In sites classified as disconnected, the enriched δ18O and elevated DOM suggest

relatively long residence times (Figures 4c-d), while lower TDN and higher Fe2+ suggest reducing

conditions (Figure 2-4e,g). Both longer residence times and reducing redox conditions are

indicative of processed floodplain water [Racchetti et al., 2011]. Although sites were classified in

a binary fashion, they exist on a continuum of residence time. As water moves across the

continuum, the accumulation of DOM and subsequent depletion electron acceptors indicate a shift

Page 28: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

19

from allochthonous to autochthonous control of biogeochemical processes [Junk et al., 1989;

Tockner et al., 2000]. While processes were observed within a temporal pattern of variability, the

degree of hydrologic connectivity was greater than normal during the observed flood pulse.

2.4.3 Carbon, Nitrogen, and Phosphorus dynamics in the ARB

Carbon Accumulation in Backwaters

In the ARB, inundation hydrology and the residence time of water in the floodplain are large

controls on local dissolved organic matter. Increased residence time in backwater areas results in

increased DOC (Figure 2-6b). Similar increases of DOM in floodplain wetlands have been

observed in both the pristine Okavango Delta of Botswana [Cawley et al., 2012] and in a restored

floodplain along the Danube in Austria [Reckendorfer et al., 2013]. Increases in DOM can come

from flushing of organic soils [Fellman et al., 2009], leachate from litter-fall [Chow et al., 2013],

and algal biomass [Hein et al., 2004]. While further analysis and characterization is needed to

understand the sources and processing of organic matter in the ARB, the relationship between

DOC and δ18O was one of the strongest signals on the floodwater biogeochemistry and suggests

residence time is an important variable in floodplain DOM dynamics.

Nitrogen Retention and Removal

In floodplain wetlands, the combination of landscape position and inundation hydrology plays a

crucial role in local redox conditions at the sediment-water interface [Welti et al., 2012], thus

controlling internal cycling of nitrogen [Wolf et al., 2013]. During both the rising and falling limbs,

sites classified as disconnected had significantly lower NO3- and significantly higher DOC:NO3

-

ratios. This suggests that NO3- was limiting in the disconnected sites, consumed by both biotic

assimilation and, more importantly, denitrification [Taylor and Townsend, 2010; Scaroni et al.,

2011]. While denitrification rates were not explicitly measured in our study, other studies have

indicated similar results when comparing connected and disconnected floodplain wetlands

[Forshay and Stanley, 2005; Racchetti et al., 2011]. Therefore, while redox conditions in

backwater areas are possibly more favorable for nitrogen removal, the increased flux of NO3- in

less hydrologically isolated sites most likely leads to greater nitrogen loss through denitrification.

Scott et al. [2014] found that during the peak of the 2011 flood in the ARB, the entire basin remove

~37% of NO3- on a daily basis, with over 60% of NO3

- in backwater sites removed. However,

during low flow when there was limited exchange of materials (e.g., NO3-) between the river and

floodplain, the ARB removed less than 10% of NO3- on a daily basis.

Phosphorus Release

In sites classified as disconnected, it appears that SRP was mobilized through redox reactions with

ferric soils (Figure 2-6b). There are three main mechanisms for phosphorus release in wetlands

[Ardon et al., 2010]: mobilization of phosphorus as sediment (e.g. Surridge et al. [2012]),

mineralization of organic phosphorus (e.g. Noe and Hupp [2005]), and desorption of Fe2+-bound

ortho-phosphate (e.g. Aldous et al. [2005]). The last two processes have been found to be

responsible for the downstream flux of SRP from floodplain ecosystems [Baldwin and Mitchell,

2000]. Because flow velocity was low and therefore sediment transport was limited in our

disconnected sites, we conclude the most likely reason for increased SRP is desorption, similar to

Banach et al. [2009], who found that local redox conditions control phosphorus export.

Page 29: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

20

2.4.4 Temporal and Spatial Variability in Floodplain Processes

Floodwater biogeochemistry was hysteretic with respect to redox conditions and connectivity over

the course of the flood (Figure 2-3), suggesting there are three dominant processes controlling

floodwater biogeochemistry: flushing (dominant during the rising limb), advective transport

(dominant during the peak), and organic matter accumulation (dominant during the falling limb).

These three processes occurred simultaneously throughout the flood, but the prevalence of each

process varied spatially and temporally depending on local conditions. Tockner et al. [1999, 2000]

suggested that a similar pattern of hysteresis occurs in the Danube floodplain, where the floodplain

transitions from a biotic phase (i.e., disconnected), to primary production phase (i.e., groundwater

connection), and a transport phase (i.e., surface connection). While groundwater flow in the ARB

is restricted [Newman and Keim, 2013], a similar continuum between allochthonous and

autochthonous control appears to occur throughout the flood period.

Because of the timing of the three synoptic events (Figure 2-1), it appears full hysteresis was

expressed from baseflow to baseflow before and after the event, because rising-limb conditions

indicated long residence time in backwater sites in the first sample period. As the ARB transitioned

from the rising limb to the peak of the hydrograph, there was a strong gradient in SRP and Fe2+

(Figure 2-2e-f and Figure 2-3). This suggests that the more hydrologically isolated sites in the

backwater wetland had developed reducing conditions resulting in the mobilizing of SRP during

iron reduction. As these sites became connected, flushing exported SRP. It is likely much of the

flushed SRP then sorbed to suspended sediment in the water column [Owens and Walling, 2002],

potentially increasing the redox potential of the flood waters. Once at the peak, there was much

less variability in water chemistry at all sites, suggesting dominance by river water and low

residence time for biogeochemical processing. During the falling limb, NO3- decreased and

dissolved organic matter increased to concentrations similar to the rising limb. However, iron and

SRP were much lower on the falling limb, suggesting that residence times were not long enough

to favor desorption of iron through redox processes.

2.4.5 Management Implications

In watershed management, it is critical to incorporate strategies and designs that maximize

pollutant removal (e.g., nitrogen through denitrification) while also minimizing the mobilization

of limiting nutrients (e.g., dissolved phosphorus). In the Lower Mississippi River Valley, it has

been proposed to use freshwater diversion and floodplain wetlands to reduce nitrogen loads to the

GOM [Mitsch et al., 2001]. While there is little scientific literature to guide the management of

the large floodplain systems, the Atchafalaya Basin is a highly studies system and can provide

insights that are applicable to other large rivers. This study highlights the need to manage redox

conditions to balance nitrogen removal with reactive solute accumulation. Here, the relatively

disconnected backwater wetland experienced much greater hysteresis in residence time and redox

conditions than the flow-through wetland, indicating a greater accumulation of reactive SRP and

DOM. Because all sites were classified as connected during the flood, it is probable that much of

the accumulated SRP and DOM were flushed downstream, thus increasing the risk of hypoxia to

the GOM. Tockner et al. [1999] found similar results in the restoration of a side arm, or backwater

floodplain wetland, of the Danube River. They suggested that to balance the nutrient retention and

export of DOM it was important to enhance hydrologic connectivity. In the case of the ARB, Scott

et al. [2014] found that greater floodplain-river exchange led to high NO3- removal rates through

perirheic exchange. However, BryantMason et al. [2013] suggested there was limited exchange

Page 30: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

21

between WMUs and the Atchafalaya River after the hydrologically isolated WMUs were filled

during the rising limb. Therefore, by increasing hydrologic connectivity between the river and

adjacent floodplain, possibly by creating flow-through hydrologic conditions, NO3- loads to the

GOM could be further reduced.

2.5 Conclusions We examined how hydrologic regime affected floodplain biogeochemistry in the Atchafalaya

River Basin, a large floodplain. Both the flow-through and backwater wetlands experienced

connected and disconnected, or hydrologically isolated, conditions depending on the timing

relative to the flood peak. Sites were classified in a binary fashion over the course of the flood, but

they actually occurred across a gradient between biogeochemically processed floodplain water and

oxic floodwater. As inundated areas transitioned between hydrologically disconnected and

connected over the course of the flood, three main processes were evident: flushing of

hydrologically isolated areas during the rising limb, advective transport during the peak of the

flood, and organic matter accumulation during the falling limb. During the flushing phase,

disconnected sites showed elevated organic matter and SRP. As backwater regions became

hydrologically connected during peak flow, water chemistry throughout the basin was relatively

uniform because advective transport dominated. However, as sites began to disconnect during the

falling limb, floodplain residence times increased and organic matter accumulation began to

dominate hydrologically remote sites. Throughout the flood, disconnected conditions (e.g. areas

dominated by OM accumulation) only occurred in the backwater wetland, suggesting more

biogeochemical processing within the flow-through wetland. These results highlight the

importance of surficial mixing and biogeochemical exchange that occur in the perirheic zone,

especially when managing the export of redox sensitive materials such as nitrogen and phosphorus.

2.6 Acknowledgements This work was supported by NSF grant 1171363 to Virginia Polytechnic Institute and State

University, NSF grant 1141417 to Louisiana State University, and the Institute for Critical

Technology and Applied Science at Virginia Tech.

2.7 Works Cited Aldous, A., P. McCormick, C. Ferguson, S. Graham, and C. Craft (2005), Hydrologic regime

controls soil phosphorus fluxes in restoration and undisturbed wetlands, Restoration Ecology,

13(2).

Ardon, M., S. Montanari, J. L. Morse, M. W. Doyle, and E. S. Bernhardt (2010), Phosphorus

export from a restored wetland ecosystem in response to natural and experimental hydrologic

fluctuations, J. Geophys. Res.-Biogeosci., 115.

Baldwin, D. S., and A. M. Mitchell (2000), The effects of drying and re-flooding on the sediment

and soil nutrient dynamics of lowland river-floodplain systems: A synthesis, Regul. Rivers-Res.

Manage., 16(5), 457-467.

Banach, A. M., K. Banach, E. J. W. Visser, Z. Stepniewska, A. J. M. Smits, J. G. M. Roelofs, and

L. P. M. Lamers (2009), Effects of summer flooding on floodplain biogeochemistry in Poland;

implications for increased flooding frequency, Biogeochemistry, 92(3), 247-262.

Bechtold, J. S., R. T. Edwards, and R. J. Naiman (2003), Biotic versus hydrologic control over

seasonal nitrate leaching in a floodplain forest, Biogeochemistry, 63(1), 53-71.

Page 31: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

22

Bonnet, M. P., et al. (2008), Floodplain hydrology in an Amazon floodplain lake (Lago Grande de

Curuaí), J. Hydrol., 349(1–2), 18-30.

Bourgoin, L. M., M.-P. Bonnet, J.-M. Martinez, P. Kosuth, G. Cochonneau, P. Moreira-Turcq, J.-

L. Guyot, P. Vauchel, N. Filizola, and P. Seyler (2007), Temporal dynamics of water and sediment

exchanges between the Curuaí floodplain and the Amazon River, Brazil, J. Hydrol., 335(1–2),

140-156.

BryantMason, A., Y. J. Xu, and M. A. Altabet (2013), Limited capacity of river corridor wetlands

to remove nitrate: A case study on the Atchafalaya River Basin during the 2011 Mississippi River

Flooding, Water Resources Research, 49(1), 283-190.

Cawley, K. M., P. Wolski, N. Mladenov, and R. Jaffe (2012), Dissolved Organic Matter

Biogeochemistry Along a Transect of the Okavango Delta, Botswana, Wetlands, 32(3), 475-486.

Cirmo, C. P., and J. J. McDonnell (1997), Linking the hydrologic and biogeochemical controls of

nitrogen transport in near-stream zones of temperate-forested catchments: a review. J. Hydrol.,

199(1), 88-120.

Chow, A. T., J. N. Dai, W. H. Conner, D. R. Hitchcock, and J. J. Wang (2013), Dissolved organic

matter and nutrient dynamics of a coastal freshwater forested wetland in Winyah Bay, South

Carolina, Biogeochemistry, 112(1-3), 571-587.

Clarke, K. R. (1993), Nonparametric multivariate analyses of changes in community structure,

Aust. J. Ecol., 18(1), 117-143.

Cory, R. M., and D. M. McKnight (2005), Fluorescence spectroscopy reveals ubiquitous presence

of oxidized and reduced quinones in dissolved organic matter, Environ. Sci. Technol., 39(21),

8142-8149.

Cory, R. M., M. P. Miller, D. M. McKnight, J. J. Guerard, and P. L. Miller (2010), Effect of

instrument-specific response on the analysis of fulvic acid fluorescence spectra, Limnology and

Oceanography-Methods, 8, 67-78.

Fellman, J. B., E. Hood, R. T. Edwards, and D. V. D'Amore (2009), Changes in the concentration,

biodegradability, and fluorescent properties of dissolved organic matter during stormflows in

coastal temperate watersheds, J. Geophys. Res.-Biogeosci., 114.

Ford, M., and J. A. Nyman (2011), Preface: an overview of the Atchafalaya River, Hydrobiologia,

658(1), 1-5.

Forshay, K. J., and E. H. Stanley (2005), Rapid nitrate loss and denitrification in a temperate river

floodplain, Biogeochemistry, 75(1), 43-64.

Harvey, J. W., M. H. Conklin, and R. S. Koelsch (2003), Predicting changes in hydrologic

retention in an evolving semi-arid alluvial stream, Advances in Water Resources, 26(9), 939-950.

Hein, T., C. Baranyi, W. Reckendorfer, and F. Schiemer (2004), The impact of surface water

exchange on the nutrient and particle dynamics in side-arms along the River Danube, Austria, Sci

Total Environ, 328(1-3), 207-218.

Hodges, J. L., Jr., and E. L. Lehmann (1963), Estimates of Location Based on Rank Tests, The

Annals of Mathematical Statistics, 34(2), 598-611.

Holm, S. (1979), A simple sequentially rejective multiple test procedure, Scandinavian Journal of

Statistics, g, 383-386.

Page 32: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

23

Hughes, D. A. (1980), Floodplain inundation - processes and relationships with channel discharge,

Earth Surf. Process. Landf., 5(3), 297-304.

Hunt, R. J., T. D. Bullen, D. P. Krabbenhoft, and C. Kendall (1998), Using stable isotopes of water

and strontium to investigate the hydrology of a natural and a constructed wetland, Ground Water,

36(3), 434-443.

Junk, W. J., P. B. Balyley, and R. E. Sparks (1989), The flood pulse concept in river-floodplain

systems., Can. Spec. Publ. Fish. Aquat. Sci., 106, 110-127.

Kaller, M. D., W. E. Kelso, B. T. Halloran, and D. A. Rutherford (2011), Effects of spatial scale

on assessment of dissolved oxygen dynamics in the Atchafalaya River Basin, Louisiana,

Hydrobiologia, 658(1), 7-15.

Kaller, M. D., R. F. Keim, B. L. Edwards, A. R. Harlan, T. Pasco, W. E. Kelso, and D. A.

Rutherford. Aquatic vegetation mediates the relationship between hydrologic connectivity and

water quality in a managed floodplain, Atchafalaya River, Louisiana, USA. In review,

Ecohydrology

Kendal, C., and J. McDonnel (1998), Isotope Tracers in Catchment Hydrology, Elsevier, The

Neatherlands.

Kendall, C., and T. B. Coplen (2001), Distribution of oxygen-18 and deuterium in river waters

across the United States, Hydrological Processes, 15(7), 1363-1393.

Kruskal, J. B. (1964), Multidimensional scaling: a numerical method, Psychometrika, 29, 115-129.

Lane, R. R., J. W. Day, and B. Thibodeaux (1999), Water quality analysis of a freshwater diversion

at Caernarvon, Louisiana, Estuaries, 22(2A), 327-336.

McCune, B., R. Rosentreter, J. M. Ponzetti, and D. C. Shaw (2000), Epiphyte habitats in an old

conifer forest in Western Washington, USA, Bryologist, 103(3), 417-427.

McGuire, K. J., J. J. McDonnell, M. Weiler, C. Kendall, B. L. McGlynn, J. M. Welker, and J.

Seibert (2005), The role of topography on catchment-scale water residence time, Water Resources

Research, 41(5).

McJannet, D., J. Wallace, R. Keen, A. Hawdon, and J. Kemei (2012), The filtering capacity of a

tropical riverine wetland: II. Sediment and nutrient balances, Hydrological Processes, 26(1), 53-

72.

Mertes, L. A. K. (1997), Documentation and significance of the perirheic zone on inundated

floodplains, Water Resour. Res., 33(7), 1749-1762.

Mielke, P. W., and K. J. Berry (2001), Permutation Methods: A Distance Function Approach. .

Misiti, T. M., M. G. Hajaya, and S. G. Pavlostathis (2011), Nitrate reduction in a simulated free-

water surface wetland system, Water Research, 45(17), 5587-5598.

Mitsch, W. J., J. W. Day, J. W. Gilliam, P. M. Groffman, D. L. Hey, G. W. Randall, and N. M.

Wang (2001), Reducing nitrogen loading to the Gulf of Mexico from the Mississippi River Basin:

Strategies to counter a persistent ecological problem, Bioscience, 51(5), 373-388.

Newman, A. E., and R. F. Keim (2013), Mesoscale connectivity through a natural levee,

Hydrology and Earth System Sciences, 17(2), 691-704.

Noe, G. B., and C. R. Hupp (2005), Carbon, nitrogen, and phosphorus accumulation in floodplains

of Atlantic Coastal Plain rivers, USA, Ecological Applications, 15(4), 1178-1190.

Page 33: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

24

Noe, G. B., C. R. Hupp, and N. B. Rybicki (2013), Hydrogeomorphology Influences Soil Nitrogen

and Phosphorus Mineralization in Floodplain Wetlands, Ecosystems, 16(1), 75-94.

Oksanen J., F. G. Blanchet, R. Kindt, P. Legendre, P.R. Minchin, R. B. O'Hara, G.L. Simpson, P.

Solymos, M.H.H.Stevens, H. Wagner (2013). vegan: Community Ecology. Package. R package

version 2.0-10. http://CRAN.R-project.org/package=vegan

Orloci, L. (1967), An Agglomerative method for classification of plant communities, Journal of

Ecology, 55, 208-212.

Owens, P. N., and D. E. Walling (2002), The phosphorus content of fluvial sediment in rural and

industrialized river basins, Water Research, 36(3), 685-701.

Poole, G. C. (2010), Stream hydrogeomorphology as a physical science basis for advances in

stream ecology, Journal of the North American Benthological Society, 29(1), 12-25.

R Development Core Team (2012). R: A language and environment for statistical computing. R

Foundation for Statistical Computing, Vienna, Austria. ISBN 3-900051-07-0, URL http://www.R-

project.org/

Racchetti, E., M. Bartoli, E. Soana, D. Longhi, R. R. Christian, M. Pinardi, and P. Viaroli (2011),

Influence of hydrological connectivity of riverine wetlands on nitrogen removal via denitrification,

Biogeochemistry, 103(1-3), 335-354.

Reckendorfer, W., A. Funk, C. Gschopf, T. Hein, and F. Schiemer (2013), Aquatic ecosystem

functions of an isolated floodplain and their implications for flood retention and management, J.

Appl. Ecol., 50(1), 119-128.

Reddy, K. R., and W. H. Patrick (1984), Nitrogen transformations and loss in flooded soils and

sediments, Crc Critical Reviews in Environmental Control, 13(4), 273-309.

Schagerl, M., I. Drozdowski, D. G. Angeler, T. Hein, and S. Preiner (2009), Water age - a major

factor controlling phytoplankton community structure in a reconnected dynamic floodplain

(Danube, Regelsbrunn, Austria), Journal of Limnology, 68(2), 274-287.

Scaroni, A., J. Nyman, and C. Lindau (2011), Comparison of denitrification characteristics among

three habitat types of a large river floodplain: Atchafalaya River Basin, Louisiana, Hydrobiologia,

658(1), 17-25.

Scott, D. T., R. F. Keim, C. N. Jones, B. Edwards, D. E. Kroes, J. R. White, and J. A. Nyman

(2014), Floodplain biogeochemical processing of floodwaters during the Mississippi River flood

of 2011, Water Resources Research. In Press.

Surridge, B. W. J., A. L. Heathwaite, and A. J. Baird (2012), Phosphorus mobilisation and

transport within a long-restored floodplain wetland, Ecol Eng, 44, 348-359.

Taylor, P. G., and A. R. Townsend (2010), Stoichiometric control of organic carbon-nitrate

relationships from soils to the sea, Nature, 464(7292), 1178-1181.

Tockner, K., F. Malard, and J. V. Ward (2000), An extension of the flood pulse concept,

Hydrological Processes, 14(16-17), 2861-2883.

Tockner, K., D. Pennetzdorfer, N. Reiner, F. Schiemer, and J. V. Ward (1999), Hydrological

connectivity, and the exchange of organic matter and nutrients in a dynamic river-floodplain

system (Danube, Austria), Freshw. Biol., 41(3), 521-535.

Page 34: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

25

Tye, R. S., & Coleman, J. M. (1989). Evolution of Atchafalaya lacustrine deltas, south-central

Louisiana. Sedimentary Geology, 65(1), 95-112.

Ward, J. H. (1963), Hierarchical grouping to optimise an objective function, Journal of the

American Statistical Association, 58, 236-244.

Welti, N., E. Bondar-Kunze, G. Singer, M. Tritthart, S. Zechmeister-Boltenstern, T. Hein, and G.

Pinay (2012b), Large-scale controls on potential respiration and denitrification in riverine

floodplains, Ecol Eng, 42, 73-84.

Wolf, K. L., G. B. Noe, and C. Ahn (2013), Hydrologic Connectivity to Streams Increases Nitrogen

and Phosphorus Inputs and Cycling in Soils of Created and Natural Floodplain Wetlands, Journal

of Environmental Quality, 42(4), 1245-1255.

Page 35: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

26

CHAPTER 3. BIOGEOCHEMICAL PROCESSING OF

DISSOLVED ORGANIC MATTER AND NITROGEN IN

FLOODWATERS ACROSS A GRADIENT OF RIVER-

FLOODPLAIN CONNECTIVITY WITHIN THE

ATCHAFALAYA RIVER FLOODPLAIN

C. Nathan Jones, Durelle T. Scott, Brandon L. Edwards, Richard. F. Keim

Submitted: anticipated, August 2015

To: Water Resources Research

Status: Draft

Abstract In large river systems, biogeochemical processing in floodplains can reduce the downstream flux

of contaminants. Thus, careful management of river-floodplain connectivity in large river systems

provides a potentially valuable conduit for mitigating downstream effects of harmful solutes. In

this study, we examined dissolved organic matter and nitrogen across spatial and temporal

gradients of river-floodplain connectivity in a highly connected region of the Atchafalaya River

floodplain. We completed three synoptic sampling campaigns across the rising limb, peak, and

falling limb of the 2013 annual flood pulse. Inundation hydrology and organic matter

biogeochemistry were characterized using hydrologic tracers and spectral indices. Samples were

collected in secondary canals at locations representing a gradient of river-floodplain connectivity.

During the rising and falling limb, isotopic ratios were relatively uniform across the sampling

gradient, suggesting the main source of water in the secondary channels was riverine and there

was little interaction with hydrologically isolated backwaters. In contrast, water isotopes

experienced much great variation during the peak, indicating enhanced river-floodplain

connectivity and biogeochemical interaction between riverine and backwater sources. While

gradients in redox sensitive solutes were observed, relatively low DOC to NO3- ratios suggest

denitrification was not NO3- limited. Further, increases in both dissolved inorganic carbon and the

fluorescence index over the course of the flood suggest that subsidies from floodwater and seasonal

increases in temperature enhanced biogeochemical processing within secondary channels.

3.1 Introduction Most of the world’s large rivers have been significantly impacted by anthropogenic modifications

[Vitousek et al., 1997]. Managed hydrology (e.g. Leopold, 1968), flood control (e.g. Sparks et al.

[1998]), and channelization (e.g. Galat et al. [1998]) have largely reduced river-floodplain

connectivity in many large river systems [Tockner and Stanford, 2002]. River-floodplain

connectivity drives water mediated exchange of material and facilitates retention and removal of

reactive solutes such as nitrogen and phosphorus [Pringle, 2003; Tockner and Stanford, 2002].

Thus, the loss or restriction of hydrologic connectivity between large rivers and their floodplains

exacerbates the downstream flux of harmful solutes typically associated with environmental

degradation (e.g. nitrogen and eutrophication; Rabalais et al. [2002]).

Page 36: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

27

One potentially attractive solution to mitigate the flux of harmful solutes is managing hydraulic

control structures to optimize, or in some cases reestablish, river-floodplain connectivity in large

river systems (e.g. Lane et al., [1999]; James et al., [2008]; Welti et al. [2012]). Because the rate

of biogeochemical processing (e.g. denitrification) in floodplains often exceeds the rate of riverine

solute delivery [Forshay and Stanley, 2005], solute retention is often transport limited [Racchetti

et al., 2011; Jones et al., 2014] and can be increased through enhanced river-floodplain

connectivity [James et al., 2008]. Increased river-floodplain connectivity significantly reduces

downstream NO3- loads [Scott et al., 2014].

While recent evidence suggests managing river-floodplain connectivity is a viable avenue to

reduce downstream flux of solutes (e.g. Mitsch et al., [2001]), there is sparse scientific literature

to guide managers and decision makers. Largely, this is due to the complexity and heterogeneity

of ecosystem processes that occur in large river floodplains [Hughes, 1980; Scown and Thoms,

2015]. Within the Atchafalaya River Basin, a highly managed large river floodplain, previous

research has shown that floodwater residence times strongly influence floodplain biogeochemistry

and thus downstream water quality (e.g. Lambou et al. [1983]; Sabo et al. [1999], Shen et al.

[2012]; Jones et al. [2014]; Scott et al. [2014]). Interestingly, these studies, along with studies in

other large river floodplains (e.g. Mladenov et al. [2005]; Fellman et al. [2013]; Hedges et al.

[2000]; Zurbrugget al. [2013]), suggest that floodplains are typically sources of dissolved organic

matter (DOM) during periods of river-floodplain connectivity. This is of particular interest,

because DOM cycling fuels biogeochemical processing in floodplains (e.g. Hein et al. [2003]) and

in turn imparts a strong signal on downstream water quality (e.g. Shen et al. [2012]).

To better understand DOM cycling during periods of river-floodplain connectivity, and the

associated effect on biogeochemical processing of floodwaters, we conducted large synoptic

sampling campaigns across a hydrologically connected region of the Atchafalaya River Basin

(ARB) over an annual flood pulse. Specifically, the goal of our study was to characterize temporal

and spatial variation in DOM quality over the course of the annual flood and infer the effect of

DOM and river-floodplain connectivity on biogeochemical processing of nitrogen.

3.2 Methods 3.2.1 Site Description

This study was conducted in the Atchafalaya River Basin (ARB) in southern Louisiana, USA. The

Atchafalaya River is the fifth largest river in North America. It starts at the convergence of the

Red and Mississippi Rivers, where it receives 30% of the combined flows of both through the US

Corp of Engineers Old River Control Structure. The river is constrained by guide levees over

approximately the upper half of its length and then flows through an interconnected system of

bayous, canals, and floodplain lakes to the Gulf of Mexico (GOM). In general, the upper reach is

incised and largely disconnected from much of the floodplain except during very large events

[Kroes and Kraemer, 2013]. The lower basin is characterized by a network of lateral secondary

channels which connect riverine flow to the floodplain. The ARB is contained by flood protection

levees on the east and west, and has internal hydraulic control structures that subdivide the basin

into 13 water management units. The basin transitions downstream from agriculture and

bottomland hardwood forest to cypress tupelo swamp. Much of the southern portion of the basin

is near permanently inundated yet hydrologically isolated from downstream transport during

normal flow conditions [Jones et al., 2014]. During annual flood pulses, up to 50% of flow routed

Page 37: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

28

into the ARB is contained within backwater regions [Scott et al., 2014; Kroes et al., 2015], but it

unclear how much river-floodplain connectivity and exchange occurs after the initial filling of

backwater areas in each WMU [BryantMason et al., 2013].

Figure 3-1. Sampling locations within the Atchafalaya River Basin in southern Louisiana, USA.

Page 38: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

29

Samples were collected in the Flat Lake WMU, a hydrologically connected region in the east-

central portion of the lower ARB (Figure 3-1). The WMU is bordered by the Atchafalaya River

on the west and the Intercoastal Waterway (ICWW) on the East, both of which flow southward

towards to the GOM and are connected by a series of canals, bayous, and floodplains lakes

(Figure 3-1). Collectively, these secondary channels act as preferential flowpaths of floodwaters

during annual flood events [Kroes and Kraemer, 2013], facilitating enhanced river-floodplain

connectivity between the main channel and hydrologically isolated backwater regions. Generally,

as floodwaters are routed from the main channel into backwater regions during the rising limb,

areas previously hydrologically isolated or disconnected become connected forming areas of

surficial mixing, or perirheic zones (e.g. Jones et al., [2014]; Mertes, [1997]). During the falling

limb, water slowly drains back to main channel allowing for prolonged hydrologic connectivity

between secondary channels and backwater regions, as illustrated by the counterclockwise

hysteresis between main channel and floodplain stage in Figure 3-2b [Hughes, 1980; Rudorff et

al., 2014]. However, because of the heterogeneous nature of the floodplain surface and the

associated inundation hydrology, secondary channels experience a variety of river-floodplain

connectivity conditions [Jones et al., 2014]. Therefore, to characterize the spatial and temporal

variation of hydrologic connectivity within this system, we chose 16 sampling locations to

represent a gradient of river-floodplain connectivity from the main channel to hydrologically

isolated backwater region. Endmembers included the Atchafalaya River (e.g. most connected) and

Murphy Lake (e.g. most disconnected), as defined by previous studies (e.g. Scott et al., [2014];

Hupp et al., [2008]). Similar to Scott et al., [2014], endmember selection is confirmed by the

gradient in isotopic content during the rising limb synoptic sampling campaign (Figure 3-3b).

Figure 3-2. A.) Hydrograph displaying the 2013 flood pulse experienced in the Atchafalaya River

main channel (USGS Gage 07381490). The blue circle, orange square, and red diamond denote

the rising limb, peak, and falling limb synoptic sampling campains, repsectively. B.) Plot

displaying counterclockwise hysterses exerpienced between the main channel flow and floodplain

stage (USGS Gage 073815963). Blue, orange, and red arrows correspond to rising limb, peak, and

falling limb segments of the hydrograph.

Page 39: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

30

3.2.2 Synoptic Sampling Design and Analysis

Sample Collection and Handling. Three synoptic sampling campaigns were conducted over the

course of the 2013 flood pulse: the rising limb (February), the peak (June), and the falling limb

(August, Figure 3-2a). These synoptic sampling campaigns mirror our sampling of the 2011

Atchafalaya Flood (e.g. Jones et al. 2014], but examines a more hydrologically connected portion

of the ARB. At each site, surface water samples were collected at a depth of 0.5 m and filtered

through a 0.45 µM Geotech capsule filter using a peristaltic pump. Samples were stored in 250 mL

plastic bottles for nutrients; 40 mL precombusted amber glass vials for dissolved organic carbon

(DOC), dissolved inorganic carbon (DIC), carbon isotopes, and fluorescence analysis; and 40 mL

glass vials with no head space for water isotope analysis. All samples were stored on ice and

transported to the laboratory within 24 hr. Once at the lab, nutrient samples and carbon samples

were stored at -20⁰C and 5⁰C, respectively, until analysis.

Water Isotope Analysis. We utilized water isotopes as hydrologic tracers to better characterize

mixing patterns over the course of the flood. Isotopes ratios of both of oxygen and hydrogen (δ18O

and δD) were measured using a Picarro Water Isotope Analyzer. Deuterium excess (d), a measure

of evaporative fractionation in water isotopes, was estimated by the deviation from the meteoric

water line [Kendall and McDonnell, 2006]. The isotopic composition at each sampling location is

a mixture of two water sources: locally derived rainwater and regionally derived floodwater [Scott

et al., 2014]. Because the Mississippi and Red Rivers drain a large portion of the United States,

latitudinal gradients in precipitation strongly influence the floodwater signal and are responsible

for depleted values of δ18O and δD [Kendall and Coplen, 2001]. In contrast, locally derived

rainwaters that are stored in backwater regions are highly enriched because of preferential

evaporation of lighter isotopes. Because of the differences between the regional water and locally

derived water, water isotopes are an ideal tracer that can be utilized to examine sources and

residence time of water within the Atchafalaya Basin [Jones et al., 2014; Scott et al., 2014].

Reactive Solute Analysis. We characterized dissolved N, DOC, and DIC to illustrate spatial and

temporal variation in biogeochemical processing over the course of the flood. NO3-concentrations

were measured through ion chromatography using a Dionex ICS-300 analyzer. Dissolved organic

carbon (DOC), dissolved inorganic carbon (DIC), and total dissolved nitrogen (TDN) were

measured using high temperature oxidative combustion using a Shimadzu TOC-CPH analyzer.

Total Kjeldahl nitrogen (TKN) was estimated by subtracting NO3- from TDN and assumed to

represent the sum of ammonia (NH4+) and dissolved organic nitrogen (DON). Here, NO3

-

represents the mobile form of N (Figure 3-3a), TKN represents the reactive or labile proportion of

dissolved N, and the ratio of DOC to NO3- is an indicator of NO3- limitation for denitrification

and other biogeochemical processes (e.g. Taylor and Townsend [2010]).

Carbon Isotope Analysis. To better understand DOM processing, and its effect on

biogeochemical processing of floodwater N, we measured stable isotopes of dissolved carbon

(δ13C-DIC and δ13C-DOC) and a variety of spectral indices over the course of the flood. δ13C-DOC

and δ13C-DIC were measured using an Aurora TOC analyzer and Picarro carbon isotope analyzer

in series, where the Aurora TOC analyzer oxidized the dissolved carbon to CO2 and the picarro

measured isotopic ratios through ring-down laser spectrometry. Both δ13C-DOC and δ13C-DIC are

tracers of dissolved carbon sources. δ13C-DOC is typically utilized to distinguish differences

between DOM parent material, where C3 and C4 plant’s isotopic composition range from -23 to -

Page 40: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

31

35‰ and -10 to -14‰, respectively [Cerling et al, 1993]. Within the Atchafalaya, potential sources

of DOC include flushing of organic soils (e.g., Fellman et al. [2009]), leachate from litterfall (e.g.

Chow et al. [2013]), and decomposition of algal biomass (e.g. Hein et al. [2003]) [Jones et al.,

2014], all of which have overlapping isotopic signatures [France, 1996]. Sources of DIC could

potentially include the byproducts from respiration (e.g. Welti et al. [2012]), carbonate dissolution

(Raymond and Cole [2003]), and atmospheric exchange of CO2 (e.g. Bianchi et al. [2013]).

Because atmospheric CO2- isotopic composition is nearly constant at -8‰ [Craig, 1953],

metabolism of plant derived DOM (low δ13C) depletes δ13C-DIC while photosynthesis typically

increases δ13C-DIC by introducing enriched atmospheric CO2 [Finlay, 2003].

Spectral Indices. We utilized the spectral indices based on carbon fluorescence and absorbance

to better understand differences in DOM composition and quality over the course of the flood.

Carbon fluorescence and absorbance was measured using a Horiba Jobin Yvon Fluoromax-4 and

a spectrometer following methods outlined by Corey et al. [2010]. Specific UV Absorbance at

254nm (SUVA) is a measure of aromaticity, where large values represent highly aromatic

molecules and are typically assumed to be of terrestrial origin [Weishaar et al., 2003]. Across the

US stream network, SUVA values in rivers and streams typically range from 2-5 L mg-1 m-1, where

values become more homogenous with larger watershed area because of increased residence time

and processing [Creed et al., 2015]. Similarly, we examined the fluorescent component of DOM

across the flood pulse using fluorescence spectrometry and the associated fluorescence index (FI)

and redox index (RI). The FI is used to examine the source of fulvic acid component of DOM,

where microbial derived sources of DOM have relatively high values (e.g. 1.74 in the McMurdo

Dry Valleys) and vegetative sources have relatively low values (e.g. 1.24 in the Suwannee River)

[Hood et al., 2003]. The RI describes the redox state of the fluorescent fraction of DOM and was

calculated using excitation-emission matrices and the Cory and McKnight [2005] Parallel Factor

Analysis (PARAFAC) model [Miller and McKnight, 2010]. Specifically, the RI was estimated as

the ratio of oxidized and reduced quinones identified using the Cory and McKnight [2005]

PARAFAC model, where larger RI values suggest oxidized DOM [Jones et al., 2014].

Statistical Analysis. To infer temporal variation in floodplain biogeochemistry over the flood, a

pairwise Wilcoxon test was used to test for differences in individual chemical analytes across the

three synoptic sampling events [Hodges and Lehmann, 1963]. All Wilcoxon tests were used in

conjunction with a Bonferroni correction to control for family wise error rates [Benjamini and

Hochberg, 1995], and significance was evaluated at p=0.05.

3.3. Results 3.3.1 Hydrologic Tracer of Water

The isotopic signal experienced across the flood was similar to the variation characterized in

previous flood events in other portions of the ARB (e.g. Scott et al. [2014]; Jones et al. [2014]; M

Kaller et al. [2015]): the rising and falling limb experienced relatively enriched values of δ18O and

δD, while the peak experienced more depleted values (Figure 3-2). The isotopic composition at

each sampling location is a mixture of two water sources: locally derived rainwater and regionally

derived floodwater [Scott et al., 2014]. Because the Mississippi and Red Rivers drain a large

portion of the United States, latitudinal gradients in precipitation strongly influence the floodwater

signal and are responsible for depleted values of δ18O and δD [Kendall and Coplen, 2001]. In

contrast, locally derived rainwater are highly enriched because of preferential evaporation of

Page 41: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

32

“light” isotopologues [Kendall and McDonnell, 2006]. Because of the differences between the

regional water and locally derived water, water isotopes are an ideal tracer that can be utilized to

examine sources and residence time of water within the Atchafalaya Basin [Jones et al., 2014;

Scott et al., 2014].

Figure 3-3. Isotopic content of samples collected over the course of the flood.Isotopic content of

samples collected over the course of the flood. Blue, orange, and red points represent samples from

the rising limb, peak, and falling limb of the hydrograph, respectively. In boxplots, black and

brown points correspond to backwater swamp and river endmembers, and statistical differences

are denoted by alphanumeric groups.

3.3.2 Reactive Solute Characterization

In this analysis, we characterized observation of NO3-, TKN, and DOC:TKN ratios to demonstrate

biogeochemical processing experienced within the floodplain over the course of the flood. There

were significantly higher concentrations of nitrate during the peak and falling limb of the

hydrograph. Further, during all three synoptic events, higher NO3- concentrations were in river

water than backwater endmembers, which suggests there is a redox gradient associated with the

degree of hydrological connectivity (e.g. Scott et al. [2014]). Similarly, TKN values were

significantly higher during the peak than rising limb, but the falling limb did not show significant

differences between the other two synoptics. Increased connectivity to backwater and the flushing

of accumulated solutes is likely the cause of the increased TKN values and suggests a pulse of

labile solutes (e.g. NH4+ and DON) occurred during the peak (e.g. BryantMason et al. [2013]).

Finally, the DOC to NO3- ratio was significantly less during the peak and falling limb of the

hydrograph, suggesting relative NO3- limitation of denitrification during the rising limb of the

hydrograph (e.g. Jones et al. [2014]).

3.3.3 DOM Characterization

Previous research has shown that residence time within backwater regions of the ARB strongly

influences DOC concentrations (e.g. Lambou et al. [1983]; Shen et al. [2012]; Jones et al. [2014];

Scott et al. [2014]), but little is known about the DOM composition and its effect on associated

biogeochemical processes. Here, we utilized a combination of DOC/DIC ratios (Figures 5a-b),

stable isotope of DOC and DIC (Figures 5c-d), and spectral indices of DOM (Figure 3-6) to further

characterize the fate and transport of DOM within the ARB.

Page 42: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

33

Figure 3-4. A.) NO3

-,B.) TKN, and C.) DOC:TKN ratio over the course of the flood. Black and

brown points represent river and backwater swamp endmembers.

DOC concentrations generally decreased over the course of the flood, where rising limb

concentrations were significantly higher than the peak and falling limb concentrations

(Figure 3-5a). Potential sources of DOC include flushing of organic soils (e.g., Fellman et al.

[2009]), leachate from litterfall (e.g. Chow et al. [2013]), and algal biomass (e.g. Hein et al.

[2003]). Inversely, DIC concentrations increased over the course of the flood, where all three

synoptic samplings were significantly different (Figure 3-5b). Sources of DIC could potentially

include the byproducts from respiration (e.g. Welti et al. [2012]), carbonate dissolution (Raymond

and Cole [2003]), and atmospheric exchange of CO2 (e.g. Bianchi et al. [2013]).

In this analysis, we characterized observation of NO3-, TKN, and DOC:TKN ratios to demonstrate

biogeochemical processing experienced within the floodplain over the course of the flood. There

Page 43: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

34

were significantly higher concentrations of nitrate during the peak and falling limb of the

hydrograph. Further, during all three synoptic events, higher NO3- concentrations were in river

water than backwater endmembers, which suggests there is a redox gradient associated with the

degree of hydrological connectivity (e.g. Scott et al. [2014]). Similarly, TKN values were

significantly higher during the peak than rising limb, but the falling limb did not show significant

differences between the other two synoptics. Increased connectivity to backwater and the flushing

of accumulated solutes is likely the cause of the increased TKN values and suggests a pulse of

labile solutes (e.g. NH4+ and DON) occurred during the peak (e.g. BryantMason et al. [2013]).

Finally, the DOC to NO3- ratio was significantly less during the peak and falling limb of the

hydrograph, suggesting relative NO3- limitation of denitrification during the rising limb of the

hydrograph (e.g. Jones et al. [2014]).

Figure 3-5. Dissolved organic and inorganic carbon concentration and isotopic content across the

flood. Black and brown points represent river and backwater swamp endmembers.

We utilized the stable isotopes of DOC and DIC to further characterize the sources and fate of

DOM across the sampling events. The rising limb had statistically lower δ13C-DOC than the peak

and falling limbs. δ13C-DOC values during the rising limb are similar to previously measured in

the lower Mississippi Basin (-28‰; Wang et al. [2004]), and enrichment of δ13C-DOC from the

rising limb to peak and falling limb suggest a shift in DOM sources, potentially shifting from

aquatic sources to greater influence of DOM derived in the floodplain [Fellman et al., 2013;

Zurbrugg et al., 2013]. Inversely, δ13C-DIC decreased over the course of the flood, where all three

synoptic samplings were statistically different. Further, there appears to have been a (decreasing)

gradient in δ13C-DIC from floodwaters to backwater across all three events. Rising limb δ13C-DIC

values are similar to typical river water draining from weathered soils (e.g. -10 to -12 ‰, Mook

Page 44: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

35

and Tan [1991]). Depletion of δ13C-DIC during the peak and falling limb synoptics, and also along

the gradient from river to backwater sites, suggest an increase in biogeochemical processing across

both spatial and temporal gradients (e.g. Wachniew [2006]).

Figure 3-6. Specific ultra violet absorbance (SUVA), fluorescence index (FI), and redox index

(RI) from across the flood pulse.

Finally, we utilized SUVA, FI, and RI indices to characterize the DOM composition and quality

over the course of the flood. We observed SUVA in the typical range for flowing waters (e.g. 2-5

L mg-1 m-1) during the rising and falling limbs. However, similar to other observation in large river

floodplains (e.g. Mladenov et al. [2005]), SUVA vales were much greater at the peak indicating a

greater dominance of increase in terrestrial derived DOM likely due to the enhanced river-

floodplain connectivity.

Page 45: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

36

Similarly, we examined the fluorescent component of DOM across the flood pulse using

fluorescence spectrometry and FI and RI indices (Figures 6b-c). Our FI observations fall in the

range typically observed in large rivers (e.g., 1.4-1.5; McKnight et al. [2001]), but increased over

the course of the flood suggesting an increase in microbial derived DOM (Figure 3-6b), suggesting

an increase in heterotrophic activity over the course of the flood. While we observed relatively

little temporal change in RI over the course of the flood, there appears to be a consistent separation

between the river and backwater endmembers (Figure 3-6c).

Figure 3-7. Linear relationships between A.) SUVA and δD, B.) NO3

- and DOC concentrations,

and C.) FI and DIC concentration.

3.4 Discussion 3.4.1 River-floodplain connectivity

Within our study area, secondary channels were highly connected to the river over the course of

the flood pulse (Figure 3-8). During the rising and falling limb, we observed minimal variation in

δD (Figure 3-3b), suggesting the river was the dominant source of water at the sampling locations.

However, during the peak, we observed greater variation and significantly enriched δD, suggesting

mixing between regional floodwaters and processed floodplain water. Connectivity across

hydrologically isolated portions of the floodplain, both permanently and seasonally inundated

areas, is further confirmed by significant increase in SUVA and TKN during the peak (Figures 4b

and 6a), suggesting a greater influence of allochthonous DOM likely originating from previously

disconnected locations. Similar patterns have been observed in other large rivers, where the flood

pulse resulted in an export of DOM to fuel downstream biogeochemical processing (e.g. the

Danube in Europe, Reckendorfer et al. [2013]; the pristine Okavango Delta in Batswana, Mladenov

et al. [2005]; and the Amazon River in South America, Junk et al. [1989]). While this pulse of

potentially reactive DOM fuels biogeochemical processing, it can also lead to excess emissions of

CO2 [Bianchi et al., 2013].

3.4.2 Biogeochemical Processing of Nitrogen within the Floodplain

Previous studies have observed NO3- limited denitrification within the ARB (e.g. Jones et al.

[2014]) and across other medium to large river floodplains (e.g. Forshay and Stanley [2005];

Racchetti et al. [2011]). Largely, this is due to a lack of hydrologic exchange and perirheic mixing

Page 46: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

37

between rivers and their adjacent floodplains [James et al., 2008; Jones et al., 2014]. In contrast,

we did not observe NO3- limitation in our study area, likely due to the high degree of hydrologic

connectivity at the sampling locations. There was a significant linear relationship between DOC

and NO3- (R2=0.52, p<0.05, Figure 3-7b), and relatively low DOC: NO3

- elemental ratios

(Figure 3-4c), both suggesting denitrification was not NO3- limited over the course of the flood

pulse [Taylor and Townsend, 2010]. Further, a redox gradient from the river to backwater

endmembers is evident in both the RI and NO3- data suggesting NO3

- removal through

denitrification (Figure 3-4a and 3-6c). Cumulatively, these measures, along evidence from

previous studies at this site (e.g. Scott et al. [2014]), suggest elevated nitrogen removal due to

increased river-floodplain connectivity. However, Scaroni et al. [2010] found relatively

homogenous (potential) denitrification rates across the ARB and indicated organic matter lability

was another key variable that could limit denitrification within the ARB.

Figure 3-8. Conceptual model of swamp channel hydrology and biogeochemistry evolution over

the course of an annual flood pulse. A.) Secondary channel hydrograph components, B.) DOM

Page 47: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

38

flux, and C.) biogeochemical processing. D.) River-floodplain connectivity during the rising limb,

peak, and falling limb of a typical flood pulse.

3.4.3 Temporal variation in Dissolved Organic Matter

Conceptually, we separate DOM into two distinct categories: highly processed allochthonous

DOM and microbial derived autochthonous DOM. The sources of allochthonous DOM include

both main channel and hydrologically disconnected regions of the floodplain and is characterized

by high aromaticity and low lability. In contrast, the autochthonous DOM is derived from internal

cycling within the secondary channel and is largely composed of labile aliphatic compounds.

Because the Atchafalaya River is composed of the Mississippi and Red rivers, riverine DOM

quality and content should be relatively uniform across the sampling period because of increased

residence time and processing associated with transport in large river systems [Creed et al., 2015;

Rosemond et al., 2015]. Duan et al. [2007] noted that this is particularly true in the lower

Mississippi River, where much of the river is hydrologically disconnected from its floodplain, as

opposed to other large systems such as the Amazon (e.g. Hedges et al. [2000]) and the Zambezi

(e.g. Zurbrugg et al. [2013]). Therefore, as river water is routed into the ARB, conceptualized as

a control volume, any changes to DOM composition are likely do to internal cycling (e.g.

production of autochthonous DOM) and river-floodplain connectivity (e.g. exchange of

allochthonous DOM with the floodplain). SUVA and δ13C-DOC signals (Figures 5c and 6a) both

experience significant changes over the course of the flood, indicating changes in a shift of

dominance of river derived DOM to floodplain derived DOM. Alone, the shift in δ13C-DOC only

suggests a shift in DOM source, largely because of overlap of isotopic content in potential

endmembers [France, 1996]. However, the simultaneous and significant increase in SUVA

suggest allochthonous DOM was the dominant source during the peak of the hydrograph.

It also appears heterotrophic activity and production of autochthonous DOM increased over the

course of the flood because of increases in DIC concentrations (Figures 3-5b and 3-7c), increases

in FI (Figures 3-6b and 3-7c), and decrease in δ13C-DIC (Figure 3-5d). Increased biogeochemical

processing can partially be explained by seasonal differences in temperature, which have been

shown to drive spatial and temporal redox gradients within the ARB [Kaller et al., 2011]. It can

also be explained by nutrient subsidies from the flood pulse (e.g., Junk et al. [1989]; Tockner et

al. [2000]), which have been shown to fuel internal cycling within the ARB [Jones et al., 2014]

and other large river floodplains like the Amazon [Hedges et al., 2000]. While the FI index values

fall within the expected range for large river systems (e.g. McKnight et al. [2001]), they increased

over the course of the flood suggesting a greater influence of autochthonous DOM. Similarly, the

increase in DIC and subsequent depletion of δ13C-DIC suggest increases in respiration and

heterotrophic activity.

These results highlight the hysteretic nature of floodwater biogeochemistry in large river

floodplains, where the floodplains fluctuate between three dominant processes: flushing (rising

limb), advective transport (peak), and organic matter accumulation (falling limb) [Jones et al.,

2014]. More specifically, our study highlights the increased potential for nitrogen retention during

the flood recession because of enhanced production of labile DOM to fuel biogeochemical

processes such as denitrification.

Page 48: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

39

3.5. Conclusion Building on previous research (e.g. Jones et al., [2014]; Scott et al., [2014]), we examined the fate

and transport of dissolved organic matter (DOM) and nitrogen (N) within secondary channels of a

highly connected portion of the Atchafalaya River floodplain. Through three synoptic sampling

campaigns across the rising limb, peak, and falling limb of the hydrograph, we were able to capture

spatial and temporal gradients of inundation hydrology and floodplain biogeochemical processing.

Primary observations included:

The secondary channels within the Atchafalaya River floodplain were largely

dominated by river water during the rising and falling limbs of the annual flood

pulse, but local floodplain water was more dominant during the peak when river-

floodplain connectivity was enhanced.

The Atchafalaya River floodplain was a major source of DOM during the flood,

exporting highly processed aromatic DOM downstream.

Seasonal increases in temperature and nutrient subsidies from the flood pulse

facilitated increases in heterotrophic activity during the falling limb of the annual

flood pulse.

Because of the unique setting of the Atchafalaya River Basin (ARB), we had the opportunity of

examining and isolating the effects of a large-river floodplain on biogeochemical processing of

DOM and N. These findings will be important as managers begin to consider reestablishing river-

floodplain connectivity in areas like the lower Mississippi River Valley, where river-floodplain

connectivity has been eliminated through engineering activities.

3.6 Works Cited Benjamini, Y., and Y. Hochberg (1995), Controlling the false discovery rate - a practical and

powerful approach to multiple testing, J. R. Stat. Soc. Ser. B-Methodol., 57(1), 289-300.

Bianchi, T. S., F. Tigreros, S. A. Lewis, M. Shields, H. J. Mills, D. Butman, C. Osburn, P.

Raymond, C. G. Shank, and S. F. DiMarco (2013), Enhanced transfer of terrestrially derived

carbon to the atmosphere in a flooding event, Geophys. Res. Lett., 40(1), 116-122.

BryantMason, A., Y. Jun Xu, and M. A. Altabet (2013), Limited capacity of river corridor wetlands

to remove nitrate: A case study on the Atchafalaya River Basin during the 2011 Mississippi River

Flooding, Water Resources Research 49, 283–190, doi:10.1029/2012WR012185.

Cerling, T. E., Y. Wang, and J. Quade (1993), Expansion of C4 ecosystems as an indicator of

global ecological change in the late Miocene, Nature, 361(6410), 344-345.

Craig, H. (1953), The geochemistry of the stable carbon isotopes, Geochimica et Cosmochimica

Acta, 3(2), 53-92.

Chow, A. T., J. N. Dai, W. H. Conner, D. R. Hitchcock, and J. J. Wang (2013), Dissolved organic

matter and nutrient dynamics of a coastal freshwater forested wetland in Winyah Bay, South

Carolina, Biogeochemistry, 112(1-3), 571-587.

Cory, R. M., and D. M. McKnight (2005), Fluorescence spectroscopy reveals ubiquitous presence

of oxidized and reduced quinones in dissolved organic matter, Environ. Sci. Technol., 39(21),

8142-8149.

Page 49: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

40

Cory, R. M., M. P. Miller, D. M. McKnight, J. J. Guerard, and P. L. Miller (2010), Effect of

instrument-specific response on the analysis of fulvic acid fluorescence spectra, Limnology and

Oceanography-Methods, 8, 67-78.

Cook, R. L., J. E. Birdwell, C. Lattao, and M. Lowry (2009), A Multi-method comparison of

Atchafalaya Basin surface water organic matter samples, Journal of Environmental Quality, 38(2),

702-711.

Creed, I. F., et al. (2015), The river as a chemostat: fresh perspectives on dissolved organic matter

flowing down the river continuum, Can. J. Fish. Aquat. Sci.

Duan, S., T. S. Bianchi, A. M. Shiller, K. Dria, P. G. Hatcher, and K. R. Carman (2007), Variability

in the bulk composition and abundance of dissolved organic matter in the lower Mississippi and

Pearl rivers, Journal of Geophysical Research: Biogeosciences (2005–2012), 112(G2).

Fellman, J. B., E. Hood, R. T. Edwards, and D. V. D'Amore (2009), Changes in the concentration,

biodegradability, and fluorescent properties of dissolved organic matter during stormflows in

coastal temperate watersheds, J. Geophys. Res.-Biogeosci., 114.

Fellman, J. B., N. E. Pettit, J. Kalic, and P. F. Grierson (2013), Influence of stream-floodplain

biogeochemical linkages on aquatic foodweb structure along a gradient of stream size in a tropical

catchment, Freshwater Science, 32(1), 217-229.

Finlay, J. (2003), Controls of streamwater dissolved inorganic carbon dynamics in a forested

watershed, Biogeochemistry, 62(3).

Ford, M., and J. A. Nyman (2011), Preface: an overview of the Atchafalaya River, Hydrobiologia,

658(1), 1-5.

Forshay, K. J., and E. H. Stanley (2005), Rapid nitrate loss and denitrification in a temperate river

floodplain, Biogeochemistry, 75(1), 43-64.

France, R. L. (1996), Scope for use of stable carbon isotopes in discerning the incorporation of

forest detritus into aquatic foodwebs, Hydrobiologia, 325(3), 219-222.

Galat, D. L., L. H. Fredrickson, and D. D. Humburg (1998), Flooding to restore connectivity of

regulated, large river wetlands natural and controlled flooding as complementary processes along

the lower Missouri River, BioScience, 48(9), 721-733.

Harrison, M. D., A. J. Miller, P. M. Groffman, P. M. Mayer, and S. S. Kaushal (2014), Hydrologic

controls on nitrogen and phosphorous dynamics in relict oxbow wetlands adjacent to an urban

restored stream, JAWRA Journal of the American Water Resources Association, 50(6), 1365-

1382.

Hedges, J. I., et al. (2000), Organic matter in Bolivian tributaries of the Amazon River: A

comparison to the lower mainstream, Limnol. Oceanogr., 45(7), 1449-1466.

Hein, T., C. Baranyi, G. J. Herndl, W. Wanek, and F. Schiemer (2003), Allochthonous and

autochthonous particulate organic matter in floodplains of the River Danube: the importance of

hydrological connectivity, Freshw. Biol., 48(2), 220-232.

Hodges, J. L., Jr., and E. L. Lehmann (1963), Estimates of location based on rank tests, The Annals

of Mathematical Statistics, 34(2), 598-611.

Page 50: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

41

Hood, E., D. M. McKnight, and M. W. Williams (2003), Sources and chemical character of

dissolved organic carbon across an alpine/subalpine ecotone, Green Lakes Valley, Colorado Front

Range, United States, Water Resources Research, 39(7).

Hughes, D. A. (1980), Floodplain inundation - processes and relationships with channel discharge,

Earth Surf. Process. Landf., 5(3), 297-304.

Hughes, F. M. R. (1997), Floodplain biogeomorphology, Prog. Phys. Geogr., 21(4), 501-529.

Hupp, C., C. Demas, D. Kroes, R. Day, and T. Doyle (2008), Recent sedimentation patterns within

the central Atchafalaya Basin, Louisiana, Wetlands, 28(1), 125-140.

James, W. F., W. B. Richardson, and D. M. Soballe (2008), Effects of residence time on summer

nitrate uptake in mississippi river flow‐regulated backwaters, River Research and Applications,

24(9), 1206-1217.

Jones, C. N., D. T. Scott, B. L. Edwards, and R. F. Keim (2014), Perirheic mixing and

biogeochemical processing in flow-through and backwater floodplain wetlands, Water Resources

Research, 50(9), 7394-7405.

Junk, W. J., P. B. Balyley, and R. E. Sparks (1989), The flood pulse concept in river-floodplain

systems., Can. Spec. Publ. Fish. Aquat. Sci., 106, 110-127.

Kaller, M., R. Keim, B. Edwards, A. Raynie Harlan, T. Pasco, W. Kelso, and D. Allen Rutherford

(2015), Aquatic vegetation mediates the relationship between hydrologic connectivity and water

quality in a managed floodplain, Hydrobiologia, 1-13.

Kaller, M. D., W. E. Kelso, B. T. Halloran, and D. A. Rutherford (2011), Effects of spatial scale

on assessment of dissolved oxygen dynamics in the Atchafalaya River Basin, Louisiana,

Hydrobiologia, 658(1), 7-15.

Kaushal, S. S., P. M. Mayer, and P. G. Vidon (2014), Land use and climate variability amplify

carbon, nutrient, and contaminant pulses: a review with management implications, JAWRA

Journal of the American Water Resources Association, 50(3), 585-614.

Kendall, C., and T. B. Coplen (2001), Distribution of oxygen-18 and deuterium in river waters

across the United States, Hydrological Processes, 15(7), 1363-1393.

Kendall, C., and J. J. McDonnell (2006), Isotope Tracers in Catchment Hydrology, Elsevier,

Amsterdam, The Netherlands.

Kreiling, R. M., J. P. Schubauer-Berigan, W. B. Richardson, L. A. Bartsch, P. E. Hughes, J. C.

Cavanaugh, and E. A. Strauss (2013), Wetland management reduces sediment and nutrient loading

to the upper Mississippi River, Journal of Environmental Quality, 42(2), 573-583.

Kroes, D. E., and T. F. Kraemer (2013), Human-induced stream channel abandonment/capture and

filling of floodplain channels within the Atchafalaya River Basin, Louisiana, Geomorphology,

201, 148-156.

Kroes, D. E., E. R. Schenk, G. B. Noe, and A. J. Benthem (2015), Sediment and nutrient trapping

as a result of a temporary Mississippi River floodplain restoration: The Morganza Spillway during

the 2011 Mississippi River Flood, Ecol Eng., 82, 91-102.

Lane, R. R., J. W. Day, and B. Thibodeaux (1999), Water quality analysis of a freshwater diversion

at Caernarvon, Louisiana, Estuaries, 22(2A), 327-336.

Page 51: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

42

Leopold, L. B. (1968), Hydrology for urban land planning: A guidebook on the hydrologic effects

of urban land use, US Government Printing Office.

Llewellyn, D. W., G. P. Shaffer, and N. J. Craig (1996), A decision‐support system for prioritizing

restoration sites on the Mississippi River alluvial plain, Conservation Biology 10(5), 1446-1455.

McKnight, D. M., E. W. Boyer, P. K. Westerhoff, P. T. Doran, T. Kulbe, and D. T. Andersen

(2001), Spectrofluorometric characterization of dissolved organic matter for indication of

precursor organic material and aromaticity, Limnol. Oceanogr., 46(1), 38-48.

Miller, M. P., and D. M. McKnight (2010), Comparison of seasonal changes in fluorescent

dissolved organic matter among aquatic lake and stream sites in the Green Lakes Valley, Journal

of Geophysical Research, 115(G1).

Mitsch, W. J., J. W. Day, J. W. Gilliam, P. M. Groffman, D. L. Hey, G. W. Randall, and N. M.

Wang (2001), Reducing nitrogen loading to the Gulf of Mexico from the Mississippi River Basin:

Strategies to counter a persistent ecological problem, Bioscience, 51(5), 373-388.

Mladenov, N., D. M. McKnight, P. Wolski, and L. Ramberg (2005), Effects of annual flooding on

dissolved organic carbon dynamics within a pristine wetland, the Okavango Delta, Botswana,

Wetlands, 25(3), 622-638.

Mook, W. G., and F. C. Tan (1991), Stable carbon isotope ratios in rivers and estuaries, in

Biogeochemistry of Major World Rivers, edited by E. T. Degens, S. Kemp and J. E. Richey, pp.

245-264, Wiley, New York.

Noe, G. B., and C. R. Hupp (2007), Seasonal variation in nutrient retention during inundation of a

short-hydroperiod floodplain, River Research and Applications, 23(10), 1088-1101.

Pringle, C. (2003), What is hydrologic connectivity and why is it ecologically important?,

Hydrological Processes, 17(13), 2685-2689.

Rabalais, N. N., R. E. Turner, and W. J. Wiseman (2002), Gulf of Mexico hypoxia, aka "The dead

zone", Annual Review of Ecology and Systematics, 33.

Racchetti, E., M. Bartoli, E. Soana, D. Longhi, R. R. Christian, M. Pinardi, and P. Viaroli (2011),

Influence of hydrological connectivity of riverine wetlands on nitrogen removal via denitrification,

Biogeochemistry, 103(1-3), 335-354.

Raymond, P. A., and J. J. Cole (2003), Increase in the export of alkalinity from North America's

largest river, Science, 301(5629), 88-91.

Reckendorfer, W., A. Funk, C. Gschopf, T. Hein, and F. Schiemer (2013), Aquatic ecosystem

functions of an isolated floodplain and their implications for flood retention and management, J.

Appl. Ecol., 50(1), 119-128.

Rosemond, A. D., J. P. Benstead, P. M. Bumpers, and V. Gulis (2015), Experimental nutrient

additions accelerate terrestrial carbon loss from stream ecosystems, Science. 347(6226), 1142-

1145.

Royer, T. V., M. B. David, and L. E. Gentry (2006), Timing of riverine export of nitrate and

phosphorus from agricultural watersheds in Illinois: implications for reducing nutrient loading to

the Mississippi River, Environ. Sci. Technol. 40(13). 4126-4131.

Page 52: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

43

Rudorff, C. M., J. M. Melack, and P. D. Bates (2014), Flooding dynamics on the lower Amazon

floodplain: 1. Hydraulic controls on water elevation, inundation extent, and river‐floodplain

discharge, Water Resources Research, 50(1), 619-634.

Sabo, M. J., C. F. Bryan, W. E. Kelso, and D. A. Rutherford (1999), Hydrology and aquatic habitat

characteristics of a riverine swamp: II. Hydrology and the occurrence of chronic hypoxia,

Regulated Rivers: Research & Management, 15(6), 525-544.

Scaroni, A. E., C. W. Lindau, and J. A. Nyman (2010), Spatial variability of sediment

denitrification across the Atchafalaya River basin, Louisiana, USA, Wetlands, 30(5), 949-955.

Scott, D. T., R. F. Keim, B. L. Edwards, C. N. Jones, and D. E. Kroes (2014), Floodplain

biogeochemical processing of floodwaters in the Atchafalaya River Basin during the Mississippi

River flood of 2011, Journal of Geophysical Research: Biogeosciences, 119(4), 2013JG002477.

Scown, M. W., and M. C. Thoms (2015), An index of floodplain surface complexity, Hydrology

and ….

Shen, Y., C. G. Fichot, and R. Benner (2012), Floodplain influence on dissolved organic matter

composition and export from the Mississippi—Atchafalaya River system to the Gulf of Mexico,

Limnol. Oceanogr.

Sparks, R. E., J. C. Nelson, and Y. Yin (1998), Naturalization of the flood regime in regulated

rivers, Bioscience.

Taylor, P. G., and A. R. Townsend (2010), Stoichiometric control of organic carbon-nitrate

relationships from soils to the sea, Nature, 464(7292), 1178-1181.

Tockner, K., and J. A. Stanford (2002), Riverine flood plains: present state and future trends,

Environmental Conservation, 29(3), 308-330.

Tockner, K., F. Malard, and J. V. Ward (2000), An extension of the flood pulse concept,

Hydrological Processes, 14(16-17), 2861-2883.

Tockner, K., D. Pennetzdorfer, N. Reiner, F. Schiemer, and J. V. Ward (1999), Hydrological

connectivity, and the exchange of organic matter and nutrients in a dynamic river-floodplain

system (Danube, Austria), Freshw. Biol., 41(3), 521-535.

Vitousek, P. M., H. A. Mooney, J. Lubchenco, and J. M. Melillo (1997), Human Domination of

Earth's Ecosystems, Science, 277(5325), 494-499.

Wachniew, P. (2006), Isotopic composition of dissolved inorganic carbon in a large polluted river:

The Vistula, Poland, Chemical Geology.

Wang, X. C., R. F. Chen, and G. B. Gardner (2004), Sources and transport of dissolved and

particulate organic carbon in the Mississippi River estuary and adjacent coastal waters of the

northern Gulf of Mexico, Marine Chemistry, 89(1-4), 241-256.

Weishaar, J. L., G. R. Aiken, B. A. Bergamaschi, M. S. Fram, R. Fujii, and K. Mopper (2003),

Evaluation of specific ultraviolet absorbance as an indicator of the chemical composition and

reactivity of dissolved organic carbon, Environ. Sci. Technol., 37(20), 4702-4708.

Welti, N., E. Bondar-Kunze, G. Singer, M. Tritthart, S. Zechmeister-Boltenstern, T. Hein, and G.

Pinay (2012), Large-scale controls on potential respiration and denitrification in riverine

floodplains, Ecol Eng, 42, 73-84.

Page 53: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

44

Zurbrugg, R., S. Suter, M. F. Lehmann, B. Wehrli, and D. B. Senn (2013), Organic carbon and

nitrogen export from a tropical dam-impacted floodplain system, Biogeosciences, 10(1), 23-38.

.

Page 54: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

45

CHAPTER 4. SEASONAL VARIATION IN FLOODPLAIN

BIOGEOCHEMICAL PROCESSING IN A RESTORED

HEADWATER STREAM

C. Nathan Jones, Durelle T. Scott, Christopher Guth, Erich T. Hester, W. Cully Hession

Submitted: May 2015

To: Environmental Science & Technology

Status: In revision

Abstract Stream and river restoration activities have recently begun to emphasize the enhancement of

biogeochemical processing within river networks through the restoration of floodplain wetlands.

It is generally accepted that this practice removes pollutants such as nitrogen and phosphorus

because the increased contact time of nutrient-rich floodwaters with reactive floodplain sediments.

Our study examines this assumption in the floodplain of a recently restored, low-order stream

through five seasonal experiments. During each experiment, a floodplain wetland was artificially

inundated for three hours. Both the net flux of dissolved nutrients and nitrogen uptake rate were

measured during each experiment. The floodplain was typically a source of dissolved phosphorus

and dissolved organic matter, a sink of NO3-, and variable source/sink of ammonia. NO3

- uptake

rates were relatively high when compared to riverine uptake, especially during the spring and

summer experiments. However, when scaled up to the entire 1 km restoration reach with a simple

inundation model, less than 0.5-1.5% of the annual NO3- load would be removed because of the

short duration of river-floodplain connectivity. These results suggest that restoring river-floodplain

connectivity is not necessarily an appropriate best management practice for nutrient removal in

low-order streams with legacy soil nutrients and provide further guidance restoration activities in

the Mid-Atlantic and North Eastern US.

Photo Credit: W. Cully Hession

Page 55: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

46

4.1 Introduction Stream and river restoration is a rapidly growing industry within the United States1. Largely, this

growth is a response to the almost 1 million km of impaired streams across the United States and

regulatory measures associated with the Clean Water Act and Endangered Species Act.2, 3

Traditional restoration objectives include streambank stabilization, riparian/instream habitat

enhancement, and more recently, a functional lift of the degraded stream ecosystem.4, 5 Here,

functional lift refers to enhancing ecosystem function as a whole, not just focusing on a single

aspect of restoration (e.g. bank stabilization), in order to maximize both ecosystem health and

services. Creating functional lift within the stream can include restoring hydrologic connectivity

between the channel and adjacent landscape.6 Specifically, connectivity between rivers and their

adjacent floodplains provide many ecosystem services such as floodwater storage, increased

ecosystem productivity, and increased biogeochemical processing of floodwaters.7 The latter is of

interest in many urban and agricultural watersheds, where water quality impairment is associated

with storm water management and legacy agricultural practices, respectively.8

Because of their transitional and dynamic nature, floodplains are hotspots for biogeochemical

activity9, 10 and can be both a source and sink of nutrients such as nitrogen (N) and phosphorus (P).

In a typical floodplain found in temperate climates, nutrient retention and removal is controlled by

a combination of antecedent moisture condition, biogeochemical processing rates, and the

residence time of floodwaters within the floodplain (Figure 4-1).11, 12 During the interflood period,

or the time between flood events, floodplain sediments can accumulate dissolved organic matter

(DOM),13 soluble reactive phosphorus (SRP),14 and ammonia (NH4+).15 Antecedent soil moisture

conditions control the export and processing of these constituents once the floodplain is

reconnected to the adjacent river channel.16 If the floodplain is relatively dry prior to inundation,

accumulated nutrients can be flushed downstream, where the NH4+ is likely transformed to

NO3- through nitrification17 and SRP likely spirals downstream through cycles of sorption and

desorption with suspended sediments.18 When the floodplain is inundated prior to hydrologic

connection with the river, an area of mixing known as the perirheic zone is established between

the existing floodplain water and river water.19 In many cases, steep redox gradients form across

the perirheic zone and at the sediment-water interface, allowing for rapid transformation of

reactive solutes through redox processes, facilitating the transition of floodwaters to

biogeochemically processed floodplain water.12

In addition to effects associated with antecedent moisture condition, biotic processing and growing

season play an important role in nutrient processing capacity of floodplains.20, 21 Biotic

assimilation, or the conversion of inorganic nutrients to organic nutrients, in benthic sediments is

typically attributed to algal and vegetative communities in floodplains.22 In forested floodplain

systems where water and/or light are limiting, biotic assimilation can account for less than 10% of

added nitrate removal (NO3-).23 In contrast, in highly productive backwater floodplain lakes,

assimilation can account for 76% of NO3- removal.24 In both cases, the floodplain is acting as

temporary sink of N, which will be available for internal cycling during the interflood period and

potentially removed from the floodplain as particulate N through a combination decay processes

(e.g., mineralization) and hydrologic export.25 Seasonal differences in biotic uptake are strongly

coupled with seasonal patterns in water availability, temperature, and solar intensity,21, 26 where

differences between seasonal processing rates can result in N load reductions ranging from 0.05%

to 60%, respectively, within the same floodplain wetland.27

Page 56: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

47

Biogeochemical processing is also strongly tied to the residence times of floodwaters,28, 29 or the

amount of time floodwaters are contained within a floodplain. In medium to large riverine systems,

residence times within floodplains can be long enough to develop conditions ideal for

biogeochemical processing.30 However, in headwater floodplains, short residence times restrict

redox dependent processes such as denitrification,20 and nutrient export is largely controlled by

the balance of biotic uptake and soil flushing processes. This suggests that the variable solute

source/sink characteristic observed headwater floodplains20, 31, 32 is largely dependent on seasonal

variables such as temperature and antecedent soil moisture. The observed variability of

biogeochemical processing across studies is inconsistent with recent efforts in the US Mid-Atlantic

and Northeast to promote river-floodplain connectivity in restoration projects in order to reduce

the downstream flux of nitrogen (N) and phosphorus (P), especially in restoration sites located in

headwater catchments. Therefore, our overall objective was to further investigate mechanisms

associated with N and P processing in a headwater floodplain over the course of an entire year and

to provide the restoration community further guidance to optimize nutrient removal through river-

floodplain connectivity. Specifically, the goals of this study were to examine the first flush of

dissolved nutrients during the initial wetting of soils; quantify the seasonal variation in source/sink

characteristics of the floodplain; and estimate N load reduction associated with river-floodplain

connectivity at the reach scale.

Figure 4-1. Conceptual Model of processes within the experimental floodplain slough. A.) Plan

view of the experimental slough highlighting the inlet and outlet structures, the three sampling

cross sections (XS-1, XS-2, XS-3), and the mixing zone between floodwater (dark grey area)

pumped into the slough and local water (light grey area) likely derived from a combination of

groundwater inputs, rainwater, and previous inundation events. Dashed lines represent

conceptualized flowpaths: advective flow and transient storage dominated flowpaths, respectively.

B.) Cross-sectional view of the inundated floodplain displaying potential mechanisms that control

nutrient cycling and retention within the experimental slough.

Page 57: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

48

4.2 Methods 4.2.1 Site Description

The study site is an abandoned slough within the floodplain of a third-order stream (Stroubles

Creek) in the Ridge and Valley physiographic province of southwestern Virginia, USA and within

the Virginia Tech Stream Restoration, Education, and Management Lab (StREAM Lab). The

contributing watershed is approximately 15 km2 which is 84% urban/residential landcover, 13%

agriculture, and 3% forest.33 The contributing area within the stream reach is predominantly row

crop agriculture, and the site itself is located along a recently restored (c. 2009) reach of stream.

The slough has an approximate surface area of 450 m2 and the flowpath length approximately 60

m (Figure 4-1a). Previous to restoration, the channel was severely incised and riparian productivity

was altered by agricultural grazing activities.34 Post restoration and cattle removal, the site has

been inundated 2-3 times a year and is dominated by reed canary grass.

Figure 4-2. Climatic data measured in the StREAM Lab floodplain. The dark grey area represents

the volumetric soil water content (Vw), the blue line is the annual hydrograph (Q), the light grey

bars represent monthly evapotranspiration (ET), the dark grey points represent mean daily

temperature (Temp), dark grey bars represent the rainfall hyetograph (Precip), and circles and

squares represent the timing of saturated and dry floods, respectively.

Hydrology within the floodplain slough is largely controlled by both evapotranspiration (ET) and

a shallow confining layer of clay. Water sources include periodic inundation from the channel,

groundwater, and direct precipitation. Previous modelling studies suggest ET drives seasonal

differences in the floodplain water table, where soils are saturated during winter/spring when ET

rates are minimal, and the water table is drawn down in the summer when ET rates are high during

Page 58: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

49

the growing season.35 This pattern was experienced during the study period, where soils were

saturated from late fall through mid-summer (Figure 4-2), although this patter was exacerbated by

a wetter-than-normal spring and early summer. Further, the shallow clay layer acts as a confining

layer, which restricts surface water-groundwater interactions and water loss to the subsurface

during the flood experiments. More information about the StREAM Lab and surface water-

groundwater interaction at the site can be found at streamlab.bse.vt.edu and in the companion study

(Hester et al., Surface water-groundwater exchange processes in an experimental floodplain along

a restored headwater stream, submitted 2015),36 respectively.

4.2.2 Experimental Design

We conducted five experimental floods over the course of one year to capture seasonal differences

in floodplain biogeochemical processes: April 8 (spring), June 29 (early summer), August 30 (late

summer), November 11 (fall), and February 7 (winter) across 2013 and 2014. Each experimental

flood lasted 3 hours to simulate natural overbank flood events, where quasi steady-state flow

conditions were achieved in the first two hours and the third hour was used to conduct a nitrogen

uptake experiment. Stream water was pumped into the floodplain slough using a Berkeley B3-

ZRMS irrigation pump at a flow rate of approximately 24 L·s-1. Inlet and outlet flows were

measured using an ultrasonic Fuji M-flow flowmeter and a 7.62 cm parshall flume with an Onset

HOBO Pressure Transducer, respectively. We collected water quality samples at the inlet (e.g., the

pump) at 30 min intervals and at three locations (XS1, XS2, and XS3) along the centerline of flow

in the slough at 15 min intervals (Figure 4-1a). During the first 15 min after water arrived at each

sampling location, we collected samples at 5 min intervals in order to capture the “first flush”

phenomena. At hour two of each experiment, we instantaneously added a solution of NaNO3 and

NaCl tracers to the inlet of the slough. As the tracer slug moved through the site, we collected

samples at 1-3 min intervals at XS2 and XS3. Here, it is important to highlight we pumped stream

water representative of baseflow, not flood flow, into the experimental slough. Because of

differences in background solute concentrations and sediment load, some biogeochemical

processes (e.g. denitrification) may have altered kinetics and processes associated with sediment

deposition (e.g. total phosphorus removal) could not be accurately represented. Further

information about sample analysis and handling can be found in the associated supplemental

material.

4.2.3 Flood Experiment Analyses

We characterized surface-water hydraulics using first-arrival and steady-state residence time

metrics. The first-arrival time represents the time between the pump starting and the start of water

flow at XS3 (e.g., the end of the slough), while steady-state residence time represents the time

from injection of the tracer to the time when the peak of the conservative solute breakthrough

curve passed XS3. Note, the tracer occurred at hour two during quasi steady-state flow conditions.

For each experimental flood, we estimated nutrient retention (or export) using a mass balance in

order to examine the source/sink characteristics of the floodplain slough. Inlet flux was calculated

using measured flow and solute concentrations at the pump, while outlet flux was calculated using

measured flow and solute concentrations at the outlet and XS3, respectfully. Here, we assumed

solute concentrations at XS3 represented solute concentrations at the outlet because of the

development of concentrated flow and proximity between the two locations

We estimated NO3- uptake using the Tracer Analysis for Spiraling Curve Characterization

(TASCC) methodology.37 While this method was designed for stream systems, it has been recently

Page 59: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

50

used within alluvial wetlands where advective flow dominates.38 Essentially, this method utilizes

the difference in the breakthrough curves of the conservative (Cl-) and nonconservative (NO3-)

tracers to estimate instantaneous spiralling metrics (e.g., Newbold et al.,1982)39 and also to

estimate background spiraling metrics through regression. Here, spiraling metrics refer to kinetic

parameters typically used to characterize solute dynamics in streams such as uptake velocity,

uptake length, and areal uptake. The calculated areal uptake (Utot, µg N m-2 min-1) and uptake

velocity (Vtot , m min-1) are applied to Michaelis-Menten kinetics model, and relationships are

developed between NO3- concentrations and spiraling metrics.

To elucidate differences in nitrogen processing across seasons and floodplain location, we

compared results from the five TASCC injections. We completed a nonparametric multivariate

analysis of variance (NPMANOVA)40 where the significance of seasonal variables (e.g., season,

soil temperature, volumetric water content, and background NO3- concentrations), location (e.g.,

cross section), and interaction between seasonal variables and location were tested. We then used

a robust analysis of covariance (robust ANCOVA)41 in a pairwise fashion to test differences in

Michaelis-Menten kinetic models developed for the two cross sections during each event. We

completed all calculations with R Statistical Software42 using the Vegan43 and WRS44 packages.

4.2.4 Reach-Scale Modeling

We estimated annual NO3- removal for the associated 1 km restoration reach using a simple well

mixed-reactor model that incorporated measured areal uptake, a simple inundation model derived

from raster analysis, and a synthetic flow record based on regional regression based on USGS

gaging data and landuse characteristics. Further model details can be found in the supplementary

material. Here, it is important to highlight the parsimonious and conservative nature of the

presented model. Feedbacks between river-floodplain interactions and biogeochemical processing

are very complex and at this point, poorly characterized.12, 45, 46 However, our model estimates

potential load reductions utilizing simple measures and provides a rough estimate of

biogeochemical processing within the StREAM Lab floodplain. While we acknowledge the

uncertainty associated with the resulting load reduction estimates, the estimates provide an

approximate characterization of load reductions in low-order floodplains that can be applied to

restoration projects in the Mid-Atlantic Ridge and Valley.

Table 4-1. Summary of residence time metrics

Flood First-arrival

Time (min)

Steady-State

Residence Time (min)

Spring 12.5 28

Summer 1 12.5 29

Summer 2 31.1 27

Fall 35.0 51

Winter 9.5 38

4.3 Results and Discussion 4.3.1 Inundation Hydrology

Topography, antecedent moisture conditions, and vegetative roughness controlled inundation

hydrology within the floodplain slough. Generally, the flood progression was similar across all

five floods: the slough would fill and then begin to drain back into the adjacent stream at the natural

Page 60: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

51

outlet point down gradient of XS3 (Figure 4-1a). When examining the first-arrival times of the

experimental floods, two groups are apparent: saturated and dry floods. The spring, early summer,

and winter floods all had relatively short first-arrival times (9.5 – 12.5 min) and were inundated

prior to pumping (i.e. saturated floods), whereas the late summer and fall floods had relatively

long first-arrival times (>30 min) and the slough was empty prior to the experiment (i.e. dry

floods). A discussion of how antecedent moisture and vegetation control surface water-

groundwater exchange and water storage in the slough can be found in Hester et al., 2015.36 Similar

patterns in flooding have been documented in the Nyack River floodplain in the western US, where

individual sloughs were characterized as hydrologic facets47 and were the scalable unit used to

model inundation hydrology.48 However, the filling and overflow of the slough does not entirely

characterize water movement within the experimental slough. The lag in conservative tracer signal

(Figure 4-3a), along with visual indicators (e.g. dye tracer and vegetation disturbance), suggest

that a mixing zone between the flood water and local water formed over the course of each

saturated flood, potentially indicating multiple surface-water flowpaths exist within the slough.

Similar to patterns observed by Mertes19 in large river systems, inundation hydrology and mixing

patterns observed in the floodplain were also controlled by antecedent inundation conditions (e.g.,

wet vs dry floods). During the saturated floods, two dominant flowpaths were conceptualized

(Figure 4-1a). Flow through the deepest portion of the slough, where local water was present prior

to flooding, was largely dominated by transient storage and experienced relatively long residence

times. A second primary flowpath formed on the inside of the floodplain slough, which essentially

bypassed the existing local water and was dominated by advective transport (Figure 4-1a). These

flowpaths were conceptualized through visual inspection of both dye tracers and vegetation

disturbance. During these experimental floods, it took 60-75 min for the slough to become

completely mixed as seen in the effluent NO3- signal (Figure 4-3b), suggesting the persistence of

the perirheic zone at the boundary of the two flowpaths. In contrast, the dry floods displayed

relatively uniform effluent signal (Figure 4-3a) and prolonged first-arrival times (Table 4-1)

because the lack of local water prior to flooding. With the available data, it is unclear if a perirheic

zone formed over the course of either dry flood. However, the fall flood had a much longer steady-

state residence time, suggesting relatively uniform flow when compared to the late summer flood

and could be attributed to greater matting of senescing vegetation within the local water column.36

Floodplain conductivity of surface-waters has been described as a measure of the floodplains

ability to route floodwaters, and is a function of the storm hydrograph, floodplain topography, and

vegetative roughness45. However, floodplains exist as a mosaic of habitats49 and therefore

experience a heterogeneous distribution of floodplain conductivity. Within the experimental

slough, the two conceptualized flowpaths experience differences in vegetation density.

Specifically, the deeper portion of the slough has much denser vegetation than the shallow portion,

potentially leading to greater mixing and reduced water velocities at the sediment water-interface.

Therefore, in addition to a decrease in conductance associated with the perirheic zone, the deeper

flowpath could also have decreased conductance because of increased vegetative roughness. As

vegetation conditions varied across the five experimental floods, it is likely the relative

contribution of the two surficial flowpaths also varied. These differences in inundation hydrology,

along with differences in the roughness due plant communities, help explain the difference in

biogeochemical processing between the 5 experimental floods.

Page 61: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

52

Figure 4-3. Inlet and outlet concentrations of dissolved reactive constituents during the five

inundation experiments. Average inlet concentrations are representative of instream solute

concentrations and are shown using bar charts on the left, where error bars represent standard error

from the mean. Outlet concentrations from the first 120 min of each flood are displayed on the

right. Red (plus sign), green(cross), yellow (triangle), orange (hollow circle), and blue (diamond)

are associated with the spring, early summer, late summer, fall, and winter floods, respectively;

outer circles and squares denote saturated and dry experimental floods, respectively.

Page 62: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

53

4.3.2 Dissolved Organic Matter Export

Floodplains are an important source of DOM to riverine networks.11 DOC export from the

floodplain slough was fairly consistent throughout all five floods, where there was a pronounced

first flush of DOC during each flood (Figure 4-3e and 4-4e). Patterns reflective of two distinct

processes appear when analyzing the DOC concentration within the floodplain effluent: dilution

and flushing. The saturated floods experienced a dilution of DOC, where mixing between stream

water and local water is apparent in the gradual decline in effluent concentration (Figure 4-3e). In

contrast, the dry floods experienced flushing of DOC from the soil surface within the slough,50

where the signal exhibited a rapid decline in effluent concentration. The experimental slough was

a relatively large source of DOC during the saturated floods, where effluent DOC mass was 30-

60% greater than the influent mass. However, the dry floods exported relatively little or no DOC

export. Our results are consistent with other observations of accumulation of DOC in inundated

floodplains18, and could be attributed to decay of algal biomass25 and plant material.13 Because of

the dominance of reed canary grass within the study slough, it is likely a dominant source of

organic matter and autochthonous nutrients.51 Further, our results suggest inundation hydrology,

and more specifically, antecedent inundation conditions controls DOC export and processing

within the experimental slough (Figure 4-4e).

4.3.3 Initial Flush of Reactive Solutes

Similar to other studies, the experimental slough was a consistent net source of SRP during each

experimental flood. Potential sources of SRP include organic matter, likely derived from reed

canary grass decay51; nutrient rich sediments and particulate organic matter deposition from

previous floods14,52; and the legacy of over 100 years of cattle grazing at the site. Possible

mechanisms responsible for SRP loss from floodplains include the oxidation of DOM,53 reduction

of iron-phosphate sediment,54 and desorption of loosely held SRP- at the soil surface52. However,

it is important to note that we did not capture depositional processes during the flood experiments,

and thus, were unable to characterize total phosphorus removal and sorption/desportion of SRP

associated with newly deposited sediments.

Here, we propose that both hydrologic and biotic processes control SRP export from the slough.

In contrast to the DOC signal, the dry floods had relatively large SRP first flushes when compared

to the saturated floods (Figure 4-3 and 4-4). During the interflood period (e.g., no overbank or

artificial flow), organic matter at the soil surface would be available for mineralization and the

production of loosely held SRP;14 thus, this mechanism is likely a dominant source of SRP during

the dry floods. While the saturated floods did not experience a strong first flush, they were still net

sources of SRP. This is likely attributed to redox conditions that developed during prolonged

inundation, where ferrous soils release SRP when reduced,55 however, thorough analysis of soil

chemistry was not conducted across the study slough to identify the persistance of ferrous soils.

In addition to hydrologic control (e.g., dry vs saturated antecedent inundation conditions), SRP

loads into and out of the floodplain varied greatly with season (Figure 4-4d) and are likely linked

to biological uptake within the floodplain.20 During the spring flood, the growing season was just

beginning when P is in high demand. Thus, available SRP was likely being utilized for plant

biological activity. Plant and microbial uptake has been widely observed in streams and associated

riverine wetlands,56 and periphyton can act as a temporary sink that completely masks

Page 63: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

54

Figure 4-4. The total load observed at the inlet (dark) and outlet (light) of the floodplain slough

during the 5 inundation experiments. Dry floods are denoted by bars with hatching, and both the

relative difference (%) and mass difference (g) between the loads is also displayed.

Page 64: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

55

P signals.57 This is a possible explanation of the minimal SRP load experienced within the spring

flood, and is in contrast to the fall flood, where much of the plant material was senescing and

possibly contributed to the large net export of SRP through mineralization of newly senesced

material. Other restored floodplains with legacy nutrients have been shown to be sources of SRP

for 20+ years post restoration.58

Over the course of the five experimental floods, the slough was a variable source/sink of NH4+.

During the fall flood, there was an observable first flush of both NH4+ and SRP (Figures 3b-c),

suggesting elevated mineralization of DOM. However, there were no apparent seasonal or

hydrologic drivers behind the NH4+ signal observed in the other four floods. This is consistent with

investigation of other headwater floodplains in both the Mid-Atlantic20 and upper Mississippi

River Valley,32 where total export varied with season and by event, respectively. NH4+ processing

within the floodplain environments is relatively complicated because of the lability of NH4+, its

affinity for sorption, and its tendency to nitrify in oxic environments.17 Noe et al., 201314 examined

mineralization rates across a gradient of floodplains in the Mid-Atlantic and found that nitrification

was dominant in headwater floodplains while ammonification was dominant in larger river

floodplains. While nitrification was not explicitly measured in this study, variability in nitrification

rates could explain the variable source/sink pattern experienced within the experimental slough.

4.3.4 Nitrate Removal

The floodplain was typically a sink of NO3-, where net removal ranged from 2 to 26%

(Figure 4-4b). Similar to the DOC effluent signal, perirheic mixing was evident as the NO3--rich

stream water mixed with the NO3- limited local water during the saturated floods (Figure 4-3b).

When comparing DOC and NO3- effluent concentrations, the DOC signal reached equilibrium

more quickly, suggesting NO3- removal through both biotic uptake and denitrification. Forshay

and Stanley23 observed similar removal of NO3- when river water mixed with local water in a

backwater floodplain, where rapid NO3- loss was attributed to denitrification. Denitrification is a

well-documented process within riverine floodplains, with recorded removal rates ranging from

undectable59 to greater than 75%29 removal of the total riverine NO3- load. Denitrification is

controlled by the amount of NO3- transported into the floodplain,30 the availability of carbon,11 and

the residence time of floodwaters.48 In many backwater floodplains, denitrification is often limited

by the amount of NO3- transported into floodplain environment, suggesting limited exchange

between channel and floodplain.60 In contrast, in environments where river-floodplain interaction

is greater and advective flow dominates in the floodplain environment, denitrification is often

limited by short residence times.59 This was likely the case in our study slough, where mean

residence times ranged from 27 to 51 min, which greatly contrasts floodplain dominated systems

like the Amazon and Danube Rivers where mean residences times can be months.61, 62

Areal uptake (Figure 4-5) was a function of both seasonality, and to a lesser extent, heterogeneity

of flowpaths within the floodplain. The NPMANOVA model accounted for 52% of variation

(R2=0.52, p<0.001) in the relationship between total areal uptake and NO3- concentration across

all five tracer experiments at both cross sections. Spatially, sampling location (e.g., XS2 vs XS3)

only accounted for 1.7% of the observed variation (p=0.01). This suggests while sampling location

was a statistically significant predictor, it is practically unimportant and the observed nitrate uptake

signal was relatively homogeneous at the spatial resolution of our sampling. Therefore, spiraling

metrics were calculated using data from both XS2 and XS3. Seasonal variables accounted for 33%

Page 65: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

56

of variation (p<0.001) in the model, and upon closer inspection, soil moisture, background NO3-

concentration, and soil temperature account for 15%,10%, and 2.5% of the observed variation

(p<0.001), respectively. Antecedent moisture conditions affect floodwater residence times (e.g.

dry vs saturated floods), can limit or enhance biological activity and uptake21, and control

mechanisms associated with the first flush (e.g. oxidation of organic matter)16. Additionally,

background NO3- concentrations have been shown to exert influence on NO3

- uptake rates, where

higher NO3- concentrations typically lead to greater NO3

- uptake but lower net removal.63, 64

However, because soil moisture and background NO3- concentration are confounding variables, it

is difficult to isolate their individual effects on the NO3- uptake with the available data. For

example, the two late summer and fall floods were both dry floods and experiences similar

background NO3- concentrations. However, the uptake rates were significantly different,

suggesting there was a shift in dominant processes affecting export.

Table 4-2. Summary of Spiraling metrics and associated statistics from tracer addition.

Experiment Background Uptake Length Michaelis-Menten Model Groups

Wilcox ANCOVA

Sw,amb

(m) Std Error

(m) p-Value

Umax (µg N m-2 min-1)

Km (µg N m-2 min-1)

p-Value

Spring 116 21.4 <0.001 2865 39.9 <0.001 A Early Summer 258 25.9 <0.001 1256 46.7 <0.001 B Late Summer 157 7.9 <0.001 2135 70.2 <0.001 C

Fall 319 44.2 <0.001 1255 123 <0.001 B Winter 304 176 0.12 1483 73.9 0.002 B

Utilizing a Wilcox ANCOVA in a pairwise fashion, the five floods separated into three statistically

different groups: 1) spring, 2) late summer, and 3) winter, early summer, and fall (Table 4-2).

Increased biological uptake is expected in the spring and summer and conversely muted uptake in

the fall and winter because of seasonal differences in temperature and its effects on the biological

communities. However, this pattern is confounded by the low uptake rates experienced in the early

summer floods. Potentially, the low uptake rates measured in the early summer flood were related

to elevated NH4+ export (Figure 4-4d), where highly labile NH4

+ satisfied N-demand within the

slough and reduced rates of NO3- uptake. Similarly, Noe and Hupp20 did not find seasonal patterns

in NO3- export from a small floodplain system, suggesting confounding factors such as nitrification

and organic matter availability/lability also controlled the export of NO3- from the floodplain.

When compared to typical instream areal NO3- uptake,65 NO3

- uptake rates were relatively high

within the slough during the five experimental floods (Figure 4-5). In a recent national assessment

of instream NO3- uptake kinetics, the LINXII study measured total instream NO3

- uptake using N15

additions across 56 different streams in North America. Represented by the dark line in Figure 4-5,

the regression model of LINXII uptake rates is lower than the rates measured across the five

experimental floods. This could partially be explained by differences in measurement techniques,

where the bulk injection alters instream kinetics because of the large increase in NO3-

concentrations.66 Further, the background NO3- concentration of the stream water used to inundate

the floodplain (Figure 4-3) was not necessarily representative of floodwater NO3- concentrations.

However, the response to the NO3- addition was minimal during all five experimental floods (e.g.,

muted rising limbs in the Michaelis-Menten curves), suggesting biogeochemical processing within

Page 66: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

57

the floodplain was not limited by NO3- , where the current N pool is derived from internal cycling

of nutrients and potentially legacy N from the 100+ years of grazing at the site.51, 67 Increased

biogeochemical processing within the floodplain can be explained by increased contact between

the floodplain soils and floodwater, increase in availability of labile organic matter, and also

increase in biological uptake.

Figure 4-5. Areal NO3- uptake calculated for spring (red, plus sign), early summer (green, cross),

late summer (yellow, triangle), fall (orange, circle), and winter (blue, diamond) experimental

floods. Points represent individual samples and lines represent Michaelis-Menten Kinetic models

developed for each flood, respectively. The black dashed line represents expected total instream

uptake based on data taken from the 58 streams across North America in the LINXII study.65

Statistical differences denoted in Table 4-2 are displayed using letters A-C.

4.3.5 Modeled Load Reductions

Annual and storm NO3- load reductions were minimal when measured uptake was extrapolated to

the restored floodplain across the synthetic flow record. For individual events, NO3- removal

ranged from 8.6 to 17.6 kg of NO3-N removed per storm, or 7.5 to 10.3% removal of the total

storm load. However, this resulted in annual removal ranging from 0 to 139 kg of NO3-N removed,

or 0 to 1.5% of the annual load of NO3-. Mean individual storm and annual load reductions were

10% and 0.60%, respectively. The relatively simple model is conservative, in that it over estimates

inundation time through a rectangular hydrograph, it assumes relatively high NO3- concentrations

during storm flow (e.g., no dilution affects were accounted for), and the maximum observed areal

uptake (spring, 2000 µg N m-2 min-1) was used for the estimate of NO3- removal. This suggests the

restored floodplain is ineffective in removal of instream NO3- and can be explained by both limited

river-floodplain connectivity and short residence times associated with short-hydroperiod

floodplains. Roley et al., 201263 found similar load reduction associated with two-stage ditch

located in the Midwestern US, where NO3- uptake was relatively high but hydrologic connectivity

and residence time were relatively low. They went on to conclude that to maximize nitrogen

removal, floodplain restoration design should enhance removal of NO3- from waters draining to

the site (e.g. upslope water). In addition, it is also important to note the modeled stream reach in

Page 67: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

58

our study represents a relatively small section of stream, and riparian buffers/wetlands have been

shown to remove significant loads of NO3- when the entire basin is taken into account.68 Here, if

we assume that 1 km of stream removes 10% of the NO3- load for individual storms, then it would

take approximately 10 km of restored floodplain to reduce the entire nitrate load being delivered

to the StREAM Lab for an average storm. While this is a vast oversimplification of

biogeochemical processing of floodwaters within floodplains, it does highlight the importance of

the cummulative effects of riparian/floodpalin systems along the stream cooridor.

4.3.6. Management Implications

Recent efforts by the Chesapeake Bay Program (CBP) have emphasized re-establishing

river-floodplain connectivity to improve downstream water quality, essentially making river-

floodplain connectivity a best management practice (BMP) for the stream restoration industry in

the eastern US.4 Here, we do not want to discount ecosystem services derived from restoration

efforts in headwater streams, the water quality benefits riparian zones and floodplains provide by

processing upslope water, or even the cumulative influence of floodplains across the river network.

However, results from studies in the Upper Midwest,32 Mid-Atlantic Piedmont,20, 27 and now the

Appalachian Ridge and Valley (i.e. this study) show that individual floodplains associated with

small to medium size streams can actually be net sources of nutrients. Therefore, these results

cumulatively suggest that restoring river-floodplain connectivity is not always an appropriate BMP

for nutrient removal, and site specific conditions such as annual inundation duration and legacy

landuse should be considered when optimizing river-floodplain connectivity for nutrient removal

and retention. Potentially, these finding would be most useful during site selection and

prioritization phase of restoration design, where multiple objectives and design parameters are

often balanced. Further, our results highlight the potential for flushing of reactive SRP from

floodplain wetlands, and more specifically, stream-wetland complexes often used for mitigation

banking and sediment removal.69 While these engineered wetlands undoubtedly remove sediment

attached P and N through deposition and denitrification,70 respectively, designers must also

balance flushing of allochthonous SRP and other reactive solutes.

4.4 Acknowledgements We want to thank the students in the StREAM Lab REU Program (NSF-EEC-REU 1156688) for

their insights and assistance in piloting and conducting the artificial floods. Specifically, Katy

Hofmeister, Tyler Weiglein, and Dylan Cooper were critical to the success of these experiments.

Further, we thank the National Science Foundation (ENG-CBET-1066817), and the Virginia Tech

Institute for Critical Technology and Applied Sciences (ICTAS) for their support. Opinions

expressed here are those of the authors and not necessarily those of the NSF or ICTAS.

4.5 Works Cited 1. Bernhardt, E. S.; Palmer, M. A.; Allan, J. D.; Alexander, G.; Barnas, K.; Brooks, S.; Carr, J.;

Clayton, S.; Dahm, C.; Follstad-Shah, J.; Galat, D.; Gloss, S.; Goodwin, P.; Hart, D.; Hassett, B.;

Jenkinson, R.; Katz, S.; Kondolf, G. M.; Lake, P. S.; Lave, R.; Meyer, J. L.; O'Donnell, T. K.;

Pagano, L.; Powell, B.; Sudduth, E. Ecology - Synthesizing US river restoration efforts. Science

2005, 308 (5722), 636-637.

2. Lave, R.; Doyle, M.; Robertson, M. Privatizing stream restoration in the US. Soc Stud Sci 2010,

40 (5), 677-703.

Page 68: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

59

3. USEPA Watershed Assessment, Tracking & Environmnetal Results.

http://www.epa.gov/waters/ir/

4. Harman, W.; Starr, R.; Carter, M.; Tweedy, K.; Clemmons, M.; Suggs, K.; Miller, C. A function-

based framework for stream assessment and restoration projects; EPA 843-K-12-006; US

Environmental Protection Agency, Office of Wetlands, Oceans, and Watersheds, Washington,

DC: 2012.

5. Bernhardt, E. S.; Sudduth, E. B.; Palmer, M. A.; Allan, J. D.; Meyer, J. L.; Alexander, G.;

Follastad-Shah, J.; Hassett, B.; Jenkinson, R.; Lave, R.; Rumps, J.; Pagano, L. Restoring rivers

one reach at a time: Results from a survey of US river restoration practitioners. Restoration

Ecology 2007, 15 (3), 482-493.

6. Kondolf, G. M.; Boulton, A. J.; O'Daniel, S.; Poole, G. C.; Rachel, F. J.; Stanley, E. H.; Wohl,

E.; Bang, A.; Carlstrom, J.; Cristoni, C.; Huber, H.; Koljonen, S.; Louhi, P.; Nakamura, K. Process-

based ecological river restoration: Visualizing three-dimensional connectivity and dynamic

vectors to recover lost linkages. Ecol Soc 2006, 11 (2), 5.

7. Tockner, K.; Stanford, J. A. Riverine flood plains: Present state and future trends. Environmental

Conservation 2002, 29 (3), 308-330.

8. Kaushal, S. S.; Groffman, P. M.; Band, L. E.; Shields, C. A.; Morgan, R. P.;Palmer, M. A.;Belt,

K. T.;Swan, C. M.;Findlay, S. E. G.;Fisher, G.T. Interaction between urbanization and climate

variability amplifies watershed nitrate export in Maryland, Environ. Sci. Technol. 2008, 42 (16),

5872-5878.

9. Vidon, P.; Allan, C.; Burns, D.; Duval, T. P.; Gurwick, N.; Inamdar, S.; Lowrance, R.; Okay,

J.; Scott, D.; Sebestyen, S. Hot spots and hot moments in riparian zones: potential for improved

water quality management1. J. Am. Water Resour. Assoc. 2010, 46 (2), 278-298.

10. McClain, M. E.; Boyer, E. W.; Dent, C. L.; Gergel, S. E.; Grimm, N. B.; Groffman, P. M.;

Hart, S. C.; Harvey, J. W.; Johnston, C. A.; Mayorga, E.; McDowell, W. H.; Pinay, G.

Biogeochemical hot spots and hot moments at the interface of terrestrial and aquatic ecosystems.

Ecosystems 2003, 6 (4), 301-312.

11. Tockner, K.; Pennetzdorfer, D.; Reiner, N.; Schiemer, F.; Ward, J. V. Hydrological

connectivity, and the exchange of organic matter and nutrients in a dynamic river-floodplain

system (Danube, Austria). Freshw. Biol. 1999, 41 (3), 521-535.

12. Jones, C. N.; Scott, D. T.; Edwards, B. L.; Keim, R. F. Perirheic mixing and biogeochemical

processing in flow-through and backwater floodplain wetlands. Water Resour. Res. 2014, 50 (9),

7394-7405.

13. Chow, A. T.; Dai, J. N.; Conner, W. H.; Hitchcock, D. R.; Wang, J. J. Dissolved organic matter

and nutrient dynamics of a coastal freshwater forested wetland in Winyah Bay, South Carolina.

Biogeochemistry 2013, 112 (1-3), 571-587.

14. Noe, G. B.; Hupp, C. R.; Rybicki, N. B. Hydrogeomorphology influences soil nitrogen and

phosphorus mineralization in floodplain wetlands. Ecosystems 2013, 16 (1), 75-94.

Page 69: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

60

15. Wolf, K. L.; Noe, G. B.; Ahn, C. Hydrologic connectivity to streams increases nitrogen and

phosphorus inputs and cycling in soils of created and natural floodplain wetlands. Journal of

Environmental Quality 2013, 42 (4), 1245-1255.

16. Baldwin, D. S.; Mitchell, A. M. The effects of drying and re-flooding on the sediment and soil

nutrient dynamics of lowland river-floodplain systems: A synthesis. Regul. Rivers-Res. Manage.

2000, 16 (5), 457-467.

17. Jicha, T. M.; Johnson, L. B.; Hill, B. H.; Regal, R. R.; Elonen, C. M.; Pearson, M. S. Spatial

and temporal patterns of nitrification rates in forested floodplain wetland soils of upper Mississippi

River pool 8. River Research and Applications 2013, 30 (5), 650-662.

18. Owens, P. N.; Walling, D. E. The phosphorus content of fluvial sediment in rural and

industrialized river basins. Water Research 2002, 36 (3), 685-701.

19. Mertes, L. A. K. Documentation and significance of the perirheic zone on inundated

floodplains. Water Resour. Res. 1997, 33 (7), 1749-1762.

20. Noe, G. B.; Hupp, C. R. Seasonal variation in nutrient retention during inundation of a short-

hydroperiod floodplain. River Research and Applications 2007, 23 (10), 1088-1101.

21. Bechtold, J. S.; Edwards, R. T.; Naiman, R. J, Biotic versus hydrologic control over seasonal

nitrate leaching in a floodplain forest. Biogeochemistry 2003, 63 (1), 53-71.

22. Vymazal, J. Removal of nutrients in various types of constructed wetlands. Sci Total Environ

2007, 380 (1–3), 48-65.

23. Forshay, K. J.; Stanley, E. H. Rapid nitrate loss and denitrification in a temperate river

floodplain. Biogeochemistry 2005, 75 (1), 43-64.

24. Friedrich, J.; Dinkel, C.; Grieder, E.; Radan, S.; Secrieru, D.; Steingruber, S.; Wehrli, B.

Nutrient uptake and benthic regeneration in Danube Delta lakes. Biogeochemistry 2003, 64 (3),

373-398.

25. Hein, T.; Baranyi, C.; Herndl, G. J.; Wanek, W.; Schiemer, F. Allochthonous and

autochthonous particulate organic matter in floodplains of the River Danube: the importance of

hydrological connectivity. Freshw. Biol. 2003, 48 (2), 220-232.

26. Kadlec, R. H.; Reddy, K. Temperature effects in treatment wetlands. Water Environ. Res. 2001,

543-557.

27. Harrison, M.; Groffman, P.; Mayer, P.; Kaushal, S. Nitrate removal in two relict oxbow urban

wetlands: a 15N mass-balance approach. Biogeochemistry 2012, 111 (1-3), 647-660.

28. Welti, N.; Bondar-Kunze, E.; Singer, G.; Tritthart, M.; Zechmeister-Boltenstern, S.; Hein, T.;

Pinay, G. Large-scale controls on potential respiration and denitrification in riverine floodplains.

Ecol Eng 2012, 42, 73-84.

Page 70: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

61

29. Scott, D. T.; Keim, R. F.; Edwards, B. L.; Jones, C. N.; Kroes, D. E. Floodplain biogeochemical

processing of floodwaters in the Atchafalaya River Basin during the Mississippi River flood of

2011. J. Geophys. Res.-Biogeosci. 2014, 119 (4), 537-546.

30. Racchetti, E.; Bartoli, M.; Soana, E.; Longhi, D.; Christian, R. R.; Pinardi, M.; Viaroli, P.

Influence of hydrological connectivity of riverine wetlands on nitrogen removal via denitrification.

Biogeochemistry 2011, 103 (1-3), 335-354.

31. Huber, B.; Luster, J.; Bernasconi, S. M.; Shrestha, J.; Pannatier, E. G. Nitrate leaching from

short-hydroperiod floodplain soils. Biogeosciences 2012, 9 (11), 4385-4397.

32. Kreiling, R. M.; Schubauer-Berigan, J. P.; Richardson, W. B.; Bartsch, L. A.; Hughes, P. E.;

Cavanaugh, J. C.; Strauss, E. A. Wetland management reduces sediment and nutrient loading to

the upper Mississippi River. J. Environ. Qual. 2013, 42 (2), 573-583.

33. Jin, S.; Yang, L.; Danielson, P.; Homer, C.; Fry, J.; Xian, G. A comprehensive change detection

method for updating the National Land Cover Database to circa 2011. Remote Sens Environ 2013

(132), 159-175.

34. Thomspson, T. W.; Hession, W. C.; Scott, D. T. The StREAM Lab at Virginia Tech. Resource

Magazine 2012, 8-9.

35. Azinheira, D. L.; Scott, D. T.; Hession, W.; Hester, E. T. Comparison of effects of inset

floodplains and hyporheic exchange induced by in-stream structures on solute retention. Water

Resour. Res. 2014, 50 (7), 6168-6190.

36. Hester, E.T.; Guth, C.R.; Scott, D.T.; Jones, C.N. Surface water-groundwater exchange

processes in an experimental floodplain along a restored headwater stream. In Review. 2015.

37. Covino, T. P.; McGlynn, B. L.; McNamara, R. A. Tracer Additions for Spiraling Curve

Characterization (TASCC): Quantifying stream nutrient uptake kinetics from ambient to

saturation. Limnology and Oceanography-Methods 2010, 8 (9), 484-498.

38. Wollheim, W. M.; Harms, T. K.; Peterson, B. J.; Morkeski, K.; Hopkinson, C. S.; Stewart, R.

J.; Gooseff, M. N.; Briggs, M. A. Nitrate uptake dynamics of surface transient storage in stream

channels and fluvial wetlands. Biogeochemistry 2014, 120 (1-3), 239-257.

39. Newbold, J. D.; O'Neill, R. V.; Elwood, J. W.; Winkle, W. V. Nutrient spiralling in streams:

implications for nutrient limitation and invertebrate activity. Am. Nat. 1982, 120 (5), 628-652.

40. Anderson, M. J. A new method for non-parametric multivariate analysis of variance. Austral

Ecol 2001, 26 (1), 32-46.

41. Wilcox, R. Robust multivariate regression when there is heteroscedasticity. Commun Stat

Simulat. 2009, 38 (1), 1-13.

42. R Core Development Team. R: A language and environment for statistical computing, R

Foundation for Statistical Computing: Vienna, Austria, 2013.

Page 71: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

62

43. Oksanen, J.; Blanchet, F. G.; Kindt, R.; Legendre, P.; Minchin, R. R.; O'Hara, R. B.; Simpson,

G. L.; Solymos, P.; Stevens, M. H. H.; Wagner, H. Vegan: Community Ecology Package.

Retreived from: http://CRAN.R-project.org/package=vegan: 2013.

44. Wilcox, R.R., & Schönbrodt, F.D. The WRS package for robust statistics in R (version 0.24).

Retrieved from http://r-forge.r-project.org/projects/wrs/: 2014.

45. Hughes, D. A. Floodplain inundation - processes and relationships with channel discharge.

Earth Surf. Process. Landf. 1980, 5 (3), 297-304.

46. Knight, D. River flood hydraulics: Theoretical issues and stage-discharge relationships. River

basin modelling for flood risk mitigation 2006, 17, 301-334.

47. Jones, K. L.; Poole, G. C.; O'Daniel, S. J.; Mertes, L. A. K.; Stanford, J. A. Surface hydrology

of low-relief landscapes: Assessing surface water flow impedance using LIDAR-derived digital

elevation models. Remote Sens Environ 2008, 112 (11), 4148-4158.

48. Helton, A. M.; Poole, G.; Payn, r.; Izurieta, C.; Stanford, J. Relative influences of the river

channel, floodplain surface, and alluvial aquifer on simulated hyrologic residence time in a

montane river floodplain. Geomorphology 2014, 205, 17-26.

49. Thorp, J. H.; Thoms, M. C.; Delong, M. D. The riverine ecosystem synthesis: biocomplexity

in river networks across space and time. River Research and Applications 2006, 22(2),123-147.

50. Fellman, J. B.; Hood, E.; Edwards, R. T.; D'Amore, D. V. Changes in the concentration,

biodegradability, and fluorescent properties of dissolved organic matter during stormflows in

coastal temperate watersheds. J. Geophys. Res.-Biogeosci. 2009, 114 (G1).

51. Kreiling, R.; De Jager, N.; Swanson, W.; Strauss, E.; Thomsen, M., Effects of flooding on ion

exchange rates in an upper Mississippi River floodplain forest impacted by herbivory, invasion,

and restoration. Wetlands 2015, 1-8.

52. Dupas, R.; Gruau, G.; Gu, S.; Humbert, G.; Jaffrézic, A.; Gascuel-Odoux, C., Groundwater

control of biogeochemical processes causing phosphorus release from riparian wetlands. Water

Research 2015, 84, 307-314.

53. Noe, G. B.; Hupp, C. R. Carbon, nitrogen, and phosphorus accumulation in floodplains of

Atlantic Coastal Plain rivers, USA. Ecological Applications 2005, 15 (4), 1178-1190.

54. Ardon, M.; Montanari, S.; Morse, J. L.; Doyle, M. W.; Bernhardt, E. S. Phosphorus export

from a restored wetland ecosystem in response to natural and experimental hydrologic fluctuations.

J. Geophys. Res.-Biogeosci. 2010, 115 (G4).

55. Reckendorfer, W.; Funk, A.; Gschopf, C.; Hein, T.; Schiemer, F. Aquatic ecosystem functions

of an isolated floodplain and their implications for flood retention and management. J. Appl. Ecol.

2013, 50 (1), 119-128.

56. Hoffmann, C. C.; Kjaergaard, C.; Uusi-Kamppa, J.; Hansen, H. C. B.; Kronvang, B.

Phosphorus retention in riparian buffers: review of their efficiency. Journal of Environmental

Quality 2009, 38 (5), 1942-1955.

Page 72: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

63

57. Jarvie, H. P.; Sharpley, A. N.; Scott, J. T.; Haggard, B. E.; Bowes, M. J.; Massey, L. B. Within-

River Phosphorus Retention: Accounting for a Missing Piece in the Watershed Phosphorus Puzzle.

Environ. Sci. Technol. 2012, 46 (24), 13284-13292.

58. Surridge, B. W. J.; Heathwaite, A. L.; Baird, A. J. Phosphorus mobilisation and transport

within a long-restored floodplain wetland. Ecol Eng 2012, 44, 348-359.

59. McJannet, D.; Wallace, J.; Keen, R.; Hawdon, A.; Kemei, J. The filtering capacity of a tropical

riverine wetland: II. Sediment and nutrient balances. Hydrological Processes 2012, 26 (1), 53-72.

60. BryantMason, A.; Xu, Y.-J.; Altabet, M. A. Limited capacity of river corridor wetlands to

remove nitrate - A case study on the Atchafalaya River Basin during the 2011 Mississippi River

flooding. Water Resourc. Res. 2012, 49(1), 283-290.

61. Bonnet, M. P.; Barroux, G.; Martinez, J. M.; Seyler, F.; Moreira-Turcq, P.; Cochonneau, G.;

Melack, J. M.; Boaventura, G.; Maurice-Bourgoin, L.; León, J. G.; Roux, E.; Calmant, S.; Kosuth,

P.; Guyot, J. L.; Seyler, P. Floodplain hydrology in an Amazon floodplain lake (Lago Grande de

Curuaí). J. Hydrol. 2008, 349 (1–2), 18-30.

62. Schagerl, M.; Drozdowski, I.; Angeler, D. G.; Hein, T.; Preiner, S. Water age - a major factor

controlling phytoplankton community structure in a reconnected dynamic floodplain (Danube,

Regelsbrunn, Austria). Journal of Limnology 2009, 68 (2), 274-287.

63. Roley, S. S.; Tank, J. L.; Stephen, M. L.; Johnson, L. T.; Beaulieu, J. J.; Witter, J. D. Floodplain

restoration enhances denitrification and reach-scale nitrogen removal in an agricultural stream.

Ecol. Appl. 2012, 22 (1), 281-297.

64. Mulholland, P. J.; Helton, A. M.; Poole, G. C.; Hall, R. O.; Hamilton, S. K.; Peterson, B. J.;

Tank, J. L.; Ashkenas, L. R.; Cooper, L. W.; Dahm, C. N.; Dodds, W. K.; Findlay, S. E. G.;

Gregory, S. V.; Grimm, N. B.; Johnson, S. L.; McDowell, W. H.; Meyer, J. L.; Valett, H. M.;

Webster, J. R.; Arango, C. P.; Beaulieu, J. J.; Bernot, M. J.; Burgin, A. J.; Crenshaw, C. L.;

Johnson, L. T.; Niederlehner, B. R.; O'Brien, J. M.; Potter, J. D.; Sheibley, R. W.; Sobota, D. J.;

Thomas, S. M., Stream denitrification across biomes and its response to anthropogenic nitrate

loading. Nature 2008, 452, (7184), 202-U46.

65. Hall, R. O.; Tank, J. L.; Sobota, D. J.; Mulholland, P. J.; O'Brien, J. M.; Dodds, W. K.; Webster,

J. R.; Valett, H. M.; Poole, G. C.; Peterson, B. J.; Meyer, J. L.; McDowell, W. H.; Johnson, S. L.;

Hamilton, S. K.; Grimm, N. B.; Gregory, S. V.; Dahm, C. N.; Cooper, L. W.; Ashkenas, L. R.;

Thomas, S. M.; Sheibley, R. W.; Potter, J. D.; Niederlehner, B. R.; Johnson, L. T.; Helton, A. M.;

Crenshaw, C. M.; Burgin, A. J.; Bernot, M. J.; Beaulieu, J. J.; Arango, C. P. Nitrate removal in

stream ecosystems measured by N-15 addition experiments: Total uptake. Limnol. Oceanogr.

2009, 54 (3), 653-665.

66. Earl, S. R.; Valett, H. M.; Webster, J. R. Nitrogen spiraling in streams: comparisons between

stable isotope tracer and nutrient addition experiments. Limnol. Oceanogr. 2007, 52, 1718-1723.

67. Covino, T.; McGlynn, B.; McNamara, R. Land use/land cover and scale influences on in-

stream nitrogen uptake kinetics. J. Geophys. Res.-Biogeosci. 2012, 117 (G2).

Page 73: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

64

68. Clement, J. C.; Aquilina, L.; Bour, O.; Plaine, K.; Burt, T. P.; Pinay, G. Hydrological flowpaths

and nitrate removal rates within a riparian floodplain along a fourth-order stream in Brittany

(France). Hydrol Process 2003, 17 (6), 1177-1195.

69. Filoso, S.; Smith, S. M. C.; Williams, M. R.; Palmer, M. A., The efficacy of constructed

stream–wetland complexes at reducing the flux of suspended solids to Chesapeake Bay. Environ.

Sci. Technol. 2015, 49, (15), 8986-8994.

70. Hernandez, M. E.; Mitsch, W. J., Denitrification in created riverine wetlands: Influence of

hydrology and season. Ecol Eng 2007, 30, (1), 78-88.

Page 74: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

65

CHAPTER S4. SUPPLEMENTAL INFORMATION

S4.1 Climatic Data Acquisition Climatic Data was collected at the site using the StREAM Lab weather station that has been

deployed since 2011. Hourly evapotranspiration (ET) was obtained using standard procedures

associated with the Bowen Ratio estimation69,70. Relevant measurements were taken hourly using

CS300-QD solar radiation sensors, HC2S3-QD air temperature and relative humidity probes,

CS106 barometric pressure sensor, and a WindSonic4-QD 2-D sonic wind sensor. Hourly rainfall

was recorded using a TE25-QD tipping bucket. Volumetric water content (%) was measured

within the slough using a Decagon GS3 soil moisture probes approximately 10 cm below the soil

surface. Stage in Stroubles Creek was recorded using a stilling basin and Campbell Scientific

CS450 pressure transducer and converted to flow using a 2-stage rating curve developed for the

site. Climatic data can be viewed real time at http://streamlab.bse.vt.edu/.

S4.2 Water Quality Sample Handling and Analysis All water quality samples were collected in 1 L plastic bottles and filtered in the field within 30

min of collection using a peristaltic pump and 0.45 µM Geotech capsule filter. Subsamples for

nutrient analysis were stored in the dark at 4C in 250 mL plastic bottles, while subsamples for

DOC analysis were stored in the dark at 4C in 40 mL pre-combusted amber glass vials. Processed

samples were transported to the VT-BSE Water Quality Lab, stored at 4C, and analyzed within 5

days. Analyses for NH4+ and SRP were completed using a SEAL Analytical AA3 analyzer using

the Berthelot reaction (saliciylate alternative) and the Murphy and Riley method, respectively; Cl-

and NO3- were analyzed using Dionex ICS-3000 ion chromatograph; and DOC and TDN were

measured using high temperature oxidative combustion with a Shimadzu TOC-V CPH.

S4.3 Estimates of Reach-Scale Load Reductions Annual NO3

- removal for the associated restoration reach was estimated using a simple well mixed

reactor model that incorporated measured areal uptake, a simple inundation model, and the

development of synthetic flow record. Here, it is important to highlight the parsimonious and

conservative nature of the presented model. Feedbacks between river-floodplain interactions and

biogeochemical processing are very complex and at this point, poorly characterized18, 44-45.

However, our model estimates potential load reductions utilizing simple measures and provides a

back-of-the-envelope estimate of biogeochemical processing within the StREAM Lab floodplain.

While we acknowledge the uncertainty associated with the resulting load reduction estimates, the

estimates provide an approximate characterization of load reductions in low-order floodplains that

can be applied to restoration projects in the Mid-Atlantic Ridge and Valley.

To estimate load reductions, the StREAM Lab reach was conceptualized as a completely mixed

reactor across a synthetic flow record. For individual storms, mass removal rate (MRemoval , µg N

m-2 min-1) was estimated simply by:

Page 75: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

66

removal tot i stormM U A d Equation S4-1

where Utot is total areal uptake, Ai is the estimated area of inundation, and dstorm is the estimated

storm duration. Utot is defined by the highest observed areal uptake observed during the during the

tracer experiments (2800 µg N m-2 min-1). The restored reach, and associated floodplain, serves

as the modeled control volume. For simplicity, Cin was held constant at 100 µM NO3-. While this

concentration is consistent for storm flow observations in the StREAM Lab, it does not take into

account seasonal differences in NO3- concentrations or dilution typically observed during storm

events. Therefore, it is important to highlight that our model estimates characterize potential

removal.

Ai was defined using a simple inundation model developed using methods similar to Jones et al.,

2008.46 A 3 meter digital elevation model (DEM) was developed for the site utilizing airborne

LiDAR data provided from the Town of Blacksburg. The DEM was then normalized to stream

slope by subtracting a “stream-slope” raster from the DEM. The stream slope raster was developed

by using inverse-distance weighting interpolation of surface-water elevations and applied across

the extent of the original DEM. Then, conditional raster algebra was utilized to identify inundated

areas for a given stage (Figure S4-1a), and stage was converted to flow using a rating curve

developed for the site. This resulted in relationship between Ai and Q (Figure S4-1b).

Figure S4-1. (Left) Map of estimated inundation experienced across the StREAM Lab Floodplain

at a streamflow of 8 cms. Note, the study site is denoted at the bottom of the page. (Right) Plot

displaying the relationship between inundation area (Ai) and streamflow (Q).

Because the StREAM Lab has a relatively short flow record (4 yrs), it was necessary to develop a

synthetic flow record to better characterize potential removal in the floodplain. This was done by

developing a flow duration curve (FDC) of mean daily flow (QMDF) from regional statistics,

Page 76: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

67

developing a 1000 yrs of synthetic flow record, and finally identifying overbank events during that

record. Using methods outlined by Ssegane et al., 2013 71, regional regression equations were

developed to estimate the FDC experienced at the StREAM Lab gage. To parameterize the

regression estimates, the USGS stream gaging network72 and associated watershed attributes from

the NHDplus database73 were utilized. More specifically, gages within the Ridge and Valley

Physiographic Region74 with watersheds less than 140 km2 were chosen for the analysis (n=227).

After optimization, regression variables were watershed area (WS), percent pasture landcover as

defined by NLCD 200132 (pasture), and mean annual rainfall (precip):

31 2

0[ ] [ ] [ ]MDFQ WS pasture precip Equation S4-2

Table S4-1. Regional Regression Coefficient Values

% Exceedence β0 β1 β2 β3 R2 P

99 -10.42 1.00 -0.05 1.82 0.69 <0.001

95 -7.35 0.96 -0.04 1.28 0.67 <0.001

90 -5.85 0.94 -0.03 1.02 0.66 <0.001

80 -4.58 0.91 -0.03 0.78 0.64 <0.001

70 -4.42 0.88 -0.02 0.72 0.61 <0.001

60 -4.92 0.56 -0.02 0.76 0.59 <0.001

50 -4.90 0.83 -0.02 0.73 0.54 <0.001

40 -4.94 0.80 -0.01 0.71 0.50 <0.001

30 -5.6 0.76 -0.01 0.78 0.46 <0.001

20 -6.26 0.73 -0.01 0.84 0.42 <0.001

10 -7.51 0.69 -6.2e-4 0.98 0.36 <0.001

5 -9.81 0.68 0.01 1.27 0.32 <0.001

1 -13.58 0.65 0.01 1.75 0.25 <0.001

where β0 , β1, β2, and β3 are regression coefficients defined for exceedence intervals 0.99, 0.95,

0.90, 0.80, … , 0.10, 0.05, 0.01 (Table S4-1). Then, the regression equations were applied to the

StREAM Lab gage’s watershed and a Log Pearson type III (LPIII) distribution was fit to the results

using a maximum-likelihood fitting technique75 (Table S4-2, Figure S4-2).

Table S4-2. Log Pearson Type 3 Parameters

Shape (α) 1.51

Rate (β) 7.65

Page 77: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

68

Figure S4-2. Parameterized Flow Duration Curve (FDC) at the StREAM Lab watershed. Grey

points result from regional regression, while the dashed line is the Log Pearson Type III

distribution fit to points at specific exceedence levels.

Then, the 1000 yr flow record was synthesized by resampling the parametrized LPIII distribution

using a random sampling technique developed by Ahrens and Dieter 1982.76 Here, it is important

to note that the synthetic record consists of QMDF, which does accurately capture overbank storm

events in low-order streams. 77-79

Similar to methods developed by Tan et al., 2007 82, the continuous flow record was discretized

into discrete subdaily hydrographs. Here, it is assumed the volume of water represented by

integrating QMDF across 24 hours equals the volume of stormflow and baseflow across that same

day:

MDF BF SFQ dt Q dt Q dt Equation S4-3

where QBF and QSF represent baseflow and stormflow, respectively. QBF was defined as 0.2 cms,

which is consistent with baseflow observed during the summer. QSF was estimated using a

rectangular hydrograph assumption and known relationship between instantaneous peak flow

(QIPF) and QMDF77, 80, 81:

IPF s MDFQ C Q Equation S4-4

( ) 1MDF BF

storm

s MDF

Q Q dayd

C Q

Equation S4-5

where CS is the steepness index defined by the relationship between QMDF and QIPF at the StREAM

Lab gage.

Bankful flow (QBKF), was used to identify individual overbank events over the course of the

synthetic flow record. Then, the relative NO3- removal was calculated as the ratio of Mremoval and

the total NO3- load and tabulated for individual storm events:

Page 78: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

69

Re% movalStorm

in

MRemoval

QC

Equation S4-6

Then, relative annual NO3- load reductions were tabulated across the synthetic flow record by

calculating the ratio of Mremoval, across the storm events in the respective year, to total annual load

experienced across the respective year:

Re

%

moval

year

Annual

in

M

RemovalQC

Equation S4-7

S4.4 Works Cited 1. Irmak, S., NEBFLUX. ASABE 2010, (53), 1098-1115.

2. Campbell Scientific Inc., C. S. Bowen Ratio Instrumentation Manual. Logan, UT, 1998.

3. Jones, C. N.; Scott, D. T.; Edwards, B. L.; Keim, R. F. Perirheic mixing and biogeochemical

processing in flow-through and backwater floodplain wetlands. Water Resour. Res. 2014, 50 (9),

7394-7405.

4. Hughes, D. A. Floodplain inundation - processes and relationships with channel discharge.

Earth Surf. Process. Landf. 1980, 5 (3), 297-304.

5. Knight, D. River flood hydraulics: Theoretical issues and stage-discharge relationships. River

basin modelling for flood risk mitigation 2006, 17, 301-334.

6. Jones, K. L.; Poole, G. C.; O'Daniel, S. J.; Mertes, L. A. K.; Stanford, J. A. Surface hydrology

of low-relief landscapes: Assessing surface water flow impedance using LIDAR-derived digital

elevation models. Remote Sens Environ 2008, 112 (11), 4148-4158.

7. Ssegane, H.; Amatya, D. M.; Tollner, E. W.; Dai, Z.; Nettles, J. E. Estimation of Daily

Streamflow of Southeastern Coastal Plain Watersheds by Combining Estimated Magnitude and

Sequence. J. Am. Water Resour. As. 2013, 49 (5), 1150-1166.

8. USGS National Water Information System. http://nwis.usgs.gov (accessed 2014).

9. Horizon Systems. National Hydrography Dataset Plus: Documentation. 2007

10. Fenneman, N. M.; Johnson, D. W. Physiographic divisions of the conterminous. In USGS, Ed.

Reston, VA, 1946.

11. Jin, S.; Yang, L.; Danielson, P.; Homer, C.; Fry, J.; Xian, G. A comprehensive change detection

method for updating the National Land Cover Database to circa 2011. Remote Sens Environ 2013

(132), 159-175.

12. Venables, W. N.; Ripley, B. D. Modern Applied Statistics with S. Fourth edition. Springer:

2002.

Page 79: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

70

13. Ahrens, J. H.; Dieter, U. Generating gamma variates by a modified rejection technique.

Communications of the ACM 1982, 25, 47-54.

14. Langbein, W. Peak discharge from daily records. Water Resour. Bull 1944, 145.

15. Fill, H. D.; Steiner, A. A. Estimating instantaneous peak flow from mean daily flow data. J

Hydrol Eng 2003, 8 (6), 365-369.

16. Tan, K.-S.; Chiew, F. H. S.; Grayson, R. B. A steepness index unit volume flood hydrograph

approach for sub-daily flow disaggregation. Hydrol. Process. 2007, 21 (20), 2807-2816.

17. Fuller, W. E. Flood flows. Trans. Am. Soc. Civ. Eng. 1914, 77, 564-617.

Page 80: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

71

CHAPTER 5. THE EFFECT OF HYDROGEOMORPHIC

GRADIENTS AND HYDROCLIMATIC SETTING ON RIVER-

FLOODPLAIN CONNECTIVITY AT THE CONTINENTAL

SCALE

C. Nathan Jones, Durelle T. Scott, Jesus D. Gomez-Velez, Judson W. Harvey

Submitted: anticipated, August 2015

To: Geophysical Research Letters

Status: Under internal review

Abstract Hydrologic connectivity between rivers and their adjacent floodplains control many ecological

processes that are consequential at catchment to continental scales; however little work has been

done to characterize river-floodplain connectivity beyond site specific studies. We developed a

method to examine the effects of contributing catchment size and physiography on river-floodplain

connectivity across the continental United States. Utilizing flow data from over 6,800 USGS

gages, we estimated transient storage volume in floodplains for individual floods across each

gage’s flow record. Floodplain transient storage varied from 5 to 50% of the total flood volume

for individual events, and there were no significant patterns between this proportion and

physiography or catchment area, highlighting the variability of flooding experienced across the

US. However, floodplain transient storage increased with stream order when generalized to the US

river network, both when examining individual reaches and cumulative storage across each stream

order. Further, floodplain transient storage was greatest in the southeastern coastal plain.

Collectively, our results highlight the importance of river-floodplain connectivity in large, lowland

rivers.

5.1 Introduction Feedbacks between regional to continental scale hydrologic and biogeochemical processes control

the timing and magnitude of water and material flux from large catchments. Because of the

complexity and heterogeneity of hydrologic processes at this scale, observations must be paired

with modeling efforts to better understand the effects of perturbations such as urbanization and

climate change. Recent efforts include estimations of nutrient flux from large river basins

[Alexander et al., 2000; Woodward et al., 2012], global and continental fluxes of riverine CO2

[Butman and Raymond, 2011; Raymond et al., 2013], and most recently, basin wide estimates of

the effects of hyporheic connectivity on solute retention and transport [Gomez-Velez and Harvey,

2014; Kiel and Cardenas, 2014]. While these contemporary modeling efforts highlight many

important processes, they typically focus on baseflow or median streamflow conditions. This

neglects discrete flood pulses, where hydrologic connection between the landscape and river

network is enhanced, and as a result, downstream flux of solutes often increases. This is

particularly relevant in catchments dominated by urban and agricultural landuses, where the

majority of the annual solute flux is accounted for during periods of elevated flow [Kaushal et al.,

2014; Royer et al., 2006].

Page 81: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

72

Flood pulses drive many ecosystem processes, ranging from controlling ecosystem composition

and structure to the exchange of energy and materials between rivers and their adjacent landscape

[Junk et al., 1989; Stanley et al., 1997; Tockner et al., 2000]. The latter is of interest for

downstream water quality, because floodplains serve as both a source and sink of materials within

a catchment [Noe and Hupp, 2007; Roley et al., 2012; Scott et al., 2014]. During overbank events,

large amounts of nutrient rich sediments and particulate organic matter are deposited on floodplain

surfaces [Curran and Hession, 2013; Hupp, 2000]. In systems that experience high rates of

deposition, these materials are often buried and effectively removed from the river network [Battin

et al., 2009; Ricker and Lockaby, 2015]. In other systems, these depositional materials provide

resources for enhanced biological processing and internal biogeochemical cycling [Welti et al.,

2012; Wolf et al., 2013], potentially increasing the biogeochemical processing of floodwaters and

improvement of downstream water quality [Mitsch et al., 2001; Tockner and Schiemer, 1997].

However, the extent of biogeochemical processing in floodplains is largely controlled by

feedbacks between redox dependent processes (i.e., denitrification) and inundation hydrology

[Helton et al., 2012; Jones et al., 2014]. For example, headwater floodplains are often a source of

nutrients because of the short residence time of floodwaters within the floodplain [Kreiling et al.,

2013; Noe and Hupp, 2007; Roley et al., 2012], while large river-floodplain systems are often a

sink because of prolonged residence times and its effect on redox conditions [Forshay and Stanley,

2005; Racchetti et al., 2011; Scott et al., 2014].

While there are many site specific investigations of feedbacks between inundation hydrology and

biogeochemistry in floodplains (e.g. Kreiling et al. [2013]; Noe and Hupp [2007]; Scott et al.

[2014]), few studies have attempted to quantify the cumulative effects of floodplains on material

expert from large river basins. Mertes [1997] hypothesized that interactions between hydroclimatic

setting and catchment area control inundation hydrology, suggesting that as rivers transitioned

from headwater streams to large rivers, they also shifted from a dominance of mixing in the

hyporheic zone to mixing in the perirheic zone. This hypothesis is partially confirmed by recent

modeling efforts, where hyporheic exchange decreased with increasing stream size and was shown

to be significant source of biogeochemical processing potential [Gomez-Velez and Harvey, 2014].

However, little work has been done to examine variation in river-floodplain interactions across

hydrogeomorphic gradients or hydroclimatic settings; and because river-floodplain interactions

drive biogeochemical processing of floodwaters (e.g., Junk et al., [1989]; Tockner et al. [2000]),

characterizing these large scale controls is an important first step in understanding the effects river-

floodplain connectivity on downstream water quality.

In this letter, we present a methodology to examine river-floodplain connectivity across large

catchments. We apply this methodology to the continental US and quantify the effects of

hydroclimatic setting and hydrogeomorphic gradients on river-floodplain connectivity. Broadly,

river-floodplain connectivity is defined as the exchange of energy and materials between the river

and floodplain [Pringle, 2003; K. Tockner and Stanford, 2002]. Here, we characterize river-

floodplain connectivity by estimating floodplain transient storage volume (VFP), or the volume of

transient storage experienced within a given reach’s floodplain, through (1) the analysis of 20 years

of flow data from over 6,800 USGS gaging station and (2) the application of observed variability

to the entire US stream network.

Page 82: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

73

5.2 Methods

Figure 5-1. A.) Rating curve from the Atchafalaya River near Simmesport, LA (USGS Gage

07381515) and a conceptual sketch of a river cross section highlighting the assumption that the

rating curve’s breakpoint represents the threshold for river-floodplain connectivity, or QBKF*. B.)

USGS gages utilized in this study from the GAGESII database across the different physiographic

regions. C.) An annual hydrograph highlighting the total flood volume (VT) as delineated by the

estimate of QBKF*. D.) Plot displaying the relationship between QBKF* and catchment area across

the gaging network. Colors correspond to physiographic regions displayed in B.

2.1 Characterization of Floodplain Transient Storage

Using flow data from over 6,800 USGS gaging stations [USGS, 2014] and the NHDplus database

[Horizon Systems, 2006], we characterized transient storage experienced in the floodplain for

individual floods across each gage’s flow record and then generalize the associated

characterization (i.e., VFP) to the US stream network. In this letter, we provide a brief accounting

of the methods used to accomplish this; however, a more thorough description of the methods and

associated assumptions can be found in the supporting information. Our analysis of the flow data

hinges on (i) identification of individual overbank floods, (ii) application of modified divided

channel method (DCM) to separate flood flow into channel and floodplain components, and (iii)

estimation of the transient volume of those components: channel volume (VC) and VFP. Then, using

information derived from the flow data analyses, we estimate annual floodplain transient storage

volume (S) across the US stream network based on both stream size and physiographic regions.

Page 83: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

74

For each gage, we established the threshold for river-floodplain connectivity using a piecewise

regression analysis of paired stage-discharge measurements (Figure 5-1a). The breakpoint between

regression segments often has physical meaning [Bacon and Watts, 1971], and in this application,

we assume it represents the point of incipient flooding, or the minimum flow where river-

floodplain connectivity occurs (QBKF*). This is consistent with hydrogeomorphic theory, which

suggests that there is an abrupt change in cross sectional geometry when stream flow transitions

from channel flow to overbank flooding [Hupp, 2000; Williams, 1978; Wolman, 1955].

Uncertainty associated with this estimate of QBKF* is largely associated with the hysteretic nature

of flood flows and river stage [Domeneghetti et al., 2012; Hughes, 1980; Knight, 2006], flow

measurement error [Kennedy, 1984; Rantz, 1982], and non-stationarity in stream/floodplain

bathymetry [Slater et al., 2015]. While this level of uncertainty may be unsuitable for site level

studies or engineering applications, we believe our estimation of QBKF* provides an acceptable

level of certainty to characterize river-floodplain connectivity at the continental scale. This is

illustrated by the emergence of power-law scaling between estimated QBKF* and catchment area

across gages utilized in this study (Figure 5-1d). The existence of this relationship does not provide

validation of the modeled QBKF* values, but it is consistent with other observations of regional and

continental scale power-law relationships between measured bankfull flow and catchment area

(e.g. Bieger et al., [2015]). See the supplemental material for more information.

We utilized a modification of the DCM (e.g. Chow [1959]) to estimate VFP for individual flood

events across each gage’s record from 1986 to 2006. Because USGS flow data is typically stored

as mean daily flow (QMDF), subdaily hydrographs were discretized using methods similar to Sangal

[1983] and Tan et al. [2007]. This allowed for the inclusion of floods with durations less than 24

hours, which is critical when characterizing floodplain transient storage in headwater streams.

Then, we identified individual overbank events using QBKF* as a threshold across each gages flow

record, respectively, and estimated the total flood volume (VT) for each flood by integrating under

the flood hydrograph with respect to time ( , Figure 5-1c). Then, we used the DCM

to divide flood flow into channel and floodplain components, where VT is assumed to be the sum

of VFP and VC. Similar to VT, we approximated VC by integrating estimated channel velocity (UC)

with respect to the cross sectional area of the channel control volume (AC) and flood duration (

). UC is assumed to be equivalent to bankful velocity and is estimated using

relationships developed by Bjerklie [2007]. AC was estimated using downstream hydraulic

geometry relationships developed by Wilkerson et al. [2014] and application of the DCM active

channel divisions. Finally, we calculated VFP by subtracting VC from VT. This process was

conducted for individual floods identified across each gage’s flow record from 1986 to 2006. See

the supplemental material for more additional information.

To better describe the observed variation in the gaging data, we developed 2 metrics: the flood

volume ratio (FVR) and the storage index (SI). The FVR is the ratio of the summation of VT and

the total volume of streamflow 2006

1986Qdt , respectively, experienced across each gage’s record

from 1986 to 2006 (Equation 5-1). This is a cumulative measure of flooding and generally

describes the upstream hydrology and the hydroclimatic setting of each gage. In contrast, SI is a

ratio of VFP and VT for individual floods and provides a discrete measure of river-floodplain

connectivity that provides an approximation of the proportion of flood flow experienced in the

T FloodV Q dt

C C CV U dA dt

Page 84: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

75

floodplain (Equation 5-2). For simplicity, the median SI from each gage was utilized in subsequent

calculations. To better understand patterns associated with stream order and physiography, we

found median values of FVR and SI for each physiographic province across the observed stream

orders. Then, we used Spearmen Correlation (e.g. Spearman [1904]) to test if there was a

significant effect of stream order on median FVR and median SI, respectively, for each

physiographic division.

2006

1986

TVFVR

Qdt

(5-1)

(5-2)

2.2 Application to the US Stream Network

In order to gain perspective on continental scale patterns in floodplain transient storage, we applied

the metrics obtained from gaging station analysis to the US stream network. While the USGS

gaging station coverage is extensive, individual gages only serve as point measurements, and

analysis of river-floodplain connectivity across the entire network is needed to fully understand

spatial variations associated with both physiography and catchment area. Thus, gages were

subdivided into their respective physiographic domains (R) and stream order (ω), as defined by

Fenneman and Johnson [1946] and Strahler [1957]; and, annual floodplain transient storage

volume (SR,ω) was approximated for each subdivision as the product of median observed floodplain

transient storage volume ( ) and median number of events annually ( )Fn .

,R FP FS V n (5-3)

In order to generalize the floodplain transient storage metrics to the US stream network, values of

SR,ω were applied to individual stream reaches with corresponding R and ω values. The NHDplus

stream network and database was used to define network geometry and ω for individual stream

reaches. To summarize the effects of catchment area on river-floodplain connectivity, we

calculated cumulative floodplain transient storage volume by stream order ( ,RS S

) and

length (lreach) weighted mean floodplain transient storage volume for each stream order (

, /R reach reachS S l l

). Similarly, to summarize the patterns associated with physiography

and river-floodplain connectivity, we calculated the length normalized cumulative floodplain

transient storage volume across each physiographic province ( , /R R reach

R

S S l ).

FP

T

VSI

V

FPV

Page 85: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

76

5.3 Results and Discussion 5.3.1 Heterogeneity of River-Floodplain Connectivity across USGS Gaging Stations

Figure 5-2. A.) The flood volume ratio (FVR) and B.) median storage index (SI) approximated

from USGS flow records. Lines represent median index values experienced at specific steam

orders across the different physiographic regions, while the 10-90% quantile interval from across

the entire dataset are displayed in grey.

There was considerable variability of river-floodplain connectivity experienced across the USGS

gaging stations, as measured by the both the SI and FVR (Figure 5-2). These results are consistent

with recent efforts to understand patterns and mechanisms associated with flood peak distributions

across the US (e.g. Mallakpour and Villarini [2015]; Slater et al. [2015]; Villarini et al. [2009]),

where variability in flooding is typically attributed to heterogeneity of upstream hydrology and the

temporal variability of hydroclimatic conditions. Because of economic and social consequences of

flooding, the majority of these analyses focus on the distribution of flood peaks and “mixtures” of

flood generating mechanisms (e.g. Hirschboeck [1988]; Murphy [2001]; Villarini and Smith

[2010]). Our analysis represents a significant departure from these studies, in that we examined

flood volumes as opposed to peak flow, which has consequences for solute transport and ecological

function. Here, we believe further investigation is needed to fully realize the benefits of

investigations of distributions of integrative flood volume.

When examining the FVR in relation to stream order, two groups of physiographic divisions

became apparent: rainfall and snowmelt dominated regions as displayed in blue and orange,

Page 86: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

77

respectively, in Figures 1b and 2. These opposing hydroclimatic regimes are large scale controls

that influence feedbacks between hydrogeomorphic and ecological processes [Montgomery,

1999]. The Appalachian Highlands, Atlantic Plain, and Interior Highlands physiographic divisions

all had increasing and statistically significant correlations between median FVR and stream order.

Similarly, in an analysis of flood generating mechanisms at USGS gages across the Eastern US,

Villarini and Smith [2010] found that flood peak distribution parameters were strongly tied to basin

area and flood generating mechanisms were typically associated with atmospheric flooding (e.g.

organized convective thunderstorms, tropical cyclones, etc.). In contrast, the Intermontane

Plateaus, Laurentian Upland, Pacific Mountains, and Rocky Mountains all experienced decreasing

or statistically insignificant trends between FVR and stream order. Here, we attribute this trend to

the overwhelming proportion of the annual water budget experienced during seasonal snowmelt

runoff events, or in the case of drylands of the Intermontane Plateaus, seasonal monsoonal events.

Interestingly, the Interior Plains physiographic domain experiences a positive, yet statistically

insignificant trend. We attribute this to the gradient from snowmelt and rain on snow/ice to summer

convective thunderstorm flood generating mechanisms experienced from north to south in the

physiographic domain [Villarini et al., 2011]. Differences in these two hydroclimatic regimes

ultimately control variability in river-floodplain connectivity experienced across the two groups

of physiographic divisions.

Our examination of median SI highlights the temporal and spatial variability of transient storage

volume (e.g. VFP) experienced across different floodplain environments. Because of computational

and measurement constraints, few studies estimate the water budget of floodplains during

overbank events (e.g. Alsdorf et al. [2005]; Scott et al. [2014]); yet, this is a critical variable in

understanding potential biogeochemical processing within floodplain environments from site to

catchment scales. Our estimation of SI provides a coarse approximation of river-floodplain

exchange, with median values ranging from 5 to 50% across physiographic regions. This is

consistent with flume studies of two stage channels (e.g., Ackers [1993]), where VFP ranged from

approximately 10-50% of VT and was a function of mean water levels in both channel and

floodplain control volumes, respectively. While SI estimates provide an important first step in

understanding the role of transient storage in floodplain across the river network, SI does not

explain mixing processes that regulate residence time of floodwater and ultimately control

depositional and biogeochemical processes [Hughes, 1980; Jones et al., 2014].

Page 87: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

78

5.3.2 Effect of Catchment Area on River-Floodplain Connectivity

Figure 5-3. A.) Map of estimated floodplain transient storage volume in 1st to 8th order streams

across the US. Note, rivers greater than 8th order (e.g. Lower Mississippi River) were not included

in this analysis and are displayed in grey. B.) Cumulative floodplain transient storage volume

across the stream network. C.) Length weighted mean floodplain transient storage volume across

the stream network. Diamonds are based on estimates from median values of floodplain transient

storage volume assigned to the stream network, while bars represent the 10th and 90th quantiles of

floodplain transient storage volume.

Page 88: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

79

When applied to the US stream network, our results show that river-floodplain connectivity

generally increases with stream order (Figure 5-3). This is consistent with hydrogeomorphic

theory, which suggests that as river networks transition from headwater streams to large rivers,

flooding transitions from short episodic pulses to longer, sustained floods [Junk et al., 1989]. This

pattern in flooding leads to annual inundation times ranging from hours in many headwater streams

(e.g. Harrison et al. [2014]) to months in large river systems (e.g. Scott et al. [2014]). Therefore,

it is intuitive and unsurprising that transient storage in floodplains would increase with increasing

stream order (Figure 5-3c). When examining cumulative SR,ω across the river network

(Figure 5-3b), floodplain transient storage volume generally increases with stream order. This

highlights the importance of river-floodplain connectivity in large river systems.

In many ways, patterns of river-floodplain connectivity experienced along the river corridor are

the inverse of patterns associated with surface water-groundwater (SW-GW) connectivity [Mertes,

1997]. The hyporheic zone is often defined as the subsurface mixing zone where water and

materials are exchanged between surface water and groundwater [Gooseff, 2010; White, 1993;

Winter et al., 1998]. Similarly, the perirheic zone is defined as the surface mixing zone where

water and materials are exchanged between the channel and floodplain [Jones et al., 2014; Mertes,

1997]. While the number and heterogeneity of hyporheic flowpaths generally increases with

increasing stream size [Stanford and Ward, 1993], hyporheic flowpaths generally convey a smaller

portion of total flow with increasing stream size [Wondzell, 2011]. Because of the dendritic nature

of stream networks, and the overwhelming number of small order streams when compared to

higher order streams, this gradient in SW-GW connectivity highlights the importance of hyporheic

flow in headwater streams at the catchment scale [Gomez-Velez and Harvey, 2014; Kiel and

Cardenas, 2014]. Paired with our results, these studies cumulatively suggest that as the river

corridor shifts from headwater streams to large rivers, mixing processes shift from a dominance of

hyporheic flowpaths and SW-GW connectivity to perirheic flowpaths and river-floodplain

connectivity.

Feedbacks between residence times of floodwaters and river-floodplain exchange largely control

biogeochemical processing within floodplains [Jones et al., 2014]. Here, we hypothesize that

similar to the observed increase in transient storage volume with stream order, there is also an

increase in floodwater residence time with increasing stream size. In headwater floodplains with

relatively short residence times (e.g. hours), the dominant biogeochemical mechanism is flushing

and the floodplain is a source of reactive solutes [Harrison et al., 2014; Kreiling et al., 2013; Noe

and Hupp, 2007; Roley et al., 2012]. In contrast, biogeochemical cycling in medium to large rivers

is often suppressed because of prolonged residence times and transport limitation [Forshay and

Stanley, 2005; Racchetti et al., 2011]. Therefore, further investigation is needed to understand

biogeochemical processing potential of floodplains across large catchments.

Page 89: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

80

5.3.3 Effect of Physiography on River-Floodplain Connectivity

Figure 5-4. The summation of length normalized floodplain transient storage volume for each

physiographic province.

Finally, we see distinct patterns of river-floodplain connectivity across the landscape (Figure 5-4).

When observing the length normalized cumulative floodplain transient storage volume across each

physiographic province, we find that floodplain transient storage volume is most prevalent in the

Coastal Plain and Central Lowlands. In the coastal plain, this is largely driven by a combination

of expansive floodplains formed during the pluvial period 18,000 to 10,000 years ago [Hupp, 2000]

and regular occurrence of tropical cyclones [Villarini and Smith, 2010]. This is in contrast to high

gradient streams with confined valleys typically found in Rocky Mountains and the relatively dry

climate experienced across the southwestern US. Similar to patterns seen in the gaging station time

series, our results confirm that the contemporary hydroclimatic setting is a dominant variable

affecting river-floodplain connectivity.

5.4 Conclusion We developed a methodology to characterize river-floodplain connectivity across large river

basins. We applied this methodology to the US River network and characterized the effects

contributing catchment area and physiography on floodplain transient storage volume. The

analysis consisted of quantifying floodplain transient storage volume at over 6,800 gages across

the US; then, we generalized information about river-floodplain connectivity to the greater river

network. Largely, our results suggest:

1. There is considerable spatial heterogeneity and temporal variability associated with

floodplain transient storage volume experienced across the USGS gaging stations.

Page 90: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

81

However, floodplain transient storage volume accounted for 5 to 50% of the total flood

volume for most flood events.

2. Floodplain transient storage volume increased along the river corridor. Our results suggest

that annual floodplain transient storage volume increased with stream order for both

individual stream reaches and cumulatively across the river network.

3. The Coastal Plain, Great Plains, and Central Lowlands are hotspots for river-floodplain

connectivity. This is due to their hydroclimatic setting and geomorphic form.

Our analysis represents an important first step in quantifying the cumulative effects of floodplains

on downstream hydrology and water quality. We characterized river-floodplain connectivity by

estimating the volume of transient storage in the floodplain. However, in order to understand

reactive transport processes, it will be critical to quantify residence times of floodwaters within

floodplains and also the exchange rates between the river and floodplain (e.g. Harvey et al. [2013]).

Our results suggest that river-floodplain connectivity is an important variable in the fate and

transport of materials, especially in medium to large river systems, and should be included in

contemporary modeling effort to understand regional to continental scale processes.

5.5 Works Cited Ackers, P. (1993), Stage-discharge functions for two-stage channels: The impact of new

research, J. Inst. Water. Env. Man., 7(1), 52-61.

Alexander, R. B., R. A. Smith, and G. E. Schwarz (2000), Effect of stream channel size on the

delivery of nitrogen to the Gulf of Mexico, Nature, 403(6771), 758-761.

Alsdorf, D., T. Dunne, J. Melack, L. Smith, and L. Hess (2005), Diffusion modeling of

recessional flow on central Amazonian floodplains, Geophys. Res. Lett., 32(21).

Bacon, D. W., and D. G. Watts (1971), Estimating the Transition between Two Intersecting

Straight Lines, Biometrika, 58(3), 525-534.

Battin, T. J., S. Luyssaert, L. A. Kaplan, A. K. Aufdenkampe, A. Richter, and L. J. Tranvik

(2009), The boundless carbon cycle, Nat Geosci, 2(9), 598-600.

Bjerklie, D. M. (2007), Estimating the bankfull velocity and discharge for rivers using remotely

sensed river morphology information, J. Hydrol., 341(3–4), 144-155.

Butman, D., and P. A. Raymond (2011), Significant efflux of carbon dioxide from streams and

rivers in the United States, Nat Geosci, 4(12), 839-842.

Chow, V. T. (1959), Open-channel hydraulics, McGraw-Hill.

Curran, J. C., and W. C. Hession (2013), Vegetative impacts on hydraulics and sediment

processes across the fluvial system, J. Hydrol., 505(0), 364-376.

Fenneman, N. M., and D. W. Johnson (1946), Physiographic divisions of the conterminous,

edited by USGS, Reston, VA.

Forshay, K. J., and E. H. Stanley (2005), Rapid nitrate loss and denitrification in a temperate

river floodplain, Biogeochemistry, 75(1), 43-64.

Page 91: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

82

Gomez-Velez, J. D., and J. W. Harvey (2014), A hydrogeomorphic river network model predicts

where and why hyporheic exchange is important in large basins, Geophys. Res. Lett., 41(18),

6403-6412.

Gooseff, M. N. (2010), Defining hyporheic zones–advancing our conceptual and operational

definitions of where stream water and groundwater meet, Geography Compass, 4(8), 945-955.

Harrison, M. D., A. J. Miller, P. M. Groffman, P. M. Mayer, and S. S. Kaushal (2014),

Hydrologic Controls on Nitrogen and Phosphorous Dynamics in Relict Oxbow Wetlands

Adjacent to an Urban Restored Stream, J. Am. Water Resour. As., 50(6), 1365-1382.

Harvey, J. W., J. Böhlke, M. A. Voytek, D. Scott, and C. R. Tobias (2013), Hyporheic zone

denitrification: Controls on effective reaction depth and contribution to whole‐ stream mass

balance, Water Resour. Res., 49(10), 6298-6316.

Helton, A. M., G. C. Poole, R. A. Payn, C. Izurieta, and J. A. Stanford (2012), Scaling flow path

processes to fluvial landscapes: An integrated field and model assessment of temperature and

dissolved oxygen dynamics in a river-floodplain-aquifer system, J. Geophys. Res.-Biogeosci.,

117, 13.

Hirschboeck, K. K. (1988), Flood Hydroclimatology, in Flood Geomorphology, p. 27-49, John

Wiley & Sons, New York.

Horizon-Systems, (2006), National Hydrography Dataset Plus, Retreived from

http://www.horizon-systems.com/NHDPlus/index.php

Hughes, D. A. (1980), Floodplain inundation - Processes and relationships with channel

discharge, Earth Surf. Process. Landf., 5(3), 297-304.

Hupp, C. R. (2000), Hydrology, geomorphology and vegetation of Coastal Plain rivers in the

south-eastern USA, Hydrological Processes, 14(16-17), 2991-3010.

Jones, C. N., D. T. Scott, B. L. Edwards, and R. F. Keim (2014), Perirheic mixing and

biogeochemical processing in flow-through and backwater floodplain wetlands, Water. Resour.

Res., 50(9), 7394-7405.

Junk, W. J., P. B. Balyley, and R. E. Sparks (1989), The flood pulse concept in river-floodplain

systems., Can. Spec. Publ. Fish. Aquat. Sci., 106, 110-127.

Kaushal, S. S., P. M. Mayer, and P. G. Vidon (2014), Land use and climate variability amplify

carbon, nutrient, and contaminant pulses: a review with management implications, J. Am. Water

Resour. As., 50(3), 585-614.

Kiel, B. A., and M. B. Cardenas (2014), Lateral hyporheic exchange throughout the Mississippi

River network, Nat Geosci, 7(6), 413-417.

Kreiling, R. M., J. P. Schubauer-Berigan, W. B. Richardson, L. A. Bartsch, P. E. Hughes, J. C.

Cavanaugh, and E. A. Strauss (2013), Wetland Management Reduces Sediment and Nutrient

Loading to the Upper Mississippi River, J. Environ. Qual., 42(2), 573-583.

Mallakpour, I., and G. Villarini (2015), The changing nature of flooding across the central

United States, Nature Clim. Change, 5(3), 250-254.

Mertes, L. A. K. (1997), Documentation and significance of the perirheic zone on inundated

floodplains, Water Resour. Res., 33(7), 1749-1762.

Page 92: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

83

Mitsch, W. J., J. W. Day, J. W. Gilliam, P. M. Groffman, D. L. Hey, G. W. Randall, and N. M.

Wang (2001), Reducing nitrogen loading to the Gulf of Mexico from the Mississippi River

Basin: Strategies to counter a persistent ecological problem, Bioscience, 51(5), 373-388.

Montgomery, D. R. (1999), Process domains and the river continuum, J. Am. Water Resour.

Assoc., 35(2), 397-410.

Murphy, P. J. (2001), Evaluation of mixed-population flood-frequency analysis, J. Hydrol. Eng.,

6(1), 62-70.

Noe, G. B., and C. R. Hupp (2007), Seasonal variation in nutrient retention during inundation of

a short-hydroperiod floodplain, River Research and Applications, 23(10), 1088-1101.

Pringle, C. (2003), What is hydrologic connectivity and why is it ecologically important?, Hydro.

Process., 17(13), 2685-2689.

Racchetti, E., M. Bartoli, E. Soana, D. Longhi, R. R. Christian, M. Pinardi, and P. Viaroli (2011),

Influence of hydrological connectivity of riverine wetlands on nitrogen removal via

denitrification, Biogeochemistry, 103(1-3), 335-354.

Raymond, P. A., et al. (2013), Global carbon dioxide emissions from inland waters, Nature,

503(7476), 355-359.

Ricker, M., and B. G. Lockaby (2015), Soil Organic Carbon Stocks in a Large Eutrophic

Floodplain Forest of the Southeastern Atlantic Coastal Plain, USA, Wetlands, 35(2), 291-301.

Roley, S. S., J. L. Tank, M. L. Stephen, L. T. Johnson, J. J. Beaulieu, and J. D. Witter (2012),

Floodplain restoration enhances denitrification and reach-scale nitrogen removal in an

agricultural stream, Ecol. Appl., 22(1), 281-297.

Royer, T. V., M. B. David, and L. E. Gentry (2006), Timing of Riverine Export of Nitrate and

Phosphorus from Agricultural Watersheds in Illinois: Implications for Reducing Nutrient

Loading to the Mississippi River, Environ. Sci. Technol., 40(13), 4126-4131.

Sangal, B. P. (1983), Practical Method of Estimating Peak Flow, J Hydraul Eng-Asce, 109(4),

549-563.

Scott, D. T., R. F. Keim, B. L. Edwards, C. N. Jones, and D. E. Kroes (2014), Floodplain

biogeochemical processing of floodwaters in the Atchafalaya River Basin during the Mississippi

River flood of 2011, J. Geophys. Res.-Biogeosci., 119(4), 537-546.

Slater, L. J., M. B. Singer, and J. W. Kirchner (2015), Hydrologic versus geomorphic drivers of

trends in flood hazard, Geophys. Res. Lett., 2014GL062482.

Spearman, C. (1904), The Proof and Measurement of Association between Two Things, Am. J.

Psychol., 15(1), 72-101.

Stanford, J. A., and J. V. Ward (1993), An Ecosystem Perspective of Alluvial Rivers:

Connectivity and the Hyporheic Corridor, J. N. Am. Benthol. Soc., 12(1), 48-60.

Stanley, E. H., S. G. Fisher, and N. B. Grimm (1997), Ecosystem expansion and contraction in

streams, Bioscience, 47(7), 427-435.

Strahler, A. N. (1957), Quantitative analysis of watershed geomorphology, Eos, Transactions

American Geophysical Union, 38(6), 913-920.

Page 93: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

84

Tan, K.-S., F. H. S. Chiew, and R. B. Grayson (2007), A steepness index unit volume flood

hydrograph approach for sub-daily flow disaggregation, Hydro. Process., 21(20), 2807-2816.

Tockner, K., and F. Schiemer (1997), Ecological Aspects of the Restoration Strategy for a River-

Floodplain System on the Danube River in Austria, Global Ecol. Biogeogr., 6(3/4), 321-329.

Tockner, K., and J. A. Stanford (2002), Riverine flood plains: present state and future trends,

Environ. Conserv., 29(3), 308-330.

Tockner, K., F. Malard, and J. V. Ward (2000), An extension of the flood pulse concept, Hydro.

Process., 14(16-17), 2861-2883.

United Station Geological Survey, (2014), National Water Information System, Retreived from

waterdata.usgs.gov/nwis.

Villarini, G., and J. A. Smith (2010), Flood peak distributions for the eastern United States,

Water Resour. Res., 46, W06504, doi:10.1029/2009WR008395.

Villarini, G., F. Serinaldi, J. A. Smith, and W. F. Krajewski (2009), On the stationarity of annual

flood peaks in the continental United States during the 20th century, Water Resour. Res., 45,

W08414, doi:10.1029/2008WR007645

Villarini, G., J. A. Smith, M. L. Baeck, and W. F. Krajewski (2011), Examining Flood

Frequency Distributions in the Midwest U.S., J. Am. Water Resour. Assoc., 47(3), 447-463.

Welti, N., E. Bondar-Kunze, G. Singer, M. Tritthart, S. Zechmeister-Boltenstern, T. Hein, and G.

Pinay (2012), Large-scale controls on potential respiration and denitrification in riverine

floodplains, Ecol Eng, 42, 73-84.

White, D. S. (1993), Perspectives on Defining and Delineating Hyporheic Zones, J. N. Am.

Benthol. Soc., 12(1), 61-69.

Wilkerson, G. V., D. R. Kandel, L. A. Perg, W. E. Dietrich, P. R. Wilcock, and M. R. Whiles

(2014), Continental-scale relationship between bankfull width and drainage area for single-

thread alluvial channels, Water Resour. Res., 50(2), 919-936.

Williams, G. P. (1978), Bank-full discharge of rivers, Water Resour. Res., 14(6), 1141-1154.

Winter, T. C., J. W. Harvey, O. L. Franke, and W. M. Alley (1998), Ground water and surface

water: A single resource, U.S. Geol. Surv. Circ., 1139, 79 pp.

Wolf, K. L., G. B. Noe, and C. Ahn (2013), Hydrologic Connectivity to Streams Increases

Nitrogen and Phosphorus Inputs and Cycling in Soils of Created and Natural Floodplain

Wetlands, J. Environ. Qual., 42(4), 1245-1255.

Wolman, M. G. (1955), The natural channel of Brandywine creek, Pennsylvania, United States

Geological Survey Professional Paper, 271, 56.

Wondzell, S. M. (2011), The role of the hyporheic zone across stream networks, Hydro.

Process., 25(22), 3525-3532.

Woodward, G., et al. (2012), Continental-Scale Effects of Nutrient Pollution on Stream

Ecosystem Functioning, Science, 336(6087), 1438-1440.

Page 94: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

85

CHAPTER S5. SUPPLEMENTAL MATERIAL

The overall goal of this study was to explore the effects of geomorphic gradients and hydroclimatic

setting on hydrologic connectivity between rivers and floodplains across the continental US.

Generally, hydrologic connectivity is defined as the water-mediated exchange of matter and energy

between the river and its adjacent landscape [Pringle, 2003; Tockner and Stanford, 2002].

Operationally, connectivity between rivers and their adjacent floodplain, or river-floodplain

connectivity, can be described by the duration of connection, the exchange and flux of water

between the floodplain and river, and the residence time of floodwaters within the floodplain

[Jones et al., 2014]. However, because of the heterogeneous nature of a typical floodplain surface

and its effects on inundation hydrology, the latter variables are difficult to measure at the site level

[Hughes, 1980], thus limiting regional to continental scale assessments of floodplain connectivity.

This includes recent efforts to understand biogeochemical cycles at the network scale (e.g.

Alexander et al., [2000]; Ensign and Doyle, [2006]; Gomez-Velez and Harvey, [2014]). Below, we

propose a simple methodology to analyze river-floodplain connectivity, as characterized by an

estimate of transient storage volume in the floodplain, at the continental scale and apply it across

the US river network.

Our methodology can be broken into two main components: (1) the analysis of 20 years of flow

data from over 6,800 USGS gaging station and (2) the application of the observed variability to

the entire US stream network. Our analysis of the flow data hinges on (i) identification of

individual overbank floods, (ii) application of modified divided channel method (DCM) to separate

flood flow into channel and floodplain components, and (iii) estimation of the volume of those

components: channel volume (VC) and VFP. Then, using information derived from the flow data

analyses, we estimate annual floodplain transient storage (S) across the US stream network based

on both stream size and physiographic regions. Table S5-1 displays definitions of pertinent

variables.

S5.1 Data Acquisition We utilized data from both the USGS stream gaging network [USGS, 2014] and the NHDplus

geospatial database [Horizon Systems, 2014] for this analysis. The USGS stream gaging network

consists of over 20,000 gaging stations that continuously collect a variety of parameters describing

streamflow, water quality, and various other environmental conditions. We chose the gaging

stations identified within GAGESII database (n=6785) for this analysis. These sites are located

along perennial streams or rivers, have a flow record of at least 20 years, and occur within a

watershed that can be reliably delineated [Falcone et al., 2010]. We obtained stage-discharge field

measurements, a corrected stage-discharge relationship, and a 20 year flow record (1986-20006)

for each gage from the National Water Information System (http://waterdata.usgs.gov/nwis).

The NHDplus geodatabase combines data from the National Hydrography Dataset (NHD), the

National Elevation Dataset (NED), and the National Watershed Boundary Dataset (WBD) [McKay

et al., 2012]. The base unit within the geodatabase is the flow line, which represents a relatively

short reach of stream or river. For each flow line, the database contains information about the

individual reach and contributing watershed such as landuse, climatic data, and annual runoff

estimates. For this analysis, we utilized channel slope (SCH), contributing catchment area (AWS),

and Strahler stream order (ω) from the NHDplus geodatabase.

Page 95: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

86

Table S5-1. Main variables used in analysis

Variable Definition Description Units

Q -- streamflow [L3T-1]

S -- stage [L]

QBKF* -- threshold for flooding [L3T-1]

QMDF -- mean daily flow [L3T-1]

QIPF , , 1 , 14

2

MDF n MDF n MDF nQ Q Q estimated instantaneous peak flow [L3T-1]

tF -- duration of a flood event [T]

nF -- number of overbank events in a given water year [-]

wBKF

0.191

0.462

2.18 , ln( ) 1.6

1.41 , ln 1.6

WS WS

WS WS

A A

A A

bankful width [L]

AWS -- contributing catchment area [L2]

UBKF 0.31 0.321.37 CHS bankful velocity [LT-1]

SCH -- channel slop [LL-1]

λ 1.1210.2 BKFw meander wavelength [L]

VT Qdt flood volume [L3]

VC *  BKFBKFQ dt U dAdt ∬ channel volume [L3]

VFP F CV V floodplain volume [L3]

FVR 2006

1986

TV

Qdt

flood volume ratio [LL-1]

SI FP

T

V

V storage index [LL-1]

ω -- Strahler stream order [-]

R -- physiographic division or province [-]

SR,ω FP fV n estimated annual floodplain transient storage

volume [L3]

Sω ,RS

cumulative floodplain transient storage volume

across a given stream order [L3]

S

,R reach

reach

S l

l

length weighted mean floodplain transient

storage volume across a given stream order [L3]

SR , /R reach

R

S l length normalized cumulative floodplain

transient storage volume for a given

physiographic province

[L3 L-1]

*Note, “~” and “-“ above letters denotes median and mean values, respectively.

Page 96: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

87

S5.2 Analysis of USGS Gage Data S5.2.1 Identifying the threshold for connectivity

In order to isolate individual storms across a gage’s flow record, we identified the threshold for

flooding through a piecewise regression analysis, or breakpoint analysis, of paired stage-discharge

measurements. The piecewise regression model consists of two or more general linear models with

each model bound to a specific range of the independent variable. A breakpoint is a point of

transition at the intersection between general linear models that may be interpreted to have physical

meaning, such as a shift in regime [Bacon and Watts, 1971]. For this application, we are utilizing

a two linear segment model, where the breakpoint is interpreted as characterizing bankfull

discharge (QBKF*) when incipient flooding is occurring. The piecewise regression model was

optimized utilizing an iterative procedure developed by Muggeo [2003] and implemented using R

statistical software [R Development Core Team, 2012].

Equation S5-1 displays the linear model chosen to represent stage-discharge measurement pairs.

Here, Q(S) is rating curve function where Q is streamflow and S is measured stage, F1 and F2

represent general linear models, I(S) is a binary function, and c is the breakpoint or threshold stage.

At each gage, we evaluated linear, exponential (log-transformed data), and power (log-log

transformed data) models, and we selected the model with the minimum Root Mean Square Error

(RMSE). The resulting model was considered adequate if: the slopes of the two segments were

significantly different (Davies Test, p<0.05), and if the estimated breakpoint was greater than the

minimum recorded S of the published USGS stage-discharge rating curve. Of the over 6,500 gages

analyzed, 5,981 produced adequate breakpoints.

1 2

1,( )

0,   

( ) ( ) ( ) ( ) [1 ( )],Q S F S I S F S I

CI

S

S

SS

C

Equation S5- 1

While the proposed technique provides an adequate method to estimate trends in river-floodplain

connectivity at the continental scale, there is considerable uncertainty associated in estimating

QBKF* utilizing stage-discharge measurements. Hydraulic geometry measures such as cross-

sectional area (AXS), S, and wetted perimeter (WP) exhibit a dramatic shift in relation to stream

flow (Q) as the flow regime shifts from in-channel flow to overbank or flood flows [Hupp, 2000;

Williams, 1978; Wolman, 1955]. Williams [1978] suggested that this shift was most evident in the

relationship between Q and AXS, but could be observed in the relationship between Q and S in areas

where floodplains persist. This is evident in methods utilized by the USGS to develop streamflow

rating curves of compound channels [Rantz, 1982], where breakpoints in the stage-discharge

relationship represent dramatic shifts in hydraulic geometry. The relationship between Q and S is

imperfect in natural channels [Chow, 1959]. Stage-discharge relationships often exhibit hysteretic

behavior over the course of individual storm-hydrographs because of hydraulic factors such as

backwatering, seasonal differences in vegetative roughness, and unsteady flow [Domeneghetti et

al., 2012; Hughes, 1980; Knight, 2006]. As a result, there is often subjectivity in the analysis of

stage-discharge rating curves that allows for experience and site-specific knowledge [Kennedy,

1984]. Other independent variables such as mean velocity and water surface slope are often utilized

as independent variables in rating curves to minimize the effect of hysteresis in the stage-discharge

relationships [Kennedy, 1984; Rantz, 1982]. Further complicating matters, extreme high flows are

Page 97: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

88

often estimated through an extrapolation from the indirect slope-area technique, which relies on

the parameterization of the Manning Equation [NRCS, 2012]. Because of their location on the

stage-discharge graph, these extreme events can exhibit substantial leverage in the pairwise

regression analysis, potentially imposing propagation error on the estimated breakpoint.

Figure S5-1. A.) Relationship between QBKF* and catchment area for 5,981 USGS gaging stations

across the continental US. B.) Relationship between bankful depth (dBKF* ) and catchment area.

The black lines represents the relationships derived from data in this study, while the orange line

represents a relationship developed by Beiber et al., [2015].

While these uncertainties in quantifying QBKF* might be unsuitable for site level hydrologic studies

or engineering applications, we believe our estimation of QBKF* provides an acceptable level of

certainty to characterize flooding at the continental scale. This is illustrated by the emergence of

power-law scaling between estimated QBKF* and catchment area across gages utilized in this study

(Figure S5-4a). The existence of this relationship does not provide validation of the modeled QBKF*

values, but it is consistent with other observations of regional and continental scale power-law

relationships between measured bankfull flow and catchment area. More specifically, we

compared the downstream hydraulic geometry relationship developed by Bieger et al., [2015] for

the continental US (n=1,310 sites) to results from our study. Here, bankful depth (dBKF*) was

estimated by subtracting minimum observed stage (Smin) from the stage corresponding to the

estimated breakpoint (SBKF*). Even with the added uncertainty of estimating dBKF*, the

Beiger et al.,[2015] model accounted for a significant amount of variation in the dBKF* values

(p<0.01), providing further evidence the breakpoint analysis adequately defined downstream

hydraulic geometry relationships.

S5.2.2 Estimation of Sub-daily hydrographs

Practical estimation of basin-scale flood characteristics across the historic flow record utilizes

records of mean daily flow (QMDF). Floods of short duration are common in low order streams and

ideally, sub-daily flow measurements would be utilized to characterize overbank floods in streams

where the majority of floods are often less than 24 hours. Several methods have been developed

A.) B.)

Page 98: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

89

to estimate instantaneous peak flow (QIPF) from QMDF records [Fill and Steiner, 2003]. The first

utilizes regional regression techniques to estimate the relationship between QMDF and QIPF for a

specific physiographic region or larger basin (e.g., Fuller [1914]; Taguas et al. [2008]). While

regional regression techniques have proven to be an effective estimators of QIPF, they do not

provide any information about the shape of the hydrograph and are not related to the prevailing

hydrologic conditions. In contrast, the second technique estimates QIPF utilizing consecutive QMDF

values (e.g., Langbein [1944]; Sangal [1983]; Fill and Steiner [2003]). Of particular interest,

Sangal [1983] utilizes a triangular hydrograph assumption, where the QIPF is the result of

geometric calculations that assume the area under the triangular hydrograph is equal to the area

under daily hydrograph (Figure S5-2 and Equation S5-2). Tan et al. [2007] combined these

approaches and found a strong linear relationship between estimated and actual QIPF in gages

where the triangular hydrograph assumption was appropriate.

, , 1 , 14

2

MDF n MDF n MDF n

IPF

Q Q QQ

Equation S5- 2

Figure S5-2. Illustration of the Sangal [1983] approximation. (Left) Plot of QMDF over the course

of three days. (Right) The same hydrograph where Equation S5-2 was applied to estimate QIPF.

Here, it is important to note Equation S5-2 is based on the assumption that the volumes represented

by V1 and V2 are equivalent.

For the purposes of this study, the Sangal [1983] triangular hydrograph approximation was used

in order to more accurately represent each gage’s historic flow record. At each gage, the

appropriateness of this correction was tested by developing a linear relationship between QIPF and

QMDF and testing for significance (p<0.05). Note, because annual maximum flows are recorded

within the USGS record, 20 pairs of QIPF and QMDF were available from 1986-2006 to develop this

relationship. As illustrated by data from the Colombia River (USGS Gage 12399500) in

Figure S5-3, most gages exhibit a strong linear relationship between QIPF and QMDF. However,

gages without a significant relationship between QIPF-QMDF were omitted form the analysis. Then,

the triangular hydrograph correction was applied to all days where QMDF was less than QBKF*. This

Page 99: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

90

allowed for the identification of floods where QMDF was less than QBKF*, thus providing a better

representation of river-floodplain connectivity in low-order streams.

Figure S5-3. Strong relationship experienced between QIPF and QMDF at the Michigan River near

the Fort Collins, CO (USGS Gage 12399500, catchment area 4 km). Of the 6,800 gages tested,

only 100 produced QIPF -QMDF relationships that were not statistically significant (p<0.05).

S5.2.3 Characterization of river-floodplain connectivity

Figure S5-4. Annual hydrograph from Boulder Creek near Boulder, CO across the 1997-1998 water

year. As delineated by QBKF*, the number of floods (nF) would be equal to 1 for this water year, the

duration of that storm (tF) would be 14 days, and the total flood volume (VT) would approximately 41.5

million m3.

Annual hydrograph characterization

To characterize river-floodplain connectivity across USGS gaging sites, measures of flood

duration (tF), annual number of events (nF), and total flood volume (VT) were estimated for

individual overbank floods across each gage’s flow record from 1986 to 2006. As illustrated by

Page 100: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

91

Figure S5-4, discrete overbank floods were identified using QBKF* as a threshold for river-

floodplain connectivity. For each identified overbank flood, values of tF and VT represent the time

stream flow was greater than QBKF* and the volume of stream flow during that time as defined by

Equation S5-3, respectively.

  ,  T FQdtV dt t Equation S5- 3

Floodplain and channel volume characterization

To better characterize river-floodplain connectivity and the potential for biogeochemical

processing at individual gages, we estimated the volume of transient storage in the floodplain, or

volume of floodplain water (VFP), and the volume of channel flow (VC) across each flood’s

hydrograph. The estimation of VFP and VC are based on a combination of estimated bankfull width

(wBKF), bankfull velocity (UBKF), and the application of the divided channel method (DCM).

To estimate the bankfull width (wBKF) at each gage, we took advantage of empirical relationships

relating wBKF to each gage’s catchment area (AWS). Originally applied to natural channels by

Leopold and Maddock [1953], hydraulic geometry measures such as bankfull width (wBKF), mean

bankfull depth (yBKF), and suspended sediment load can often be predicted using power-law

relationships where QBKF or contributing catchment area (AWS) are independent variables [Dunne

and Leopold, 1978; Hey and Thorne, 1986; Wilkerson et al., 2014]. Of particular interest, wBKF is

often utilized as a dependent variable because it can be measured remotely (e.g. Allen and Pavelsky

[2015]) and is less sensitive to measurement error when compared to other hydraulic geometry

measures such as bankfull stage or cross-sectional area [He and Wilkerson, 2011; Wahl, 1977].

For the purposes of this study, wBKF was estimated using relationships developed by Wilkerson et

al. [2014] (Equation S5-4). These relationships are ideal because they are based on a large data set

(n=1018) that were developed from sites at or near USGS gaging stations across the US.

0.191

0.462

2.18 , ln( ) 1.6

1.41 , ln 1.6  B

WS WS

WS WS

KF

A Aw

A A

Equation S5- 4

To estimate UBKF, we utilized empirical relationships where channel parameters associated with

energy regulation (e.g. channel slope, SCH, and meander wavelength, λ) serve as independent

variables. Unlike other hydraulic geometry measures, AWS is not a strong predictor of UBKF [Singh,

2003]. This is because downstream velocity tends to be constant, or gradually increase, as AWS

increases [Leopold, 1953]. For use in remote sensing applications, Bjerklie [2007] derived two

relationships estimating UBKF. The first highlights the connection between meanders and flow

resistance, and similar to Manning’s Equation, estimates UBKF as a function of water surface slope,

modified meander wavelength, and a fitting factor “m” that represents an arbitrary fraction of the

meander length. The derivation of the theoretical equation provides justification for the second

model, an empirical model that relates UBKF with SCH and λ (Equation S5-5). The model is based

on geomorphic surveys and flume studies completed by Church and Rood [1983] and Leopold et

al. [1960], respectively. Bjerklie [2007] goes on to suggest that in the absence of remotely sensed

data, wavelength could be estimated using relationships presented in Leopold and Wolman [1964]

and, more preferable, Williams [1986] (Equation S5-6).

Page 101: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

92

0.31 0.321.37 CHBKFU S Equation S5- 5

1.1210.2 BKFw Equation S5- 6

Figure S5-5. Estimate of UBKF at USGS gaging stations using relationship developed by

Bjerklie [2007].

Figure S5-5 displays estimates of UBKF from Equations S5-5 and S5-6. Because Equation S5-5 is

based on a relatively small sample size (n=78), it is less robust than the relationships presented in

Equation S5-4. However, the estimates of UBKF are fairly constant across stream order, which is

consistent with general hydrogeomorphic theory (e.g. Leopold, [1953]). Another alternative to

estimating UBKF would be to divide QBKF* by an estimate of AXS at bankful conditions derived from

continental scale hydraulic geometry relationships (e.g. Andreadis et al. [2013]). While this

approach could potentially provide an adequate estimate of UBKF, error in the estimate would be

compounded by the uncertainty associated with the calculation of QBKF*.

One dimensional flow in compound channels is often modeled with DCM, where flow is calculated

separately in the channel and floodplains compartments [Chow, 1959; Knight, 2006]. The classic

method presented by Chow [1959] does not account for the exchange of momentum or matter.

However, other methods (e.g. Bousmar and Zech [1999]; Knight and Demetriou [1983]; Myers

[1978]) have improved upon this and are extensively reviewed by Knight [2006]. While we are

not modeling one dimensional flow here, we are conceptualizing floods by separating the active

flood flows as occurring in both channel and floodplain compartments (Figure S5-6).

Page 102: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

93

Figure S5-6. (Bottom) Conceptualization of the Divided Channel Method (DCM) where the active

storm channel cross section is subdivided into floodplain and channel compartments. (Top) Cross

section displaying the subdivisions of the channel into bankful channel and bankful excess

compartments.

Flood flow is defined conceptually as the sum of channel flow (QC) and transient storage in the

floodplain. Integrating across a single overbank flood hydrograph, VT is the sum of channel flood

volume, VC, and floodplain flow volume, VFP:

 T C FPV V V Equation S5- 7

To find VC, it is necessary to further subdivide the channel compartment into the bankful channel

and bankful excess compartments (Figure S5-6). VC is the sum of total volume of flow in the

bankful channel (VBKF) and a bankfull excess compartment that remains associated with the

channel (VBFE).

   C BKF BFEV V V

Equation S5- 8 Here, VBKF is calculated by integrating the estimated bankfull flow (QBKF*) over the course of

the storm using:

* BKF BKF

V Q dt Equation S5- 9

Similarly, VBKE is calculated by integrating UBKF across the bankfull excess cross sectional area

over the course of the individual storm:

Page 103: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

94

 BFE BKFV U dAdt∬

Equation S5- 10

The DCM method’s relatively simple geometry specifies the cross sectional area of the excess

bankfull compartment as equivalent to the product of the bankful width (wBKF) and the stage excess

of bankful (Δy). Equation S5-10 is further simplified by assuming UC is equivalent to UBKF:

BFE BKF BKFV U w y t Equation S5- 11

Once VT and VC are estimated, we calculated VFP by rearranging Equation S5-7 and simply

subtracting VC from VT.

Flood Volume Ratio and Storage Index characterization

Finally, two indices are calculated to characterize river-floodplain connectivity across each gage’s

flow record: the flood volume ratio (FVR) and the storage index (SI). The FVR provides an

understanding of each gages upstream hydrologic conditions and is the ratio of the sum of discrete

estimates of VT and total streamflow experienced across each gages flow record:

2006

1986

TVFVR

Qdt

Equation S5- 12

In contrast, the SI characterizes river-floodplain connectivity experienced during individual floods

and is the ratio of VFP and VT :

   FP

T

VSI

V

Equation S5- 13

To explore patterns associated with both physiography and contributing catchment area, we

summarized FVR and SI by physiographic divisions (R) and stream order (ω), as defined by

Fenneman and Johnson [1946] and Strahler [1957] (Tables S5-2 and S5-3). Note, because each

gage experiences a range of SI, the median SI from each gage was used to tabulate values across

the different R and ω subdivisions.

Table S5-2. Median Flood Volume Ratio by Stream Order (ω) and Physiographic Division (R)

Stream

Order

Appalachian

Highlands

Atlantic

Plain

Interior

Highlands

Interior

Plains

Intermontane

Plateaus

Laurentian

Upland

Pacific

Mountains

Rocky

Mountains

1 0.34 0.34 0.18 0.39 0.58 NA 0.56 0.64

2 0.30 0.40 0.22 0.31 0.69 0.26 0.54 0.69

3 0.31 0.38 0.28 0.36 0.47 0.48 0.51 0.66

4 0.36 0.45 0.27 0.34 0.53 0.39 0.50 0.64

5 0.42 0.39 0.24 0.34 0.46 0.34 0.57 0.55

6 0.57 0.63 0.26 0.40 0.48 0.41 0.58 0.54

7 0.48 0.59 0.60 0.52 0.52 NA 0.47 0.57

8 NA 0.72 NA 0.32 NA NA NA NA

Page 104: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

95

Table S5-3. Median Storage Index by Stream Order (ω) and Physiographic Division (R)

Stream

Order

Appalachian

Highlands

Atlantic

Plain

Interior

Highlands

Interior

Plains

Intermontane

Plateaus

Laurentian

Upland

Pacific

Mountains

Rocky

Mountains

1 0.26 0.20 0.28 0.23 0.18 NA 0.17 0.08

2 0.22 0.20 0.13 0.20 0.13 0.15 0.23 0.17

3 0.27 0.24 0.21 0.18 0.12 0.13 0.26 0.12

4 0.25 0.23 0.16 0.20 0.16 0.17 0.22 0.12

5 0.24 0.22 0.26 0.19 0.16 0.19 0.21 0.13

6 0.27 0.24 0.25 0.20 0.15 0.11 0.20 0.12

7 0.24 0.19 0.14 0.21 0.15 NA 0.16 0.08

8 NA 0.13 NA 0.19 NA NA NA NA

Then, we used the Spearmen Correlation (e.g. Spearman [1904]) to test if there was a significant

effect of stream order on median FVR and median SI, respectively, for each physiographic

division. We did not find any significant correlations between stream order and SI, but did find

significant correlations between stream order and FVR for a subset of the physiographic divisions

(Table S5-4).

Table S5-4. Spearman Correlation (RS) between FVR

and Stream Order by Physiographic Division

Physiographic Division RS p-value

Appalachian Highlands 0.83 0.02**

Atlantic Plain 0.90 <0.01**

Interior Highlands 0.72 0.07*

Interior Plain 0.27 0.52

Intermontane Plateaus -0.56 0.19

Laurentian Upland 0.31 0.61

Pacific Mountains -0.29 0.53

Rocky Mountains -0.81 0.03**

S5.3 Application to the National Stream Network We then applied the metrics obtained from the USGS gaging stations to the US stream network in

order to gain perspective on continental scale patterns in river-floodplain connectivity. While the

USGS gaging station coverage is extensive, individual gages only serve as point measurements,

and analysis of river-floodplain connectivity across the entire network is needed to fully

understand spatial variations associated with both physiography and catchment area. Thus, gages

were subdivided into their respective physiographic division (R) and stream order (ω), as defined

by Fenneman and Johnson [1946] and Strahler [1957]. We tabulated median values for floodplain

volume (VFP) and annual number of events (nf) for each subdivision, and then we estimated annual

floodplain transient storage (SR,ω):

,R FP fS V n Equation S5-14

Page 105: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

96

The results for each subdivision can be found in Table S5-5. We then applied estimates of SR,ω

from each subdivision to flow lines of matching properties within the NHDplus flow network. For

example, the median d for gages along 4th Order streams in the Atlantic Plain was 2 days. All flow

lines of 4th order streams in the Atlantic plain were then assigned d = 2 days. Note, because there

was a lack of USGS gaging stations along 8th order streams in most physiographic divisions, we

utilized median values from all gages along 8th order streams and applied them to 8th order

subdivisions.

Table S5-5. Estimated annual floodplain transient storage (106 m3) by Stream Order (ω) and

Physiographic Division (R)

Strea

m

Order

Appalachi

an

Highlands

Atlant

ic

Plain

Interior

Highlan

ds

Interi

or

Plains

Intermonta

ne

Plateaus

Laurenti

an

Upland

Pacific

Mountai

ns

Rocky

Mountai

ns

1 0.01 0.04 1.19 0.04 0.02 NA 0.02 0.01

2 0.05 0.08 0.02 0.05 0.01 0.01 0.04 0.03

3 0.18 0.22 0.28 0.14 0.03 0.10 0.25 0.06

4 0.63 1.11 0.82 0.57 0.11 0.47 0.77 0.13

5 2.17 2.97 2.82 1.61 0.33 2.25 2.82 0.43

6 8.60 14.37 3.36 3.67 1.40 2.28 2.81 1.11

7 23.24 31.67 7.40 17.83 1.05 NA 41.81 1.42

8 55.23 55.23 55.23 55.23 55.23 55.23 55.23 55.23

To summarize river-floodplain connectivity across hydrogeomorphic gradients (i.e. stream order),

we estimated cumulative floodplain transient storage (Sω) and length weighted mean floodplain

transient storage

( S ) using SR,ω values from the entire NHDplus network (Table S5-6):

Table S5-6. Cumulative floodplain transient storage (Sω) and length weighted mean floodplain

transient storage ( S ) across stream order from the US stream network.

Stream Order

(ω)

Cumulative Floodplain

Transient storage

(1012 m3)

Length Weighted Mean

Floodplain Transient Storage

(106 m3)

1 0.73 0.52

2 0.25 0.40

3 0.40 1.19

4 0.79 4.22

5 1.23 11.5

6 2.52 44.5

7 3.44 134

8 1.09 110

Page 106: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

97

,RS S

Equation S5- 15

,R reach

reach

S l

Sl

Equation S5- 16

Finally, we summarized river-floodplain connectivity across physiographic provinces by

calculating the cumulative length normalized floodplain transient storage (SR) values across each

physiographic province (Table S5-7):

, /R R reach

R

S S l Equation S5- 17

Note, here we used physiographic province, as opposed to physiographic division, to examine

regional differences in river-floodplain connectivity. As a subset of physiographic divisions, the

provinces provided a more appropriate resolution to differentiate variation in river-floodplain

connectivity across regions.

Table S5-7. Cumulative length normalized floodplain transient storage (SR) across

physiographic provinces

Physiographic Province Cumulative Floodplain transient storage

(x109 m3/km)

Coastal Plain 9.89

New England 1.36

St. Lawrence Valley 0.08

Valley and Ridge 3.64

Adirondack 0.27

Appalachian Plateaus 2.73

Blue Ridge 0.53

Central Lowland 5.31

Piedmont 4.35

Superior Upland 0.22

Interior Low Plateaus 0.64

Ozark Plateaus 1.41

Ouachita 0.75

Great Plains 8.77

Middle Rocky Mountains 0.09

Southern Rocky Mountains 0.02

Wyoming Basin 0.08

Northern Rocky Mountains 0.19

Basin and Range 0.29

Colorado Plateaus 0.63

Cascade-Sierra Mountains 0.34

Columbia Plateau 0.23

Pacific Border 2.03

Lower Californian 0.01

Page 107: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

98

S5.4 Works Cited Alexander, R. B., R. A. Smith, and G. E. Schwarz (2000), Effect of stream channel size on the

delivery of nitrogen to the Gulf of Mexico, Nature, 403(6771), 758-761.

Allen, G. H., and T. M. Pavelsky (2015), Patterns of river width and surface area revealed by the

satellite-derived North American River Width (NARWidth) dataset, Geophys. Res. Lett.,

2014GL062764.

Andreadis, K. M., G. J. P. Schumann, and T. Pavelsky (2013), A simple global river bankfull

width and depth database, Water Resources Research, 49(10), 7164-7168.

Bacon, D. W., and D. G. Watts (1971), Estimating the Transition between Two Intersecting

Straight Lines, Biometrika, 58(3), 525-534.

Bieger, K., H. Rathjens, P. M. Allen, and J. G. Arnold (2015), Development and Evaluation of

Bankfull Hydraulic Geometry Relationships for the Physiographic Regions of the United States,

JAWRA Journal of the American Water Resources Association, n/a-n/a.

Bjerklie, D. M. (2007), Estimating the bankfull velocity and discharge for rivers using remotely

sensed river morphology information, J. Hydrol., 341(3–4), 144-155.

Bousmar, D., and Y. Zech (1999), Momentum transfer for practical flow computation in

compound channels, Journal of hydraulic engineering, 125(7), 696-706.

Chow, V. T. (1959), Open-channel hydraulics, McGraw-Hill.

Church, M., and K. Rood (1983), Catalogue of alluvial river channel regime data, first ed. ,

Department of Geography, University of British Colombia, Vancouver, Canada.

Domeneghetti, A., A. Castellarin, and A. Brath (2012), Assessing rating-curve uncertainty and

its effects on hydraulic model calibration, Hydrology and Earth System Sciences, 16(4), 1191-

1202.

Dunne, T., and L. Leopold (1978), Water in environmental planning, New York, 566-580.

Ensign, S. H., and M. W. Doyle (2006), Nutrient spiraling in streams and river networks, Journal

of Geophysical Research: Biogeosciences, 111(G4), G04009.

Falcone, J. A., D. M. Carlisle, D. M. Wolock, and M. R. Meador (2010), GAGES: A stream gage

database for evaluating natural and altered flow conditions in the conterminous United States,

Ecology, 91(2), 621-621.

Fenneman, N. M., and D. W. Johnson (1946), Physiographic divisions of the conterminous,

edited by USGS, Reston, VA.

Page 108: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

99

Fill, H. D., and A. A. Steiner (2003), Estimating instantaneous peak flow from mean daily flow

data, J Hydrol Eng, 8(6), 365-369.

Fuller, W. E. (1914), Flood flows., Trans. Am. Soc. Civ. Eng. , 77, 564-617.

Gomez-Velez, J. D., and J. W. Harvey (2014), A hydrogeomorphic river network model predicts

where and why hyporheic exchange is important in large basins, Geophys. Res. Lett., 41(18),

2014GL061099.

He, L., and G. V. Wilkerson (2011), Improved Bankfull Channel Geometry Prediction Using

Two-Year Return-Period Discharge1, JAWRA Journal of the American Water Resources

Association, 47(6), 1298-1316.

Hey, R., and C. Thorne (1986), Stable Channels with Mobile Gravel Beds, Journal of Hydraulic

Engineering, 112(8), 671-689.

Horizon Systems (2014), NHDplus Dataset, accessed [December, 2014], at URL

[http://www.horizon-systems.com/nhdplus/]

Hughes, D. A. (1980), Floodplain Inundation - Processes And Relationships With Channel

Discharge, Earth Surf. Process. Landf., 5(3), 297-304.

Hupp, C. R. (2000), Hydrology, geomorphology and vegetation of Coastal Plain rivers in the

south-eastern USA, Hydrological Processes, 14(16-17), 2991-3010.

Jones, C. N., D. T. Scott, B. L. Edwards, and R. F. Keim (2014), Perirheic mixing and

biogeochemical processing in flow-through and backwater floodplain wetlands, Water Resources

Research, 50(9), 7394-7405.

Kennedy, E. J. (1984), Discharge ratings at gaging stations: U.S. Geologial Survey Techniques

of Water-Resources Investigations, book 3, chap A10, edited.

Knight, D. (2006), River flood hydraulics: theoretical issues and stage-discharge relationships,

River basin modelling for flood risk mitigation, 17, 301-334.

Knight, D., and J. Demetriou (1983), Flood Plain and Main Channel Flow Interaction, Journal of

Hydraulic Engineering, 109(8), 1073-1092.

Langbein, W. (1944), Peak discharge from daily records, Water Resour. Bull, 145.

Leopold, L. B., and T. Maddock (1953), The hydraulic geometry of stream channels and some

physiographic implications.

Leopold, L. B. (1953), Downstream change of velocity in rivers, American Journal of Science,

251(8), 606-624.

Page 109: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

100

Leopold, L. B., and M. G. Wolman (1964), Fluvial processes in geomorphology, edited, San

Francisco: WH Freeman.

Leopold, L. B., R. A. Bagnold, M. G. Wolman, and L. Brush (1960), Flow resistance in sinuous

or irregular channels, paper presented at The Physics of Sediment Transport by Wind and Water,

ASCE.

Mallakpour, I., and G. Villarini (2015), The changing nature of flooding across the central

United States, Nature Clim. Change, 5(3), 250-254.

McKay, L., T. Bondelid, T. Dwald, J. Johnson, R. Moore, and A. Rea (2012), NHDPlus Version

2 User Guide.

Muggeo, V. M. R. (2003), Estimating regression models with unknown break-points, Stat Med,

22(19), 3055-3071.

Myers, W. (1978), Momentum transfer in a compound channel, Journal of Hydraulic Research,

16(2), 139-150.

NRCS. (2012), National Engineering Handbook: Part 630—Hydrology, USDA Soil Conservation

Service: Washington, DC, USA.

Pringle, C. (2003), What is hydrologic connectivity and why is it ecologically important?,

Hydrological Processes, 17(13), 2685-2689.

Pierson, M. S., B. J. Rosenbaum, L. D. McKay, and T. G. Dewald (2008), Strahler Stream Order

and Strahler Calculator Values in NHDPlusRep.

R Development Core Team (2012), R: A Language and Enviroment for Statistical Computing, R

Found For Stat. Compput, Vienna. [Available at http://ww.R-project.org/]

Rantz, S. (1982), Measurement and computation of streamflow; Volume 1, measurement of stage

and discharge; Volume 2, computation of discharge, US Geological Survey water-supply paper,

2175, 631.

Sangal, B. P. (1983), Practical Method of Estimating Peak Flow, J Hydraul Eng-Asce, 109(4),

549-563.

Singh, V. P. (2003), On the theories of hydraulic geometry, International Journal of Sediment

Research, 18(3), 196-218.

Taguas, E. V., J. L. Ayuso, A. Pena, Y. Yuan, M. C. Sanchez, J. V. Giraldez, and R. Pérez

(2008), Testing the relationship between instantaneous peak flow and mean daily flow in a

Mediterranean Area Southeast Spain, CATENA, 75(2), 129-137.

Page 110: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

101

Tan, K.-S., F. H. S. Chiew, and R. B. Grayson (2007), A steepness index unit volume flood

hydrograph approach for sub-daily flow disaggregation, Hydrological Processes, 21(20), 2807-

2816.

Tockner, K., and J. A. Stanford (2002), Riverine flood plains: present state and future trends,

Environmental Conservation, 29(3), 308-330.

USGS (2014), National Water Information System data available on the World Wide Web

(Water Data for the Nation), accessed [December, 2014], at URL [waterdata.usgs.gov/nwis.]

Wahl, K. L. (1977), Accuracy of channel measurements and implications in estimating

streamflow characteristics, Journal of Research of the Us Geological Survey, 5(6), 811-814.

Wilkerson, G. V., D. R. Kandel, L. A. Perg, W. E. Dietrich, P. R. Wilcock, and M. R. Whiles

(2014), Continental-scale relationship between bankfull width and drainage area for single-

thread alluvial channels, Water Resources Research, 50(2), 919-936.

Williams, G. P. (1978), Bank-full discharge of rivers, Water Resources Research, 14(6), 1141-

1154.

Williams, G. P. (1986), River meanders and channel size, J. Hydrol., 88(1–2), 147-164.

Wolman, M. G. (1955), The natural channel of Brandywine creek, Pennsylvania, United States

Geological Survey Professional Paper, 271, 56.

Page 111: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

102

CHAPTER 6. CONCLUDING REMARKS AND QUESTIONS

FOR FUTURE STUDY

In the presented research, floodwater residence time emerged as a dominant variable controlling

the fate and transport of solutes within floodplain systems (Figure 6-1). At the extremes, short

residence times restrict biogeochemical processes, whereas prolonged residence times lead to

solute limitation. In both of these cases, dominant biogeochemical process shift from removal and

retention to export (i.e., flushing) and accumulation of reactive solutes, respectively. Intrinsically,

river networks experience this gradient in processes (e.g. flushing-retention-accumulation) as the

network transitions from relatively short, sporadic floods in headwater catchments to longer,

seasonal flood pulses in large-river systems. However, there are confounding factors at the site

level that also control distributions of floodwater residence times and thus the dominant

biogeochemical process.

Figure 6-1. Conceptual model of large scale controls on net solute removal in floodplains. After

James et al., [2008].

In the StREAM Lab floodplain, representative of the headwater stream endmember of river-

floodplain connectivity, we measured relatively high rates of nitrate uptake in the spring and late

summer. However, because inundation period is so short (<4 hrs/yr), annual load reductions were

minimal at the reach scale. Further, because of the legacy of agriculture landuse at this site, the

experimental floodplain was a net source of reactive solutes, suggesting flushing was a dominant

biochemical mechanism during inundation events. While our findings do not diminish the value

of small floodplains’ ability to process upslope water, or the cumulative effects of small

floodplains across large-river basins, our results did suggest that the restoration industry should

focus on enhancing river-floodplain connectivity in medium to large streams when optimizing

restoration projects for solute removal.

Page 112: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

103

On the other end of the spectrum, we observed a dominance of both solute accumulation and

retention in the Atchafalaya River floodplain. At hydrologically isolated locations, there were

increased concentrations of reactive solutes suggesting accumulation, whereas in more connected

areas, reduced solute concentrations suggested enhanced retention and removal. Here, we observed

increased river-floodplain connectivity in reference to flow-through conditions (as compared to

backwater conditions) and preferential flow of floodwaters in swamp channels. Our results, along

with a handful of other studies, suggest river-floodplain connectivity could be further optimized

for solute retention within the Atchafalaya River Basin.

When examining the cumulative effects of river-floodplain connectivity across larger-river basins,

it is clear that large-river floodplains have greater capacity to process floodwaters than smaller

headwater floodplains, both at the site level and when taken cumulatively across large river basins.

This suggest that ecological engineering efforts should focus on enhancing river-floodplain

connectivity in medium to large-rivers when optimizing for nutrient removal. However, there is

uncertainty in how to move forward and implement optimization strategies. Below, we present

three potential areas of study that build off of the research presented in this dissertation:

Characterization of river-floodplain exchange: At this point, there are relatively few studies

that characterize river-floodplain exchange at the reach scale (e.g. Rudorff et al., 2014; Helton et

al., 2014). However, developing an accurate water balance for individual floodplains is critical in

understanding floodwater residence times and the associated biogeochemical processes (e.g.

Hughes, 1980). Largely, this has not occurred due to computational limitations associated with

2D hydraulic models and the difficulty of measuring floodplain microtopography and diffuse flow

associated with overbank flooding. However, recent advances in both computing power and

LiDAR should encourage new modeling efforts to simulate floodplain inundation and flow.

Ecological Engineering in the Atchafalaya River Basin: Currently, several studies have been

completed on river-floodplain connectivity within the Atchafalaya River Basin, and loosely, the

consensus is that nitrogen removal could be increased through enhanced river-floodplain

connectivity. I believe the next logical progression is to model maximum load reduction,

experimentally increase river-floodplain connectivity in collaboration with the Corps of Engineers,

and further characterize organic matter fate and transport.

Inclusion of floodplains in continental scale network modeling efforts: We are currently on the

cusp of a new understanding of solute fate and transport within large river basins (Creed et al.,

2015). However, contemporary analyses typically focus on baseflow or median flow conditions

(e.g. Gomez-Velez and Harvey, 2014; Kiel and Cardenas, 2014; Woodward et al., 2012), while

the majority of the solute balance often occurs during flood events (e.g. Royer et al., 2006) when

rivers are typically connected to their adjacent floodplains. Therefore, it is critical that floodplains

be included in any large-scale understanding of solute flux in riverine systems.

“We shall never achieve harmony with the land, anymore than we shall achieve absolute justice

or liberty for people. In these higher aspirations the important thing is not to achieve but to strive.”

― Aldo Leopold, Round River: From the Journals of Aldo Leopold

Page 113: Floodplain Hydrology and Biogeochemistry › bitstream › handle › 10919 › 75169 › Jo… · Floodplain Hydrology and Biogeochemistry Charles N. Jones Dissertation submitted

104

6.1 Works Cited Creed, I. F., et al. (2015), The river as a chemostat: fresh perspectives on dissolved organic

matter flowing down the river continuum, Can. J. Fish. Aquat. Sci.

Gomez-Velez, J. D., and J. W. Harvey (2014), A hydrogeomorphic river network model predicts

where and why hyporheic exchange is important in large basins, Geophys. Res. Lett., 41(18),

2014GL061099.

Helton, A. M., G. C. Poole, R. A. Payn, C. Izurieta, and J. A. Stanford (2014), Relative

influences of the river channel, floodplain surface, and alluvial aquifer on simulated hydrologic

residence time in a montane river floodplain, Geomorphology, 205, 17-26.

Hughes, D. A. (1980), Floodplain inundation - processes and relationships with channel

discharge, Earth Surf. Process. Landf., 5(3), 297-304.

James, W. F., W. B. Richardson, and D. M. Soballe (2008), Effects of residence time on summer

nitrate uptake in mississippi river flow‐ regulated backwaters, River Research and Applications,

24(9), 1206-1217.

Royer, T. V., M. B. David, and L. E. Gentry (2006), Timing of Riverine Export of Nitrate and

Phosphorus from Agricultural Watersheds in Illinois: Implications for Reducing Nutrient

Loading to the Mississippi River, Environ. Sci. Technol. 40(13). 4126-4131.

Rudorff, C. M., J. M. Melack, and P. D. Bates (2014), Flooding dynamics on the lower Amazon

floodplain: 1. Hydraulic controls on water elevation, inundation extent, and river‐floodplain

discharge, Water Resources Research, 50(1), 619-634.

Woodward, G., et al. (2012), Continental-Scale Effects of Nutrient Pollution on Stream

Ecosystem Functioning, Science, 336(6087), 1438-1440.


Top Related